Next Article in Journal
The Adsorption Mechanism of Hydrogen on FeO Crystal Surfaces: A Density Functional Theory Study
Next Article in Special Issue
MOF-Derived Co Nanoparticles Catalyst Assisted by F- and N-Doped Carbon Quantum Dots for Oxygen Reduction
Previous Article in Journal
Extraction of Silicon-Containing Nanoparticles from an Agricultural Soil for Analysis by Single Particle Sector Field and Time-of-Flight Inductively Coupled Plasma Mass Spectrometry
Previous Article in Special Issue
Effect of Calcination Temperatures on Surface Properties of Spinel ZnAl2O4 Prepared via the Polymeric Citrate Complex Method—Catalytic Performance in Glycerolysis of Urea
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbon Capture Using Porous Silica Materials

by
Sumedha M. Amaraweera
1,
Chamila A. Gunathilake
2,3,*,
Oneesha H. P. Gunawardene
2,
Rohan S. Dassanayake
4,*,
Eun-Bum Cho
5,* and
Yanhai Du
3,*
1
Department of Manufacturing and Industrial Engineering, Faculty of Engineering, University of Peradeniya, Peradeniya 20400, Sri Lanka
2
Department of Chemical and Process Engineering, Faculty of Engineering, University of Peradeniya, Peradeniya 20400, Sri Lanka
3
Department of Applied Engineering & Technology, College of Aeronautics and Engineering, Kent State University, Kent, OH 44242, USA
4
Department of Biosystems Technology, Faculty of Technology, University of Sri Jayewardenepura, Homagama 10200, Sri Lanka
5
Department of Fine Chemistry, Seoul National University of Science and Technology, Seoul 01811, Republic of Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(14), 2050; https://doi.org/10.3390/nano13142050
Submission received: 2 June 2023 / Revised: 4 July 2023 / Accepted: 7 July 2023 / Published: 11 July 2023
(This article belongs to the Special Issue New Trends in Mesoporous Materials for Catalysis and Sensors)

Abstract

:
As the primary greenhouse gas, CO2 emission has noticeably increased over the past decades resulting in global warming and climate change. Surprisingly, anthropogenic activities have increased atmospheric CO2 by 50% in less than 200 years, causing more frequent and severe rainfall, snowstorms, flash floods, droughts, heat waves, and rising sea levels in recent times. Hence, reducing the excess CO2 in the atmosphere is imperative to keep the global average temperature rise below 2 °C. Among many CO2 mitigation approaches, CO2 capture using porous materials is considered one of the most promising technologies. Porous solid materials such as carbons, silica, zeolites, hollow fibers, and alumina have been widely investigated in CO2 capture technologies. Interestingly, porous silica-based materials have recently emerged as excellent candidates for CO2 capture technologies due to their unique properties, including high surface area, pore volume, easy surface functionalization, excellent thermal, and mechanical stability, and low cost. Therefore, this review comprehensively covers major CO2 capture processes and their pros and cons, selecting a suitable sorbent, use of liquid amines, and highlights the recent progress of various porous silica materials, including amine-functionalized silica, their reaction mechanisms and synthesis processes. Moreover, CO2 adsorption capacities, gas selectivity, reusability, current challenges, and future directions of porous silica materials have also been discussed.

1. Introduction

With the exponential growth of industrialization, global warming and climate change have become worldwide concerns and have attracted much attention in recent decades [1]. Furthermore, human activities have significantly contributed to the increased levels of CO2 in the atmosphere. For example, atmospheric CO2 measured at NOAA’s Mauna Loa Atmospheric Baseline Observatory peaked for 2021 at a monthly average of 419 parts per million (ppm), and it is reported as the highest level since accurate measurements began 63 years ago [2].
The increase in CO2 concentration leads to the rise in global temperature and sea levels, alternative of rainfall patterns, extinction of species, natural disasters such as severe weather events, ranging from flash floods, hurricanes, freezing winters, severe droughts, heat waves, urban smog, and cold streaks [3].
The main CO2 stationary emission sources are power plants, refineries, chemical and petrochemical, iron and steel, gas processing, and cement industries. More irreversible and adverse environmental impacts should be expected if atmospheric carbon dioxide continues to rise. Therefore, the international communities led by the United Nations reached a landmark global accord, the Paris Agreement, adopted by 196 nations in 2015 to address climate change and related issues. Moreover, countries around the globe made their “nationally determined contributions (NDCs)” of greenhouse gas reduction. Different approaches employed in different countries to reduce CO2 emissions are shown in Table 1. Table 1 also summarizes the major advantages and disadvantages of each approach.
Among these approaches, the CO2 capture and storage (CSS) can reduce CO2 emissions by 85–90% from large emission sources [4]. CCS includes different CO2 capture, separation, transport, storage technologies, and chemical conversion, which are discussed in detail below.

2. CO2 Capture

2.1. CO2 Capture Technologies

Capture and sequestration of CO2 (CCS) from aforementioned stationary emission sources has been identified as a paramount option for the issues of global warming and climate change. CCS includes four primary steps known as CO2 capture, compression, transport, and storage, therefore, developing an efficient and economically feasible technology for the capture and sequestration of CO2 produced by anthropogenic emissions is critically important. CO2 capture is the central part of the CCS technology process and gained around 70–80% of the total expensive. However, CSS methods can be classified as, for example, (i) Post-combustion (ii) Pre-combustion, and (iii) Oxy-fuel combustion (Oxygen-fired combustion) [5,6].
In post-combustion capture technology, it collects and separates the CO2 from the emission gases of a combustion system [7,8,9,10,11]. Firstly, flue gas (mainly consists of CO2, H2O, and N2) passes through denitrification and desulphurization treatments. As the next step, the flue gas is fed to an absorber which contains solvent. Herein, CO2 regeneration occurs. Then the CO2-rich absorbent is sent to a CO2-stripper unit to release the CO2 gas. Moreover, CO2-lean absorbent is sent back to the CO2-absorber unit [1]. Next, the captured CO2 is then compressed into supercritical fluid and then transported [1] as shown in Figure 1.
Pre-combustion capture is a technology where CO2 is captured before the combustion process and CO2 is generated as an intermediate co-product of conversion process [12]. The pre-combustion technologies are mainly used in power plants, production of fertilizers and natural gas [13,14].
In oxyfuel combustion, the carbon-based fuel consumes in re-circulated flue gas and oxygen (O2) stream. CSS capture technology is considered expensive due to the high cost of O2 separation and production. However, the capture and separation of CO2 are reasonably easy compared to other methods and is considered as an energy-saving method [15].
Among the currently available technologies, post-combustion capture has grabbed much attention because it can be easily accomplished, applicable for large scale-power plants, easily managed and required short time for CO2 capture compared to other available methods [1]. Post-combustion capture uses different methods for gas separation, and collects CO2 by adsorption/desorption, as shown in Table 2, including absorption [6,16], adsorption [6,17], membrane-based technologies [18,19], and cryogenics [20]. Table 2 also depicts the efficiency, advantages, and disadvantages of the different types of post-combustion capture technologies.
Absorption process mainly uses liquids to capture CO2. During adsorption, once CO2 is separated from the gas, the sorbent should be regenerated by using a stripper, heating, or depressurization. Moreover, this method is considered as the most established process for CO2 separation [21]. In general, adsorbents can be divided into two types, namely, chemical and physical adsorbents (see Table 2 for details).

2.2. Criteria for Selecting CO2 Sorbent Material

Certain economical and technical properties are required in order to select the best solid adsorbent candidate for a particular CO2 capture application. These criteria are listed and described below.
  • Adsorption capacity for CO2:
The equilibrium adsorption capacity of a sorbent material is represented by its equilibrium adsorption isotherm. The adsorption capacity is an important parameter when considering the cost. Moreover, which causes reduction in the sorbent quantity, and in the size of the adsorption column. However, to enhance the adsorption capacity of solid sorbents, functionalization has been carried out with existing monoethanolamine (MEA) [24]. The CO2 working capacity should be in the range of 2–4 mmol/g of the sorbent [25].
  • Selectivity for CO2:
The adsorption selectivity or selectivity of CO2 is explained as the sorption uptake ratio of a target gas species compared to another type (as example N2) contained in a gaseous mixture under given operation conditions. Therefore, it depends on the purity of the adsorbed gas in the effluent [21]. However, the purity of CO2 influences transportation and sequestration and, therefore, this criterion plays an important role in CO2 sequestration [24].
  • Adsorption and desorption kinetics:
It is necessary to have fast adsorption/desorption kinetics for CO2 and it controls the cycle time of a fixed-bed adsorption system. Fast kinetics results in a sharp CO2 breakthrough curve in which effluent CO2 concentration changes are measured as a function of time, while slow kinetics provides a distended breakthrough curve. However, both fast and slow adsorption and desorption kinetics impact on the amount of sorbent required. In functionalized solid sorbents, the overall kinetics of CO2 adsorption mainly depend on the functional groups present, as well as the mass transfer or diffusional resistance of the gas phase through the sorbent structures. The porous support structures of functionalized solid sorbents also can be tailored to minimize the diffusional resistance. The faster an adsorbent can adsorb CO2 and be desorbed, the less of it will be needed to capture a given volume of flue gas [24].
  • Mechanical strength of sorbent particles:
The sorbent must show the stable microstructure and morphological structure in adsorption and regeneration steps. Mainly disintegration of the sorbent particles occurs due to the high volumetric flow rate of flue gas, vibration, and temperature. Apart from that, this could also happen due to abrasion or crushing. Therefore, a sufficient mechanical strength of a sorbent particles is required to keep CO2 capture process cost-effective [24].
  • Chemical stability/tolerance towards impurities:
Solid CO2 capture sorbents such as amine-functionalized sorbents should be stable in an oxidizing environment of flue gas and should be resistant to common flue gas contaminants [24].
  • Regeneration of sorbents:
The regeneration of the sorbent is energy saving and is one of the most important parameters required for improving energy efficiency [26]. Regeneration can be achieved through the adjustment of the thermodynamics of the interaction between CO2 and the solid adsorbent [24]. Considering regeneration, physisorption is mostly favored over chemisorption since the latter involves high energy consumption for regeneration.
  • Sorbent costs:
The production cost is the main key point when considering industrial applications at reasonable gas selectivity and adsorption performance [24].

2.3. Liquid Amine for CO2 Capture

Development of solvents for CO2 chemical absorption is a major area of research [27]. The ideal solvent should have a high CO2 absorption capacity and react rapidly and reversibly with CO2 with minimal heat requirement. The solvent should exhibit the following properties such as stability in oxidative and thermal environment, low vapor pressure, toxicity, flammability, and reasonable production cost [27].
Recently, a most promising CO2 capture method with chemical absorption is by using liquid amine which can be divided mainly into two groups known as simple alkanolamines and sterically hindered amines [28]. Examples for simple alkanolamines are monoethanolamine (MEA), diethanolamine (DEA), and triethanolamine (TEA) [29,30]. Furthermore, alkanolamines are the most widely used sorbents for CO2 capture. The structures of alkanolamines include primary, secondary, ternary amines containing at least one hydroxyl (-OH) group and amine group-(N-R) as shown in Table 3.
However, these different amine classes have different reaction kinetics with CO2, CO2 absorption capacity and equilibria, stability, and corrosion [28]. Advantages and disadvantages among the alkanolamines are shown in Table 3. As shown in Equations (1) and (2) below, both primary and secondary amines react with CO2 to form a carbamate and protonated amine, consuming approximately two moles of amine per mole of CO2 according to the zwitterion mechanism [31]. According to Equation (3), tertiary amines react with CO2 gas molecules in the presence of H2O while forming bicarbonates.
C O 2 + 2 R 1 N H 2 R 1 N H 3 + + R 1 N H C O O
C O 2 + 2 R 1 R 2 N H 2 R 1 R 2 N H + + R 1 R 2 N C O O
C O 2 + 2 R 1 R 2 R 3 N + H 2 O R 1 R 2 N H + + H C O 3
(where R1, R2, and R3 are aryl/alkyl groups).
However, García-Abuín et al. [32] observed that MEA produced a mixture of carbamate and bicarbonate as the main reaction products during CO2 absorption. The reaction starts with the reversible reactions between MEA and CO2 to form carbamate at low CO2 loading, followed by the CO2 hydration to form HCO3/CO32− under high CO2 loading, and accompanied by the hydrolysis of carbamate. The reaction mechanism of CO2 capture into MEA solution with different CO2 loadings is shown in Figure 2.
Table 3. Comparison between different liquid amines [33,34,35,36,37,38].
Table 3. Comparison between different liquid amines [33,34,35,36,37,38].
CriteriaAlkanolaminesSterically Hindered Amines
PrimarySecondaryTertiary
ExamplesMonoethanolamine (MEA)Diethanolamine (DEA)N-methyldiethanolamine (MEDA)2-amino-2-methyl-1-
propanol (AMP)
StructureNanomaterials 13 02050 i001Nanomaterials 13 02050 i002Nanomaterials 13 02050 i003Nanomaterials 13 02050 i004
CO2 loading at 59.85 °C
(mol CO2/mol amine)
0.426
(MEA 30 wt%)
0.404
(DEA 30 wt%)
0.141
(TEA 30 wt%)
0.466
(AMP 30 wt%)
Regeneration efficiency (%) at 90 °C75.584.8995.09
Advantages
  • Inexpensive solvent
  • Reversible absorption
  • High selectively (between acid and other gases)
  • Reacts with CO2 more rapidly
  • Inexpensive solvent
  • Reversible absorption
  • High selectively (between acid and other gases)
  • Reacts with CO2 more rapidly
  • Inexpensive solvent
  • Reversible absorption
  • High selectively (between acid and other gases)
  • High CO2 absorption capacity
  • Requires low regeneration energy
  • High CO2 absorption capacity
  • Requires low regeneration energy
Disadvantages
  • Lower CO2 absorption capacity
  • Requires high regeneration energy
  • Oxidative degradation occurs in the presence of other gas components
  • Corrosive
  • High capital costs
  • Lower CO2 absorption capacity
  • Requires high regeneration energy
  • Oxidative degradation occurs in the presence of other gas components
  • Corrosive
  • High capital costs
  • Reaction rate with CO2 is low
  • Compared to MEA and DEA
  • Corrosive
  • High capital costs
  • Low reaction rate
According to Table 3, there are three categories of alkanolamines that show increased capital costs due to requirement of specialized and expensive materials for construction [28]. On the contrary, degradation of alkanolamine causes operational, and environmental problems including high amount of absorbent required, corrosion of equipment, and demanding of energy [24].
Among three different alkanolamines, MEA is commonly considered as a well-established solvent to separate CO2 because it can be regenerated easily [35]. On the other hand, Rinprasertmeechai et al. reported the order of CO2 absorption capacity of the different alkanolamines as MEA > DEA > TEA (see Table 3) [33]. Moreover, they have further showed the regeneration ability of the amines in the following order: MEEA > > DEA > MEA. According to Table 3, MEA exhibits high CO2 adsorption capacity as it reacts more rapidly with CO2 compared to MEDA by forming carbamates. However, MEDA shows high regeneration efficiency and requires lower energy [36]. Moreover, Wang et al. found that, when MEA and MEDA are mixed with the appropriate ratio, the energy consumption for CO2 regeneration is reduced significantly [37].
Sterically hindered amines are based on primary or secondary amines with bulky alkyl groups, which is inhibited from reacting with CO2 through the effect of steric hindrance [28]. One example of sterically hindered amines is 2-amino-2-methyl-1-propanol (AMP). Steric factor reduces the stability of the formed carbamate due to the weak interaction between the CO2 molecule and the NH2 group, promoting fast hydrolysis to form bicarbonate and reducing regeneration energy. Due to the immediate regeneration process of AMP, the NH2 group can react with CO2 molecules over and over, increasing CO2 adsorption (see Table 3). Moreover, Dave et al. [38] compared the CO2 absorption of different liquid amine classes and showed a lower regeneration energy requirement for 30 wt% AMP over 30% MEA, 30% MEDA, 2.5% NH3, and 5% NH3 [38].
Recently, ionic liquids (IL) have also been investigated as liquid solvents for CO2 capture due to their low vapor pressure, thermal stability, non-toxicity, and adsorption capacity [39,40,41]. The widely studied ILs include bis(trifluoromethylsulfonyl)imide (TF2N), tetrafluoroborate (BF4), and hexafluorophosphate (PF6) [39,40,41]. However, the main drawbacks of the ILs are high viscosity and production high cost.

2.4. Comparison between Major Non-Carbonaceous Solid Sorbents for CO2 Capture and Importance of Silica Materials

Due to the low contact area between gas and liquid, low CO2 loading, and absorbent corrosion associated with liquid amine-based sorbents, solid sorbents for CO2 capture have attracted significant attention in recent years [42,43]. Various solid adsorbents have been proposed according to their structures and compositions, adsorption mechanisms, and regeneration process [43]. Many solid sorbents are cheap and readily available and show low heat capacities, fast adsorption kinetics, high CO2 adsorption capacities and selectivity, and high thermal, chemical, and mechanical stabilities [43].
Commercially available solid adsorbents for CO2 capture include carbonaceous materials such as activated carbons, nanofibrillated cellulose (CFCs), carbon nanotubes (CNTs), and non-carbonaceous materials, including silica, zeolites, hollow fibers, and alumina [6]. These materials show different surface morphologies, pore structures, specific surface areas, and functional groups.
Carbonaceous adsorbents are widely used for CO2 capture due to their relative abundance, low cost, renewability, and high thermal stability. However, the weak CO2 adsorption capacities of carbonaceous materials at 50–120 °C make it challenging to use in industrial CO2 capture [44]. Therefore, much research focus has been given to non-carbonaceous materials. Table 4 tabulates commonly tested non-carbonaceous solid adsorbents for CO2 capture and their advantages and setbacks.
As mentioned earlier, carbonaceous adsorbents such as activated carbon have been widely used for CO2 capture due to their wide availability, low cost, and high thermal stability. However, weak CO2 adsorption of carbonaceous materials in the range of 50–120 °C leads to high sensitivity in temperature and relatively low selectivity in operation [44]. Therefore, many research works have focused on non-carbonaceous materials such as mesoporous silica, and zeolites due to their advantages, as shown in Table 4.
Zeolites are aluminosilicates with ordered three-dimensional (3D) microporous structures with high crystallinity and surface area [44]. The adsorption efficiencies of zeolites are primarily affected by their size, charge density, and chemical composition of cations in their porous structures [37]. It has been reported that the CO2 adsorption of zeolites increases as the Si/Al ratio increases and is exchanged with alkali and alkaline-earth cations in the structure of zeolites [45]. However, zeolites present several drawbacks, such as relatively low CO2/N2 selectivity and high hydrophilicity [46]. Apart from the above, zeolites show reduced CO2 adsorption capacity when CO2/N2 mixtures contain moisture, and zeolites require high temperatures (>300 °C) for regeneration [47].
Recently, metal-organic frameworks (MOFs) have gained much attention owing to their unique properties, such as tunable pore structure and high surface area [48]. However, when exposed to gas mixtures, the MOFs show decreased adsorption capacities [46]. Moreover, previous reports indicate that MOFs are promising materials for CO2 capture in laboratory settings; however, further research is required to confirm their practical applicability [49]. Water vapor also negatively affects the application of these sorbents by competing and adsorbing them onto physisorbents, thus decreasing their CO2 adsorption capacity [50].
Ordered mesoporous silica materials are good candidates because of their high surface area, high pore volume, tunable pore size, and good thermal and mechanical stability. So far, mesoporous silica includes the families of MCM (Mobil Company Matter: M41S, Santa Barbara Amorphous type material (SBA-n), anionic surfactant-template mesoporous silica (AMS) [44]. However, the CO2 adsorption capacities of them observed at atmospheric pressure are not high. Therefore, many studies have been recently reported on the functionalized mesoporous and nanoporous silica for efficient CO2 capture [51,52].
Several reviews have recently focused on the potential applications of porous silica materials as CO2 adsorbents. Reddy et al. [53] reported CO2 adsorption based on porous materials of MOFs, clay-based adsorbents, porous carbon-based materials, and polymer-based adsorbents. Liu et al. [54] also discussed different porous materials, including silica, for post-combustion CO2 capture [54]. However, more information on silica-based sorbents and their synthesis methods still needs to be available. Therefore, this review mainly discusses CO2 capture onto different porous and functionalized silica materials. In addition, an overview of synthesis processes and a comparison between the adsorption capacities are also profoundly discussed. Finally, the technical challenges and the future research directions of the porous silica materials for CO2 adsorption are also presented in this review.

3. CO2 Capture Methods

Two general mechanisms are involved in CO2 capturing using solid sorbents: chemisorption and physisorption. Table 5 represents the major comparison between chemisorption and physisorption. However, the two mechanisms differ in the interactions between the gas molecules and the sorbent surface. During chemisorption, gas molecules are chemically bonded to the surface, whereas in physisorption, there is no chemical binding of the gas molecules to the surface, see Figure 3.
CO2 capturing using solid adsorbent is a selective separation [24]. The critical parameters for solid sorbents are surface tension, pore size, temperature, and pressure [24,59]. The adsorption process involves repeated cycles of adsorption and desorption, also known as regeneration. The four main adsorption processes are: (i) Pressure Swing Adsorption (PSA), (ii) Temperature Swing Adsorption (TSA), (iii) Electric Swing Adsorption (ESA), and (iv) Vacuum Swing Adsorption (VSA). Figure 4 shows the four different adsorption processes and their unique characteristics.
In the PSA process, adsorption happens at low pressure, and desorption occurs at high pressure. The adsorption of the TSA process occurs in the temperature range of 40–120 °C and the desorption process in the temperature range of 120–360 °C, respectively [3]. The VSA process involves CO2 uptake at ambient pressure, then swings to a vacuum condition to regenerate the adsorbent. The ESA process conducts the adsorption–desorption process by changing the electrical supply [3]. Activated carbons, MOF, zeolites, activated alumina, and silica gel are mainly used sorbents in TSA and PSA processes, while ESA is considered less costly compared to those of both TSA and VSA [59].
The microwave-swing adsorption (MWSA) is another adsorption process that has recently received considerable attention due to its efficient energy management. Unlike in conventional heating, where solids heat through conduction and convection, the MWSA process can transfer energy directly to the adsorbate without transferring the heat to both the adsorbate and adsorbent [11,60].

4. CO2 Adsorption Using Mesoporous Silica Materials (Physisorbents)

4.1. Mesoporous Silica Materials

Mesoporous silica materials are used for various applications, including catalysis and wastewater treatment [61]. Mesoporous silica has unique properties such as uniformity of pore distribution (with size between 0.7 and 50 nm), high surface area (around 1000 m2/g), and good thermal stability [62]. The first synthesized mesoporous silica material was M41S in the 1990s [63]. However, the development of surfactants and synthesis protocols have been able to prepare many types of mesoporous silicas such as MCM-41, SBA-15, SBA-16, FDU-2, MCM-50, and KIT-5 with a diverse range of pore geometries such as cubic, and hexagonal, and morphologies such as rods, spheres, and discs [64].
In 1990, Mobil Oil Corporation discovered molecular sieves of the M41S family consisting of silicate/aluminosilicate [65]. Typically, these materials are prepared via the sol-gel method. Three well-defined structural arrangements have been identified after studying the effect of surfactant concentration, and those are hexagonal (MCM-41), cubic (MCM-48), and lamellar (MCM-50) structures. Therefore, these materials (M41S family) exhibit mesoporous arrays with amorphous walls of about 10 Å (1 nm) [65]. Moreover, the structural ordering of these M41S family materials can be changed with increasing hydrothermal synthesis temperature and time [65]. These M41S molecular sieves are mainly applied in catalysis [66], adsorption [65], and controlled release of drugs [67]. The main advantage of this mesoporous silica is its unique chemical structure consisting of the high density of functional silanol groups (Si–OH), pore size and shape can be molded during the synthesis process, and the internal surface can be easily modified with organic and inorganic groups [65,68,69].
Santa Barbara Amorphous family (SBA) first prepared silica-based materials with well-ordered mesoporous in 1998 [65]. This material group consists of SBA-2 (hexagonal close-packed array), SBA-12 (three-dimensional hexagonal network), SBA-14 (cubic structure), SBA-15 (two-dimensional hexagonal), and SBA-16 (structured in a cubic cage) [65,70]. These nanostructured mesoporous materials comprise a silica-based framework with uniform and well-ordered mesopores, large pores, thick and porous walls, high surface area, and high thermal stability [69,71]. The most widely investigated members of the SBA-n family in the literature are SBA-15 and SBA-16. The SBA-15- and SBA-16-based mesoporous arrays are commonly utilized as adsorbents [69], catalysts or catalytic [72], and drug deliveries [73].
The Fudan University synthesized mesoporous materials family (FDU-n)-based mesoporous silica arrays with well-ordered mesostructures and pore arrangements, high surface area, large and uniform distribution of pore diameter, amorphous pore-wall structures, and thermal and mechanical stability [74]. FDU-1-based mesoporous materials have a 3D face-centered cubic (FCC) structure with large cage-like mesopores, while FDU-2 mesoporous array possesses a mesostructured FCC unit cell and well-ordered 3D architecture [69].
On the contrary, the mesoporous material series of the KIT-n family, where n = 1, 5, or 6, are mainly represented by the KIT-1, KIT-5, and KIT-6. However, KIT-1-based mesoporous silicas exhibit a 3D architecture in a disordered framework with high surface area, large pore volume and pore diameter, and thermal and hydrothermal stability [75]. KIT-5-based nanostructured mesoporous materials have well-ordered 3D cage-like mesopores in a face-centered close-packed cubic lattice architecture [69]. In addition, KIT-6 shows 3D mesoporous amorphous walls with large pore size, uniform pore distribution, high surface area, and thermal stability [69].
Moreover, mesoporous silica materials of the M41S, SBA-n, FDU-n, and KIT-n families are used in a wide range of applications such as separation, catalysis, drug release adsorption, sensors, matrix solid-phase dispersion (MSPD) and solid-phase extraction [69].

4.2. Synthesis Procedures of Mesoporous Silica

Initially, Stöber et al. [76] discovered an effective method for synthesizing monodispersed silica particles. This process consists of hydrolysis of tetraethyl orthosilicate (TEOS) using ammonia as a catalyst in water and ethanol solution. This method leads to the synthesis of silica particles [77]. In this reaction, TEOS undergoes hydrolysis in an ethanol/ammonia solution. As a result, it produces silanol monomer (-Si-OH) with the epoxy groups (-Si-OEt), as shown in Equation (4). Then silanol groups undergo condensation to produce branched siloxane clusters, which causes to initiate the nucleation and growth of silica particles, see Equation (5). Simultaneously, silanol monomers react with the unhydrolyzed TEOS via condensation (see Equation (6)) and participate in the nucleation and growth of silica particles [30]. Moreover, the particle size of Stöber silica depends on the concentration of the aqueous ammonia solution and water in the ethanol reaction [30].
S i ( O E t ) 4 + X H 2 O H y d r o l y s i s S i O ( O E t ) 4 x O H x + X E t O H
S i O O E t 4 x O H x C o n d e n s a t i o n O E t 4 2 x ( O H ) 2 x 2 + H 2 O
S i O E t 4 + S i O O E t 4 x O H x C o n d e n s a t i o n O E t 7 x ( O H ) x 1 + E t O H
Many experimental factors control hydrolysis, silica condensation rate, assembly kinetics, nucleation, and growth rates [65,78]. The pH is an essential factor that influences the charges of silica species. Rates of hydrolysis of silane and condensation of the siloxane bond depend strongly on the charge states. Hydrolysis of the Si–OR bond in silanes could be catalyzed by acid and base conditions, but its rate is prolonged near the neutral conditions [78].
Sakamoto et al. [79] prepared silica nanoparticles (NPs) via the evaporation and self-assembly of silicate and quaternarytrialkylmethylammonium as a surfactant. This study shows that the size of NPs depends on the ratio between the surfactant and silica precursor. Apart from that, Sihler et al. [80] used dye-stabilized emulsion to synthesize SiO2 NPs. Moreover, this synthesis method provides silica capsules and sub-particles with precise size control. Monodispersed colloidal silica NPs (diameter of 15–25 nm) were prepared by Murray et al. [81]. In this study, as the silica source, octadecyltrimethoxysilane (OTMS) was used.
Simple synthesis methods called soft and hard templating are also applied to increase the pore volume and loading capacity of prepared hollow mesoporous SiO2 [82]. Template synthesis of mesoporous materials typically enrolls in three steps: template preparation, template-directed synthesis of the target materials using sol-gel, precipitation, hydrothermal synthesis, and template removal [83,84].
The hard-templating method involves nano-casting using pre-synthesized mesoporous solids [85]. Hard templating is a facile synthesis method for fabricating porous materials with a stable porous structure. The structure replication is very straightforward [83]. This approach utilizes porous hard templates such as mesoporous silica. The pores of these templates are impregnated with a precursor compound for the desired product, which is then thermally converted into the product. The template is finally removed to yield the desired mesoporous material as a negative structural replica of the hard template [83]. However, the method is costly and time-consuming. Moreover, the mesoporous parameters, such as mesostructure and pore sizes, are difficult to change [84].
In contrast, soft templating methods use cationic and anionic surfactants or block copolymers as templates [78]. During the synthesis, surfactant or block copolymers are used as a soft template. Moreover, the increase in surfactant micelle concentration causes the formation of a large assembly or self-assembly of 3D mesoporous [30]. Different 3D micelle structures can be obtained by varying the solvent ratio between the aqueous and non-aqueous and adding co-solvents. Moreover, the silica source interacts with the structure-directing agent (SDA) without any phase separation. The interactions between ions or charged molecules are vital in forming well-defined porous nanostructures [85].
The soft templating method mainly depends on the self-assembly of the surfactant [83]. The process is based on the interactions between inorganics. The mesoporous structure of the final material is obtained after the removal of the pore-templating surfactant or block copolymers by low-temperature calcination (up to 600 °C) or by different washing techniques (extraction) [83]. Figure 5 represents the synthesis mechanism of mesoporous silica in the presence of a cationic surfactant. The synthesis process of mesoporous silica is carried out using TEOS as the silica source [30]. In this process, surfactant plays a significant role in defining the pore size and volume of silica [30]. Cationic surfactant forms micelle structures with water, which arranges the cationic “heads” of the surfactant molecules to the outer side. It resulted in the hydrophobic “tails” collected in the center of each micelle. As the next step, silica molecules cover the micelle surface. Finally, the surfactant is removed via calcination or extraction, and it results in porous silica [30,86,87].
Figure 6 shows the schematic diagram for synthesizing mesoporous silica using block copolymer. As can be seen from Figure 6, titania-incorporated organosilica-mesostructures (Ti-MO) are synthesized via condensation method using silica precursors ([3-(trimethoxysilyl) propyl] isocyanurate and tetraethylorthosilicate) and titanium precursor (titanium isopropoxide) in the presence of the triblock copolymer, Pluronic P123 [88]. This method consists of template removal using two independent steps (i) extraction with a 95% ethanol solution and (ii) calcination of the sample at 350 °C. This method improves the adsorption capacity and enhances the structural properties such as specific surface area, micro-porosity, and pore volume.
Figure 5. Mechanism for the synthesis of mesoporous silica in the presence of a cationic surfactant (Reprinted with permission from Kim et al. [89]).
Figure 5. Mechanism for the synthesis of mesoporous silica in the presence of a cationic surfactant (Reprinted with permission from Kim et al. [89]).
Nanomaterials 13 02050 g005
The synthesis of MCM-41 and SBA-15 is performed using cetrimoniumbromide (CTAB) and Pluronic P123 surfactant. The CTAB is an ionic surfactant and acts as stearidonic acid (SDA) and which causes the formation of a hexagonal array of mesostructured composites [12]. However, as the final step, surfactants are removed by heating in air at high temperatures or by solvent extraction to obtain MCM-41 and SBA-15 [30]. Wu et al. [79] and Hao et al. [88] reported a detailed description of the mechanism. Paneka and co-workers have reported the synthesis of MCM-41 from fly ash using a hydrothermal process. However, the synthesis of MCM-41 shows reduced BET surface area, increased pore volume, and pore size [89].
Recently, Singh and Polshettiwar [90] reported the synthesis of silica nano-sheets using ammonium hydroxide. They have developed a method to synthesize silica nano-sheets using lamellar micelles as soft templates in a water-cyclohexane solvent mixture. Zhang et al. [19] also reported the large-scale synthesis of mesoporous silica nanoparticles. Reported data show that various morphologies and particle sizes have been obtained during the synthesis. For synthesis process, the reaction occurred at atmospheric pressure with a sol–gel technique using CTAB as a template.

4.3. Importance of Micro-Porosity and CO2 Adsorption Capacity of Mesoporous Silica Materials

The textural properties, including surface area, pore diameter and volume of mesoporous materials, are usually measured by studying nitrogen adsorption–desorption isotherms. The specific surface area is calculated using the volume adsorbed at different relative pressure data by the Brunauer–Emmett–Teller (BET) method [65]. Apart from that, the pore volume and pore size distribution are determined using the Barrett–Joyner–Halenda (BJH) method [65].
Furthermore, the textural properties are important parameters when considering CO2 adsorption using physisorbents. Moreover, microporosity plays a major role in CO2 gas adsorption because it involves the diffusion of CO2 molecules into the physisorbent [91,92,93]. Table 6 represents the textural properties and CO2 absorption capacity recorded for different ordered mesoporous silica materials studied.
MCM-41 has high porosity and an ordered hexagonal pore structure arrangement. However, it showed a low CO2 adsorption capacity of 0.63 mmol/g at 25 °C and 1 bar (see Table 6). This behavior may be due to the weak interactions between the hydroxyl groups of MCM-41 and CO2 molecules [93]. Son et al. prepared KIT-6, SBA-15, SBA-16, MCM-48, and MCM-41 and their textural properties of the materials are tabulated in Table 6 [94]. The pore size of mesoporous materials varied in the descending order of KIT-6 > SBA-15 > SBA-16 > MCM-48 > MCM-41. The KIT-6 exhibited the largest pore volume among the other sorbents. These combined features of large pore size and large pore volume would enable KIT-6 to better accommodate the bulky polyethyleneimine (PEI) with little hindrance, allowing higher loadings inside silica particles than other silica-supported materials. Moreover, Zelěnák and co-workers prepared three mesoporous silica materials with different pore sizes (3.3 nm MCM-41; 3.8 nm SBA-12; 7.1 nm SBA-15) [95]. During their studies, amine functionalization was investigated with the effect of pore size and architecture on CO2 sorption. According to the data, SBA-15 showed the highest CO2 adsorption of 1.5 mmol/g due to the highest amine surface density in SBA-15 [95].
Lashaki and Sayari [96] also investigated the impact of the support pore structure on the CO2 adsorption performance of SBA-15 silica. In this study, SBA-15 silica supports were used to obtain different pore sizes and intra-wall pore volumes. These materials were functionalized further with triamine through dry and wet grafting. CO2 sorption measurements showed the positive impact of support with large pore size and high intra-wall pore volume on adsorptive properties, with the former being dominant. Large pore volume influenced the load of more amine groups, CO2 uptakes, and CO2/N2 ratios and faster kinetics. When the intra-wall pore volume decreased by 53%, it caused a reduction in CO2 uptake capacity by up to 63% and CO2/N2 ratios by up to 62% and slower adsorption kinetics. Moreover, it was inferred that large pore size and high intra-wall pore volume of the support improved the adsorptive properties via enhanced amine accessibility [96].
Table 6. The textural properties and CO2 absorption capacity of various ordered mesoporous silica materials.
Table 6. The textural properties and CO2 absorption capacity of various ordered mesoporous silica materials.
Types of Mesoporous SilicaMesostructureSilica
Source
Surfactant/
Block Co-Polymer
BET Specific Surface Area (m2/g)Pore Volume (cm3/g)Pore Size (nm)Adsorption
Capacity
(mmol/g)
Adsorption ConditionsRef.
Temp. (°C)Pressure (Bar)
KIT-53D-cubic TEOSPluronic P1237111.058.040.48 301[97]
KIT-63D-cubicTEOSPluronic P123895 1.226.0---[94]
MCM-41Hexagonal Na2SiO3CTAB9941.003.030.63251[93]
Na2SiO3CTAB993 1.00 3.1 0.63 251[98]
Na2SiO3CTAB9800.924.08 [90]
MCM 48Cubic SiO2CTAB12871.13.5 251[99]
SBA-152D hexagonalTEOSP1231254 2.44 11.4---[100]
SBA-16Cubic cageTEOSPluronic F127736 0.754.1---[94]
SNS TEOSPluronic F1273940.1021.12.06251[101]
SNT TEOSPluronic F1273190.0726.02.46251[101]
Where CTAB: cetyltrimethylammoniumbromide and hexadecyltrimethylammoniumbromide, F127: tri-block copolymer F127, Na2SiO3: sodium silicate, P123: triblock copolymer (Pluronic P123), SiO2: silica, SNS: silica nano spheres, SNT: silica nano tube, TEOS: tetraethyl orthosilicate.

5. Chemisorbents (Amine Functionalized Si-Based Materials)—Application at Low and High Temperature CO2 Sorption

In physisorption, CO2 molecules attach to the pore walls through weak Van der Waals and pole–pole interactions [102]. However, the unmatched pore size of the mesoporous silica and the small diameter of the CO2 gas molecule causes low CO2 adsorption capacities. The heat of adsorption of the physisorption process ranges from −25 to −40 kJ/mol [103], which is approximately closer to the heat of sublimation [104]. Recently, it has been reported about mesoporous silica materials with improved CO2 sorption capacity with amine functionalization [105]. Hence, the adsorption capacity of CO2 depends on the nature of the amine groups and the spacing between the amino silanes [106]. Figure 7 represents the different types of amino silanes and polymer-containing amino groups used during the functionalization of mesoporous silica for enhanced adsorption or separation.

5.1. Synthesis of Amine-Functionalized Silica

Amine-based adsorbents are generally synthesized using three approaches: the selection of solid scaffolds with high amine loading ability, use of amines with high nitrogen content, and use of effective methods for introducing amine groups [44]. Synthesis methods of amine-functionalized silica materials include three main pathways: impregnation, grafting, and in-situ polymerization. Figure 8 shows the three different synthesis processes of amine-functionalized silica materials.
In impregnation, amines are physically trapped in the pores of silica materials. Moreover, the performance of amine-silica adsorbents is influenced by the pore structure of silica. For example, Chen et al. [107,108] reported that the CO2 adsorption capacity decreases as the pore diameter decreases. Moreover, surfactants, surface functional groups, amine types and heteroatom incorporation affect the impregnation process [54]. In this method, the amine loading is also influenced by the total pore volume of the silica materials and the amine density.
Moreover, if the amount of amine exceeds the capacity of the support, the amine species agglomerate on the support. The main advantage of this method is the simplicity and easy synthesis procedure. Further, many amine species can be incorporated with mesoporous silica due to the large pore volume of the porous silica materials [109].
Grafting occurs between an aminosilane and silica, as shown in Figure 8, where amine groups are grafted on the silica surface via covalent bonds [110]. Mainly, three methods are used for grafting amine onto silica support: post-synthesis grafting, direct synthesis by co-condensation (one-pot synthesis), and anionic template synthesis [111]. In a typical process, silica is dispersed in a solvent, amino silanes are added, and the mixture is heated under reflux. However, the amount of amine incorporated is related to the number of hydroxyl groups on the silica surface [109]. In-situ polymerization is another promising method for functionalizing porous silica, such as hyperbranched aminosilica (HAS). This category of supported sorbents can be considered a hybrid of grafting and impregnation [112].
Figure 8. Different types of synthesis processes of amine-functionalized silica materials (Schematic shows supported amines (yellow) in the pores (blue)) (Reprinted with permission from Bollini et al. [113]).
Figure 8. Different types of synthesis processes of amine-functionalized silica materials (Schematic shows supported amines (yellow) in the pores (blue)) (Reprinted with permission from Bollini et al. [113]).
Nanomaterials 13 02050 g008
Solvents, including toluene are also used for grafting. Moreover, the impregnating technique is widely employed because of its simplicity, low cost, environmental friendliness, and convenience for large-scale production [114]. However, to overcome the challenges caused by grafting, researchers have recently investigated aminosilane gas-phase grafting and supercritical fluid impregnation [115].
Supercritical fluid impregnation is one of the most effective, simple, and reproducible methods for producing homogeneous, covalently bonded, and high-density silane [115]. López-Aranguren et al. [115] synthesized functionalized silica via supercritical CO2 grafting of aminosilanes. This study used silica gels (4.1 and 8.8 nm pore diameter), mesoporous silica MCM-41 (3.8 nm pore diameter), and mono- and di-aminotrialkoxysilane.
The double-functionalization method of mesoporous materials is also widely used in recent years. Several studies prepared amine–silica composites using the double-functionalization method [116,117,118]. Those studies employed impregnation and grafting to improve CO2 uptake [116].

5.2. Comparison of Adsorption Capacities of Silica-Based Sorbents

Nigar et al. [99] synthesized the ordered mesoporous (MCM-48) silica with different silane molecules, including 3triethoxysilylpropylamine, 3-(2-aminoethylamino) propyl] trimethoxysilane and 2-[2-(3-trimethoxysilylpropylamino)ethylamino]ethylamine. Here-in, silane groups were covalently bound with the silica groups, as shown in Figure 9. The functionalization caused the reduction in the surface area and the pore volume compared to the non-functionalized MCM-48 (1287 m2/g and 1.1 cm2/g) (see Table 7). Most importantly, it is seen that the increment of the number of amine groups in silane molecules leads to a decrease in CO2 absorption capacity governed via chemisorption [99].
Moreover, Park et al. [29] synthesized functionalized silica using silane molecules, similar to the study conducted by Niger et al. [99]. However, they compared in-situ polymerization and grafting. According to the data (see Table 7), the sorbent prepared through in-situ polymerization shows enhanced CO2 adsorption capacity. Ahmed et al. [93] reported a detailed study about the functionalization of mesoporous Si-MCM-41 with different loadings of PEI. According to their work, with increasing PEI loading, the CO2 adsorption capacity also increased (see Table 7). They mentioned that the enhanced adsorption is due to branched PEI with many amino groups, providing potential sites for CO2 molecules. Moreover, the hierarchical mesoporous structure of Si-MCM-41 made these sites accessible to CO2 by improving the dispersion of PEI [119].
Gargiulo and co-workers investigated the effect of temperature on CO2 adsorption capacity on SBA-15 and PEI. CO2 adsorption was evaluated at 25, 40, 55, and 75 °C temperatures [120]. The experimental data showed a significant dependence of the CO2 adsorption capacity on temperature (Table 7). The effect of pore dimension on CO2 adsorption over amine-modified mesoporous silicas was reported by Heydari-Gorji et al. [100]. The pore lengths of the silica supports were 25, 1.5, and 0.2 μm. It showed that the small pore size of silica materials exhibited the highest adsorption capacities due to the enhanced amine accessibility inside the pores. Heydari-Gorji and Sayari [121] showed PEI impregnation for CO2 removal applications. They demonstrated that PEI-functionalized silica materials were thermally stable at mild temperatures. Kuwahara et al. [122] synthesized poly(ethyleneimine)/silica composite adsorbents by incorporating zirconium (Zr) into the silica support. The authors observed Zr sites with increased CO2 adsorbent capacity (see Table 7), regeneration, and stability.
Apart from that, Kishor and Ghoshal [123] investigated the effects of the structural parameters such as pore size, pore volume, and surface area of the silicas and amine-functionalized silica on the CO2 sorption capacity. The authors used various silica materials such as KIT-6, MCM-41, SBA-15, and HV-MCM-41. The wet impregnation method was employed to prepare the pentaethylenehexamine (PEHA) functionalized silica. The CO2 capture capacities of the amine-functionalized silicas were measured at 105 °C and 1 bar pressure conditions (see Table 7). The KIT-6 showed the highest CO2 capture capacity of 4.48 mmol/g of CO2 at 105 °C and 1 bar pressure) among all the sorbents investigated (MCM-41 < HVMCM-41 < SBA-15 < KIT-6). Furthermore, KIT-6 showed enhanced amine density distribution due to large pore volume. All the other silica sorbents remained stable up to ten adsorption–desorption cycles.
Table 7. CO2 adsorption capacities and structural properties of amine functionalized silica-based adsorbents.
Table 7. CO2 adsorption capacities and structural properties of amine functionalized silica-based adsorbents.
Silica-Based SorbentAmine TypesCO2 Adsorption
Performance
Capacity
(mmol/g)
ConditionsBET Specific Surface Area (m2/g)Pore Volume (cm3/g)Pore Size (nm)Preparation MethodsRef.
Temperature (°C)Pressure (Bar)
DWSNT-0.125 830.58 Immobilization[124]
DWSNTAPTMS1.025 1120.72 Immobilization[124]
DWSNTMAPTMS1.525 1140.79 Immobilization[124]
DWSNTDEAPTMS1.825 68.90.49 Immobilization[124]
DWSNTAEAPTMS2.2525 60.90.45 Immobilization[124]
HASAziridines3.2525 7150.15 [125]
HPSPEI2.44 75 10.5 0.009 Impregnation[126]
HVMCM-41PEHA4.07105 1 Impregnation[123]
KIT-6PEHA4.48105 1 Impregnation[123]
MCM-41EDA 1.1935 Impregnation[127]
MCM-41DETA1.4335 Impregnation[127]
MCM-41TEPA1.9635 Impregnation[127]
MCM-41PEHA2.3435 Impregnation[127]
MCM-41MEA (3%)11.3925 4260.423.12Impregnation[128]
MCM-41PEI0.3940 0.15443 0.340 2.95Impregnation[49]
MCM-41PEI0.2275 15901.413.6Impregnation[120]
MCM-41PEI
Aziridine
0.9875 1 In-situ grafted polymerization[129]
MCM-41APTS9425 1100.01 Grafting[114]
MCM-41APTS0.70 300.1 [130]
MCM-41APTS2.48201170.0420.1Grafting[131]
MCM-41PEHA4.5105 1 Impregnation[120]
MCM-41MEA0.8925 1190.82 Impregnation[98]
MCM-41DEA0.8025 1130.07 Impregnation[98]
MCM-41TEA0.6325 12130.17 Impregnation[98]
MCM-41Branched PEI 1.08100 160-Impregnation[93]
MCM-41Branched PEI 0.79100 1120.04-Impregnation[93]
MCM-41Branched PEI—(30 wt%)0.70100 1800.14-Impregnation[93]
MCM-41Branched PEI 28100 11040.122.05Impregnation[93]
MCM-41Branched PEI 17.5100 12910.172.05Impregnation[93]
MCM-41TEPA1.2425 1110.051.8Impregnation[132]
MCM-48APTES0.62 251.0110720.522.9Grafting[99]
MCM-48TRI0.46251.016980.392.6Grafting[99]
MCM-48TRI0.4425 1.014630.232.5Grafting[99]
MSiNTsPEI 2.7592 52.40.1712.4Impregnation[133]
OMSPEI1.4 25 3520.79 Grafting[120]
SAB-15PEHA4.0105 1 Impregnation[123]
SBA-15PEI0.6525 6831.198.5Impregnation[122]
SBA-15PEI/Zr41.3425 6421.088.6Impregnation[122]
SBA-15PEI/Zr71.5625 6741.239.5Impregnation[122]
SBA-15PEI/Zr141.4125 6010.697.0Impregnation[122]
SBA-15PEI/Ti1.40.2425 5100.394.4Impregnation[122]
SBA-15NH2OH1.6525 1435.60.546.85 Grafting[134]
SBA-15APTMS1.46250.15820.165Grafting[135]
SBA-15TEPA2.4570 50.03 Grafting[100]
SBA-15AMP1.7970 3720.21 Grafting[120]
SBA-15
(0.2 µm)
PEI5.8410015901.4413.6Impregnation[120]
SBA-15 (1.5 µm)PEI-10017460.807.2Impregnation[120]
SBA-15 (25 µm)PEI5.81100 15800.9510.5Impregnation[120]
SiO2APTES4.3 30 670.51 In-situ polymerization[29]
SiO2AEAPTMS5.730 450.37 In-situ polymerization[29]
SiO2TRI5.630 250.22 In-situ polymerization[29]
SiO2APTES0.530 2161.11 Grafting[29]
SiO2AEAPTMS0.330 2061.10 Grafting[29]
SiO2TRI0.830 1720.99 Grafting[29]
SMCM-41MEA 10.4025 4050.393.01Impregnation[128]
SBA-15TEPA4.5751121.1 0.327 Impregnation[136]
MPSMTEA4.27751340.089.5Impregnation[50]
MCM-41TRI1.74250.05678.31.47 Grafting[137]
MCM-41APTES1.203011045.212.5930Grafting[138]
MCM-41PEI0.983016.60.010.8Grafting[139]
MCM-41PEI4.684518941.285.1Grafting[116]
MCM-41PEI 2.92500.1508 0.98 2.54Impregnation[140]
MCM-41TEPA2.25500.14310.832.21Impregnation[140]
MCM-41-KOHPEI-3.38500.13911.082.33Impregnation[140]
MCM-41-Ca(OH)2PEI-3.81500.14111.122.50Impregnation[140]
MCM-41-CsOHPEI-5.02500.13060.912.14Impregnation[140]
MCM-41-KOHTEPA-3.93500.13220.972.15Impregnation[140]
MCM-41-Ca(OH)2TEPA-3.76500.14050.942.31Impregnation[140]
PET-CsOHTEPA-5.42500.12930.972.61Impregnation[140]
MCM 48PEI1.09800.2479.30.021.68Impregnation[141]
MCM-41PEI1.23800.2459.10.021.80Impregnation[141]
SBA-15PEI1.07800.2462.10.015.2Impregnation[141]
SBA-15PEI1.77017830.037.0Impregnation[142]
SBA-15PEI1.26450.153990.798.2Impregnation[143]
MCM 41PEI 3.53251240.012 Impregnation[144]
MCM 41APTS 2.412517360.37 Grafting[144]
SBA-15PEI1.84251.21950.397.0Grafting[145]
SBA-15-APES 1.78251.21900.377.2Grafting[145]
SBA-15-APESPEI1.54251.2240.212.7Grafting[145]
OMSPEI2.43251.21670.337.6Grafting[145]
OMS-APES 3.03251.21800.377.2Grafting[145]
OMS-APESPEI1.18251.2390.182.3Grafting[145]
OMS-NCCAmidoxime5.5412013150.699.3 [146]
MPS-MCC * 2.41120 3020.447.0 [147]
MPS-MCC ** 3.85120 2850.406.7 [147]
OMS-MgO 4.7112012610.487.25 [148]
OMS-CaO 3.8512011630.256.76 [148]
SiO2-Al2O3APTS2.642517401.245.1Grafting[149]
SiO2-Al(NO3)3APTS0.782513190.632.9Grafting[149]
OMS-Ti 0.81251487 [88]
MSiNTsAPTES2.87251.22930.7922Grafting[101]
SNSAPTES2.13251.22100.3119.6Grafting[101]
Al(NO3)3AP0.982513590.6210.0 [150]
OMS-Al-Zr 2.606014410.616.9 [151]
Where, ** MCC-mesoporous silica with amidoxime functionalities, * MCC-mesoporous silica with cyanopropyl groups, APTMS: 3-[2-(2-aminoethylamino)ethylamino]propyltrimethoxysilane, AEAPTMS: [3-(2-aminoethyl) aminopropyl]trimethoxysilane, AMP: 2-amino-2-methyl-1-propanol, AP: 3-aminopropyltriethoxysilane, APTMS: (3-aminopropyl) trimethoxysilane, APTS: 3-aminopropyltrimethoxysilane, DEA: diethanolamine, DEAPTMS: [3-(diethylamino) propyl]trimethoxysilane, DETA: diethylenetriamine, DWSNT: double-walled silica nano tube, EDA: ethylenediamine, HPS: Hierarchically porous silica, MAPTMS: [3-(methylamino) propyl]trimethoxysilane, MCC: microcrystalline cellulose, MEA: monoethanolamine, MPSM: monodispersed porous silica microspheres, MSiNTs: mesoporous silica nanotubes, NCC: nanocrystalline cellulose, OMS: ordered mesoporous organosilica, OMS: Oxide-templated silica monoliths, PEHA: pentaethylenehexamine, PEI: polyethylenimine, SNS: silica nano spheres, TEA: triethanolamine, TEPA: tetraethylenepentamine, TRI: 3-[2-(2-Aminoethylamino)ethylamino]propyltrimethoxysilane.
Sim and co-workers [145] studied the CO2 absorption capacity of the silica-based composites papered using SBA-15 and organosilica as silica precursors and N-[3-(trimethoxysilyl)propyl]ethylenediamine as an aminosilane precursor. Herein, PEI was grafted to the silica composites. Results exhibited that organosilica composites (see Table 7) showed the highest CO2 adsorption capacity, selectivity, and reproducibility. Another silica composite was prepared by Dassanayake et al. [146] using nanocrystalline cellulose (NCC) and reported that their NCC/mesoporous silica composite showed high CO2 absorption capacity (see Table 7), recyclability and thermal stability. Gunathilake et al. [147] synthesized microcrystalline cellulose (MCC) mesoporous silica composites using two MCC-mesoporous silica composites: MCC-mesoporous silica with cyanopropyl groups and MCC mesoporous silica amidoxime groups. CO2 adsorption was evaluated at 25 and 120 °C. According to the results, MCC-mesoporous silica with amidoxime functionalities exhibited the highest absorption capacity (see Table 7) at 120 °C due to the oxime and amine groups in amidoxime and hydroxyl groups in MCC which serve as active sites.
Rao et al. [144] determined the effect of impregnation and grafting of the amine-functionalized MCM-41. The results showed (see Table 7) grafted sorbents with higher thermal stability than the impregnation ones. They concluded that adsorbents modified by impregnation exhibited higher amine-loading efficiencies and, thus, higher CO2 adsorption capacities, whereas those prepared by grafting had better thermal and cyclic stability.
Moreover, Tang and co-workers have investigated the effect of inorganic alkalis such as (KOH, Ca(OH)2 and CsOH) on the CO2 absorption capacity [140]. The results showed that all three kinds of inorganic alkali-containing adsorbents exhibited higher CO2 adsorption capacities than tetraethylenepentamine (TEPA) and PEI-modified samples (see Table 7). This may be due to the introduction of inorganic alkali, which changes the chemical adsorption mechanism between adsorbate-CO2 and the adsorbent surface due to more hydroxyl groups. Moreover, they reported that CO2 adsorption capacities have a linear dependency with the amounts of alkali adsorbents. Apart from that, Gunathilake and Jaroniec [148] reported the incorporation of magnesium oxide (MgO) and calcium oxide (CaO) into mesoporous silica surface (OMS) and applied those materials for CO2 sorption at ambient and elevated temperatures. The materials were synthesized using the sol–gel method. However, composite sorbents performed relatively high adsorption capacities (see Table 7). It suggested that MgO and CaO enhanced CO2 adsorption via physisorption and chemisorption. Those synthesized CaO-SiO2 and MgO-SiO2 composites possessed high surface area, surface properties and thermal and chemical stability.
Alumina materials also possess high surface area, porosity, and thermal and mechanical stability. Therefore, researchers have recently used amine-grafted mesoporous silica and impregnated alumina as solid sorbents for CO2 capture [149]. Alumina-based materials for CO2 capture include basic Al2O3, amine-impregnated or amine-modified mesoporous Al2O3 and Al2O3–organosilica [149]. Gunathilake et al. [149] synthesized Al2O3–organosilica by introducing three different silica precursors such as tris [3-(trimethoxysilyl)propyl] isocyanurate (ICS), 1,4-bis(triethoxysilyl)benzene (BTEB), and bis(triethoxysilyl)ethane (BTEE)). This study used two alumina precursors, aluminum nitrate nanahydrate and aluminum isopropoxide, whereas grafting of amine groups was performed using 3-aminopropyltriethoxysilane (APTS). SiO2-Al2O3 showed the highest absorption capacity (Table 7), and the adsorption properties of the materials were dependent on the surface area of the sample, alumina precursor, and structure and functionality of the organosilica bridging group. Moreover, Choi et al. [152] used epoxy-functionalized PEI to synthesize CO2 sorbents. According to the reported data, epoxy-functionalized PEI exhibited a CO2 capacity of 2.2 mmol/g at 120 °C and 100% regeneration capability at similar temperatures. This can be attributed to the heat-resistant properties of epoxy butane, which enhanced the CO2 capture capacity and thermal stability of the silica-epoxy-PEI sorbent.
However, according to the reported data by Hu et al. [153], Li4SiO4 exhibited attractive prospects for CO2 capture. The main advantage of this material was the high CO2 sorption capacity (theoretical sorption capacity of 0.367 g CO2/g sorbent) and lower regeneration temperature (<750 °C) in comparison with other reported materials such as CaO, which requires a regeneration temperature of over 900 °C [153].

5.3. Sorbent Selectivity, Regeneration, and Stability in the Cyclic CO2 Adsorption–Desorption

During industrial applications, high adsorption capacity along with good regenerability of the sorbents in the cyclic adsorption–desorption process is vital [117]. The practical application of an adsorbent requires high sorption capacity, easy regeneration, stability in normal atmospheric conditions, and stable performance during cyclic use for long-term operation.
For instance, Ahmed et al. [93] reported a detailed study about the functionalization of mesoporous MCM-41 with different loadings of polyethylenimine (PEI). In this study, the selectivity measurement was conducted for CO2 over N2 and H2 and the adsorption capacities of N2 and H2 on 50 wt% PEI-Si-MCM-41 were 3.89 mg/g and 6.51 mg/g, respectively (see Table 8). Table 8 summarizes the gas selectivity values of previous studies performed for porous SiO2.
Wang et al. [154] prepared SBA-15 using silica-ethanol extraction and conventional high-temperature calcination template removal methods. Then, the silica was subjected to amine (3-aminopropyl) grafting and studied for its CO2 adsorption properties. This study aimed to increase the surface silanol density by grafting amine groups, increasing CO2 adsorption capacity and CO2/N2 selectivity. According to the reported data, CO2/N2 selectivity changed from 46 to 13 (see Table 8), and these results ensured that solvent extraction also enhanced CO2/N2 selectivity. Moreover, the authors performed a test to measure the stability of amine-SBA-15 (solvent extracted). According to the results, amine-SBA-15 (solvent extracted) was regenerated under a vacuum after each adsorption step.
In industrial applications of adsorbents, it is essential to remain stable during cyclic operations. This section summarizes the previous studies on sorbent regeneration and stability in cyclic CO2 adsorption–desorption by amine–silica composites, and the reported data are tabulated in Table 9. The regeneration of the amine-impregnated and grafted silica composites was mainly conducted by pressure and temperature swing adsorptions. Typically, the sorbent was regenerated at 50~120 °C in N2, He, or Ar flow. As depicted in Table 9, the amine-impregnated silica composites show a loss of CO2 capture capacity in the cyclic CO2 adsorption–desorption due to amine leaching from the silica surface and degradation [110]. Amine leaching is closely related to the amine types introduced and the operation temperature, while the degradation of amine is related to the operation temperature and gas atmosphere [109].
Guo et al. [128] conducted the adsorption/desorption cycles for hierarchically porous silica (HPS) grafted PEI at 75 °C. In this experiment, the modelled flue gas flow rate was maintained at 70 mL/min, and the CO2 partial pressure was held at 1 bar. According to the data, adsorption capacities are similar in eight adsorption/desorption cycles, showing that the aforementioned sorbents with good stability and regenerability.
Wang et al. [117] investigated the regenerability of the amine-modified MCM-41 (MCM-41-TEPA and MCM-41-AMP). The authors conducted fifteen cycles to verify the regenerability. According to the reported data, after fifteen cycles, the adsorption capacity decreased from 3.01 mmol/g to 2.88 mmol/g, and it was shown that both sorbents showed good regenerability. This may be due to the hydrogen-bonding interactions among TEPA, AMP and MCM-41, TEPA.
Kishor and Ghoshal [123] measured the stability of the pentaethylenehexamine (PEHA) impregnated KIT-6. The sorbent was aged for 6 months, and its adsorption performance was explored at 90–105 °C. The results showed that PEHA-impregnated KIT-6 had 4.0 and 4.3 mol CO2/kg sorption capacities at 90 and 105 °C at 1 bar even after 6 months. Moreover, the sorption performance of the adsorbent was tested for ten consecutive adsorption/desorption cycles. The sorption capacity of the sorbent decreased by less than 4% at 90–105 °C at 1 bar without any structural degradation. Moreover, the results exhibited that PEHA-impregnated KIT-6 had better sorption performance than those of earlier reported adsorbents, except for silica aerogel.
Liu and co-workers performed a regeneration test for zeolite-mesoporous silica-supported-amine hybrids sorbent [160]. Their data showed that, after 10 cycles, the adsorption capacity remained unchanged. Therefore, the sample performed a very stable cyclic adsorption–desorption performance. In contrast, López-Aranguren et al. [129] examined the regeneration of CO2 from branched PEI—mesoporous silica. In this study, CO2 adsorption–desorption cycles showed that the uptake measured in the first cycle was successfully maintained even after 20 cycles. Zhang et al. [174] examined the stability of the adsorbents based on linear PEI supported on silica. According to the reported data, the adsorbent maintained its adsorption capacity. Still, the adsorption capacity was reduced by approximately 5.6% when the temperature was increased to 100 °C, which was attributed to amine leaching. Furthermore, Subagyono et al. [162] found that the branched PEI-containing adsorbent decreased CO2 adsorption–desorption capacity during cycling, attributed to the by-product formation.

6. Technical Challenges and Future Trends

Financial, technical, and environmental concerns are the main barriers to CCS technologies. For instance, one major challenge with CCS is moving CO2 captured to remote storage sites using pipelines, as laying these pipelines is costly and associated with numerous environmental issues.
Several studies reported the requirements and a working definition for carbon dioxide capture (CCS). Advanced physical adsorbents must be developed with high CO2 selectivity and gas uptake. Stability (over 1000 cycles), CO2 affinity, scalability, reusability, resistance against surface erosion, and high energy requirement are the major concerns in CO2 capture technologies. The sorbent cost is the most significant part of an air capture system; however, it is difficult to estimate the price of a particular sorbent in lab-scale experiments. According to the reported data, the value of a kilogram of sorbent is equal to the net present value of the CO2 revenue collected during its lifetime. Therefore, a sorbent must possess constant stability and performance for its lifetime [178,179].
The other main challenges associated with sorbents are stability, kinetics, and sorbent capacity. However, many sorbents are thermodynamically strong enough to capture CO2 from ambient air and allow for easy regeneration. Despite the reported data, further studies on stability, kinetics and capacity still need to be improved in SiO2-based adsorbents. Another factor is sorbent loading and unloading cycles, which are essential for reducing costs. Moreover, adsorption kinetics is affected by binding energies, diffusion into porous materials, and the geometry of sorbent materials and many sorbents require longer sorption times. Therefore, improved kinetics can lower the cost. High adsorption capacity can reduce the cost of CO2 capture by reducing the amount of sorbent required. Physisorbents that selectively separate CO2 from gaseous mixtures formed a revolution in CCS since it requires less energy for recycling, with enhanced CO2 capacity.
Amine-based sorbents are widely used in CCS technologies. However, amine sorbent depends on the molecular weight of the sorbent and the pore sizes of the sorbent. To improve the capacity of moisture-swing sorbents, the ion exchange resins can be prepared with a higher charge density, and materials with different cation distances can be used under different humidity conditions. The potential of solid sorbents to remove CO2 from flue gas is enormous compared to conventional liquid amine processes in terms of regeneration energy and significant cost reduction. However, as discussed previously, solid sorbents have limitations and challenges to address before being deployed commercially in post-combustion CO2 capture.
There is limited literature available on CO2 capture using low-cost silica-based materials such as rice husks. These sources lead to the reduction in production costs. Nevertheless, novel silica-based materials such as lithium orthosilicate (Li4SiO4), silica nanotubes, silica nanospheres, silica-based composites, and silica aero gels give rise to high CO2 capture at elevated temperatures.
Moreover, most studies have used sol-gel and hydrothermal processes to synthesize silica-based sorbent. However, apart from the aforementioned methods, microwave treatment can also be used, which is cost-effective and timeserving. Moreover, different surfactants can prepare silica with varying pore sizes and morphologies. Another area for improvement with silica-based sorbent is the need for more literature on kinetic data at different adsorption temperatures, which are helpful in industrial implementations.

7. Summary

CO2 capture by porous SiO2 materials, their reaction mechanisms and synthesis processes were extensively discussed in this review. Chemical absorption of CO2 is more suitable than physical absorption owing to high adsorption capacity, relatively easy synthesis routes, and lower regeneration energy requirements. Among many chemisorbents, SiO2-based adsorbents, including amine-functionalized SiO2, possess higher CO2 selectivity and adsorption capacities, making them ideal candidates for CO2 capture. However, the performance of currently available amine-functionalized SiO2 needs to be further developed and improved in terms of stability, gas selectivity and resistivity to thermal degradation. Furthermore, the review highlighted major financial, technical, and environmental barriers and prospects associated with porous silica-based materials during the industrial scale-up process.

Author Contributions

Conceptualization, S.M.A., C.A.G., Y.D., R.S.D. and E.-B.C.; methodology, S.M.A. and C.A.G.; software, S.M.A., O.H.P.G. and C.A.G.; validation, S.M.A., C.A.G., Y.D. and R.S.D.; formal analysis, S.M.A. and C.A.G.; investigation, S.M.A., C.A.G., Y.D., R.S.D. and E.-B.C.; resources, S.M.A. and C.A.G.; data curation, S.M.A. and C.A.G.; writing—original draft preparation, S.M.A., O.H.P.G., C.A.G. and R.S.D.; writing—review and editing, S.M.A., C.A.G., Y.D., R.S.D. and E.-B.C.; visualization, S.M.A., C.A.G., Y.D. and R.S.D.; supervision, C.A.G., Y.D., R.S.D. and E.-B.C.; project administration, C.A.G., Y.D., R.S.D. and E.-B.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by the National Research Foundation of Korea.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors would like to express their gratitude to the Department of Chemical and Process Engineering, University of Peradeniya. This study was supported by the Human Resource Development Programs for Green Convergence Technology funded by the Korea Ministry of Environment (MOE). E.-B. Cho acknowledges supports under the National Research Foundation of Korea (NRF-2022R1I1A2054213).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations and Acronyms

AbbreviationDefinition
CO2Carbon dioxide
Li4SiO4Lithium orthosilicate
CCSCarbon dioxide capture
PEHAPentaethylenehexamine
HSSPHollow silica spherical particles
PEIPolyethylenimine
APTS3-aminopropyltriethoxysilane
MCCMesoporous silica with amidoxime functionalities
APTMS3-[2-(2-aminoethylamino)ethylamino]propyltrimethoxysilane,
TRI3-[2-(2-aminoethylamino)ethylamino]propyltrimethoxysilane
TEPATetraethylenepentamine
TEATriethanolamine
SNSSilica nano spheres
OMSOxide-templated silica
NCCNanocrystalline cellulose
MSiNTsMesoporous silica nanotubes
MPSMMonodispersed porous silica microspheres
MCCMicrocrystalline cellulose
HPSHierarchically porous silica
MSPDMatrix solid phase dispersion
FCCFace-centered cubic
SBASanta Barbara amorphous family
MWSAMicrowave-swing adsorption
PSAPressure swing adsorption
TSATemperature swing adsorption
ESAElectric swing adsorption
VSAVacuum swing adsorption

References

  1. Osman, A.I.; Hefny, M.; Maksoud, M.A.; Elgarahy, A.M.; Rooney, D.W. Recent advances in carbon capture storage and utilisation technologies: A review. Environ. Chem. Lett. 2020, 19, 797–849. [Google Scholar] [CrossRef]
  2. Ahn, D. Quantifying the Emissions of Carbon Dioxide (CO2), Carbon Monoxide (Co), and Nitrogen Oxides (Nx) from Human Activities: Top-Down and Bottom-Up Approaches Doctoral Dissertation; University of Maryland: College Park, ML, USA, 2021. [Google Scholar]
  3. Gunawardene, O.H.; Gunathilake, C.A.; Vikrant, K.; Amaraweera, S.M. Carbon Dioxide Capture through Physical and Chemical Adsorption Using Porous Carbon Materials: A Review. Atmosphere 2022, 13, 397. [Google Scholar] [CrossRef]
  4. Leung, D.Y.; Caramanna, G.; Maroto-Valer, M.M. An overview of current status of carbon dioxide capture and storage technologies. Renew. Sustain. Energy Rev. 2014, 39, 426–443. [Google Scholar] [CrossRef] [Green Version]
  5. Omoregbe, O.; Mustapha, A.N.; Steinberger-Wilckens, R.; El-Kharouf, A.; Onyeaka, H. Carbon capture technologies for climate change mitigation: A bibliometric analysis of the scientific discourse during 1998–2018. Energy Rep. 2020, 6, 1200–1212. [Google Scholar] [CrossRef]
  6. Lee, S.Y.; Park, S.J. A review on solid adsorbents for carbon dioxide capture. J. Ind. Eng. Chem. 2015, 23, 1–11. [Google Scholar] [CrossRef]
  7. Bui, M.; Gunawan, I.; Verheyen, V.; Feron, P.; Meuleman, E.; Adeloju, S. Dynamic modelling and optimisation of flexible operation in post-combustion CO2 capture plants—A review. Comput. Chem. Eng. 2014, 61, 245–265. [Google Scholar] [CrossRef]
  8. Cotton, A.; Patchigolla, K.; Oakey, J.E. Minor and trace element emissions from post-combustion CO2 capture from coal: Experimental and equilibrium calculations. Fuel 2014, 117, 391–407. [Google Scholar] [CrossRef] [Green Version]
  9. Goto, K.; Yogo, K.; Higashii, T. A review of efficiency penalty in a coal-fired power plant with post-combustion CO2 capture. Appl. Energy 2013, 111, 710–720. [Google Scholar] [CrossRef]
  10. Zhang, N.; Pan, Z.; Zhang, Z.; Zhang, W.; Zhang, L.; Baena-Moreno, F.M.; Lichtfouse, E. CO2 capture from coalbed methane using membranes: A review. Environ. Chem. Lett. 2020, 18, 79–96. [Google Scholar] [CrossRef]
  11. Sreenivasalu, B.; Gayatri, D.V.; Sreedhar, I.; Ragharan, K.V. A journey into the process and engineering aspects of carbon capture technologies. Renew. Sustain. Energy Rev. 2015, 41, 1324–1350. [Google Scholar] [CrossRef]
  12. Singh, J.; Bhunia, H.; Basu, S. Development of sulfur-doped carbon monolith derived from phenol-formaldehyde resin for fixed bed CO2 adsorption. Environ. Innov. 2020, 20, 101104. [Google Scholar] [CrossRef]
  13. Sánchez, J.M.; Maroño, M.; Cillero, D.; Montenegro, L.; Ruiz, E. Laboratory-and bench-scale studies of a sweet water–gas-shift catalyst for H2 and CO2 production in pre-combustion CO2 capture. Fuel 2013, 114, 191–198. [Google Scholar] [CrossRef]
  14. Babu, P.; Kumar, R.; Linga, P. A new porous material to enhance the kinetics of clathrate process: Application to precombustion carbon dioxide capture. Environ. Sci. Technol. 2013, 47, 13191–13198. [Google Scholar] [CrossRef]
  15. Wienchol, P.; Szlȩk, A.; Ditaranto, M. Waste-to-energy technology integrated with carbon capture—Challenges and opportunities. Energy 2020, 198, 117352. [Google Scholar] [CrossRef]
  16. Kim, Y.E.; Moon, S.J.; Yoon, Y.I.; Jeong, S.K.; Park, K.T.; Bae, S.T.; Nam, S.C. Heat of absorption and absorption capacity of CO2 in aqueous solutions of amine containing multiple amino groups. Sep. Purif. Technol. 2014, 122, 112–118. [Google Scholar] [CrossRef]
  17. Roth, E.A.; Agarwal, S.; Gupta, R.K. Nanoclay-based solid sorbents for CO2 capture. Energy Fuels 2013, 27, 4129–4136. [Google Scholar] [CrossRef]
  18. Kenarsari, S.D.; Yang, D.; Jiang, G.; Zhang, S.; Wang, J.; Russell, A.G.; Wei, Q.; Fan, M. Review of recent advances in carbon dioxide separation and capture. RSC Adv. 2013, 3, 22739–22773. [Google Scholar] [CrossRef]
  19. Zhang, X.; He, X.; Gundersen, T. Post-combustion carbon capture with a gas separation membrane: Parametric study, capture cost, and exergy analysis. Energy Fuels 2013, 27, 4137–4149. [Google Scholar] [CrossRef]
  20. Scholes, C.A.; Ho, M.T.; Wiley, D.E.; Stevens, G.W.; Kentish, S.E. Cost competitive membrane—Cryogenic post-combustion carbon capture. Int. J. Greenh. Gas Control 2013, 17, 341–348. [Google Scholar] [CrossRef]
  21. Khraisheh, M.; Mukherjee, S.; Kumar, A.; Al Momani, F.; Walker, G.; Zaworotko, M.J. An overview on trace CO2 removal by advanced physisorbent materials. J. Environ. Manag. 2020, 255, 109874. [Google Scholar] [CrossRef]
  22. Anwar, M.N.; Fayyaz, A.; Sohail, N.F.; Khokhar, M.F.; Baqar, M.; Khan, W.D.; Rasool, K.; Rehan, M.; Nizami, A.S. CO2 capture and storage: A way forward for sustainable environment. J. Environ. Manag. 2018, 226, 131–144. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, G.; Liang, F.; Yang, Y.; Hu, Y.; Zhang, K.; Liu, W. An improved CO2 separation and purification system based on cryogenic separation and distillation theory. Energies 2014, 7, 3484–3502. [Google Scholar] [CrossRef] [Green Version]
  24. Samanta, A.; Zhao, A.; Shimizu, G.K.; Sarkar, P.; Gupta, R. Post-combustion CO2 capture using solid sorbents: A review. Ind. Eng. Chem. Res. 2012, 51, 1438–1463. [Google Scholar] [CrossRef]
  25. Patel, H.A.; Byun, J.; Yavuz, C.T. Carbon dioxide capture adsorbents: Chemistry and methods. ChemSusChem 2017, 10, 1303–1317. [Google Scholar] [CrossRef]
  26. Sumida, K.; Rogow, D.L.; Mason, J.A.; McDonald, T.M.; Bloch, E.D.; Herm, Z.R.; Bae, T.H.; Long, J.R. Carbon dioxide capture in metal–organic frameworks. Chem. Rev. 2012, 112, 724–781. [Google Scholar] [CrossRef]
  27. Reynolds, A.J.; Verheyen, T.V.; Adeloju, S.B.; Meuleman, E.; Feron, P. Towards commercial scale postcombustion capture of CO2 with monoethanolamine solvent: Key considerations for solvent management and environmental impacts. Environ. Sci. Technol. 2012, 46, 3643–3654. [Google Scholar] [CrossRef]
  28. Mumford, K.A.; Wu, Y.; Smith, K.H.; Stevens, G.W. Review of solvent based carbon-dioxide capture technologies. Front. Chem. Sci. Eng. 2015, 9, 125–141. [Google Scholar] [CrossRef]
  29. Park, J.H.; Celedonio, J.M.; Seo, H.; Park, Y.K.; Ko, Y.S. A study on the effect of the amine structure in CO2 dry sorbents on CO2 capture. Catal. Today 2016, 265, 68–76. [Google Scholar] [CrossRef]
  30. Gunathilake, C.; Gangoda, M.; Jaroniec, M. Mesoporous alumina with amidoxime groups for CO2 sorption at ambient and elevated temperatures. Ind. Eng. Chem. Res. 2016, 55, 5598–5607. [Google Scholar] [CrossRef]
  31. Lv, B.; Guo, B.; Zhou, Z.; Jing, G. Mechanisms of CO2 capture into monoethanolamine solution with different CO2 loading during the absorption/desorption processes. Environ. Sci. Technol. 2015, 49, 10728–10735. [Google Scholar] [CrossRef]
  32. García-Abuín, A.; Gomez-Diaz, D.; Lopez, A.B.; Navaza, J.M.; Rumbo, A. NMR characterization of carbon dioxide chemical absorption with monoethanolamine, diethanolamine, and triethanolamine. Ind. Eng. Chem. Res. 2013, 52, 13432–13438. [Google Scholar] [CrossRef]
  33. Kim, Y.E.; Lim, J.A.; Jeong, S.K.; Yoon, Y.I.; Bae, S.T.; Nam, S.C. Comparison of carbon dioxide absorption in aqueous MEA, DEA, TEA, and AMP solutions. Bull. Korean Chem. Soc. 2013, 34, 783–787. [Google Scholar] [CrossRef] [Green Version]
  34. Rinprasertmeechai, S.; Chavadej, S.; Rangsunvigit, P.; Kulprathipanja, S. Carbon dioxide removal from flue gas using amine-based hybrid solvent absorption. Int. J. Chem. Eng. 2012, 6, 296–300. [Google Scholar]
  35. Aaron, D.; Tsouris, C. Separation of CO2 from flue gas: A review. Sep. Sci. Technol. 2005, 40, 321–348. [Google Scholar] [CrossRef]
  36. Songolzadeh, M.; Soleimani, M.; Takht Ravanchi, M.; Songolzadeh, R. Carbon dioxide separation from flue gases: A technological review emphasizing reduction in greenhouse gas emissions. Sci. World J. 2014, 2014, 828131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Wang, M.; Lawal, A.; Stephenson, P.; Sidders, J.; Ramshaw, C. Post-combustion CO2 capture with chemical absorption: A state-of-the-art review. Chem. Eng. Res. Des. 2011, 89, 1609–1624. [Google Scholar] [CrossRef] [Green Version]
  38. Dave, N.; Do, T.; Puxty, G.; Rowland, R.; Feron, P.H.M.; Attalla, M.I. CO2 capture by aqueous amines and aqueous ammonia–A Comparison. Energy Procedia 2009, 1, 949–954. [Google Scholar] [CrossRef] [Green Version]
  39. Zeng, S.; Zhang, X.; Bai, L.; Zhang, X.; Wang, H.; Wang, J.; Bao, D.; Li, M.; Liu, X.; Zhang, S. Ionic-liquid-based CO2 capture systems: Structure, interaction and process. Chem. Rev. 2017, 117, 9625–9673. [Google Scholar] [CrossRef]
  40. Borhani, T.N.; Wang, M. Role of solvents in CO2 capture processes: The review of selection and design methods. Renew. Sustain. Energy Rev. 2019, 114, 109299. [Google Scholar] [CrossRef]
  41. Aghaie, M.; Rezaei, N.; Zendehboudi, S. A systematic review on CO2 capture with ionic liquids: Current status and future prospects. Renew. Sustain. Energy Rev. 2018, 96, 502–525. [Google Scholar] [CrossRef]
  42. Jin, X.; Ge, J.; Zhang, L.; Wu, Z.; Zhu, L.; Xiong, M. Synthesis of Hierarchically Ordered Porous Silica Materials for CO2 Capture: The Role of Pore Structure and Functionalized Amine. Inorganics 2022, 10, 87. [Google Scholar] [CrossRef]
  43. Zagho, M.M.; Hassan, M.K.; Khraisheh, M.; Al-Maadeed, M.A.A.; Nazarenko, S. A review on recent advances in CO2 separation using zeolite and zeolite-like materials as adsorbents and fillers in mixed matrix membranes (MMMs). Chem. Eng. J. Adv. 2021, 6, 100091. [Google Scholar] [CrossRef]
  44. Yu, C.H.; Huang, C.H.; Tan, C.S. A review of CO2 capture by absorption and adsorption. Aerosol Air Qual. Res. 2012, 12, 745–769. [Google Scholar] [CrossRef] [Green Version]
  45. Kumar, S.; Bera, R.; Das, N.; Koh, J. Chitosan-based Zeolite-Y and ZSM-5 porous biocomposites for H2 and CO2 storage. Carbohydr. Polym. 2020, 232, 115808. [Google Scholar] [CrossRef] [PubMed]
  46. Kumar, S.; Srivastava, R.; Koh, J. Utilization of zeolites as CO2 capturing agents: Advances and future perspectives. J. CO2 Util. 2020, 41, 101251. [Google Scholar] [CrossRef]
  47. Balashankar, V.S.; Rajendran, A. Process optimization-based screening of zeolites for post-combustion CO2 capture by vacuum swing adsorption. ACS Sustain. Chem. Eng. 2019, 7, 17747–17755. [Google Scholar] [CrossRef]
  48. Avci, G.; Erucar, I.; Keskin, S. Do new MOFs perform better for CO2 capture and H2 purification? Computational screening of the updated MOF database. ACS Appl. Mater. Interfaces 2020, 12, 41567–41579. [Google Scholar] [CrossRef]
  49. Modak, A.; Jana, S. Advancement in porous adsorbents for post-combustion CO2 capture. Microporous Mesoporous Mater. 2019, 276, 107–132. [Google Scholar] [CrossRef]
  50. Le, M.U.T.; Lee, S.Y.; Park, S.J. Preparation and characterization of PEI-loaded MCM-41 for CO2 capture. Int. J. Hydrogen Energy 2014, 39, 12340–12346. [Google Scholar]
  51. Kim, M.I.; Choi, S.J.; Kim, D.W.; Park, D.W. Catalytic performance of zinc containing ionic liquids immobilized on silica for the synthesis of cyclic carbonates. J. Ind. Eng. Chem. 2014, 20, 3102–3107. [Google Scholar] [CrossRef]
  52. Alkhabbaz, M.A.; Khunsupat, R.; Jones, C.W. Guanidinylated poly (allylamine) supported on mesoporous silica for CO2 capture from flue gas. Fuel 2014, 121, 79–85. [Google Scholar] [CrossRef]
  53. Reddy, M.S.B.; Ponnamma, D.; Sadasivuni, K.K.; Kumar, B.; Abdullah, A.M. Carbon dioxide adsorption based on porous materials. RSC Adv. 2021, 11, 12658–12681. [Google Scholar] [CrossRef] [PubMed]
  54. Liu, R.S.; Shi, X.D.; Wang, C.T.; Gao, Y.Z.; Xu, S.; Hao, G.P.; Chen, S.; Lu, A.H. Advances in Post-Combustion CO2 Capture by Physical Adsorption: From Materials Innovation to Separation Practice. ChemSusChem 2021, 14, 1428–1471. [Google Scholar] [CrossRef] [PubMed]
  55. Boot-Handford, M.E.; Abanades, J.C.; Anthony, E.J.; Blunt, M.J.; Brandani, S.; Mac Dowell, N.; Fernández, J.R.; Ferrari, M.C.; Gross, R.; Hallett, J.P.; et al. Carbon capture and storage update. Energy Environ. Sci. 2014, 7, 130–189. [Google Scholar] [CrossRef]
  56. MacFarlane, D.R.; Tachikawa, N.; Forsyth, M.; Pringle, J.M.; Howlett, P.C.; Elliott, G.D.; Davis, J.H.; Watanabe, M.; Simon, P.; Angell, C.A. Energy applications of ionic liquids. Energy Environ. Sci. 2014, 7, 232–250. [Google Scholar] [CrossRef] [Green Version]
  57. Tuci, G.; Iemhoff, A.; Ba, H.; Luconi, L.; Rossin, A.; Papaefthimiou, V.; Palkovits, R.; Artz, J.; Pham-Huu, C.; Giambastiani, G. Playing with covalent triazine framework tiles for improved CO2 adsorption properties and catalytic performance. Beilstein J. Nanotechnol. 2019, 10, 1217–1227. [Google Scholar] [CrossRef] [Green Version]
  58. Madden, D.G.; O’Nolan, D.; Chen, K.J.; Hua, C.; Kumar, A.; Pham, T.; Forrest, K.A.; Space, B.; Perry, J.J.; Khraisheh, M.; et al. Highly selective CO2 removal for one-step liquefied natural gas processing by physisorbents. Chem. Commun. 2019, 55, 3219–3222. [Google Scholar] [CrossRef]
  59. Balsamo, M.; Budinova, T.; Erto, A.; Lancia, A.; Petrova, B.; Petrov, N.; Tsyntsarski, B. CO2 adsorption onto synthetic activated carbon: Kinetic, thermodynamic and regeneration studies. Sep. Purif. Technol. 2013, 116, 214–221. [Google Scholar] [CrossRef]
  60. Cherbanski, R.; Komorowska-Durka, M.; Stefanidis, G.D.; Stankiewicz, A.I. Microwave Swing Regeneration Vs Temperature Swing Regeneration Comparison of Desorption Kinetics. Ind. Eng. Chem. Res. 2011, 50, 8632–8644. [Google Scholar] [CrossRef]
  61. Maity, A.; Belgamwar, R.; Polshettiwar, V. Facile synthesis to tune size, textural properties and fiber density of dendritic fibrous nanosilica for applications in catalysis and CO2 capture. Nat. Protoc. 2019, 14, 2177–2204. [Google Scholar] [CrossRef]
  62. Miricioiu, M.G.; Niculescu, V.C. Fly ash, from recycling to potential raw material for mesoporous silica synthesis. Nanomaterials 2020, 10, 474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Ying, W.; Zhao, D. On the controllable soft-templating approach to mesoporous silicates. Chem. Rev. 2007, 107, 2821–2860. [Google Scholar]
  64. Narayan, R.; Nayak, U.Y.; Raichur, A.M.; Garg, S. Mesoporous silica nanoparticles: A comprehensive review on synthesis and recent advances. Pharmaceutics 2018, 10, 118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Costa, J.A.S.; de Jesus, R.A.; Santos, D.O.; Mano, J.F.; Romao, L.P.; Paranhos, C.M. Recent progresses in the adsorption of organic, inorganic, and gas compounds by MCM-41-based mesoporous materials. Microporous Mesoporous Mater. 2020, 291, 109698. [Google Scholar] [CrossRef]
  66. Brezoiu, A.M.; Deaconu, M.; Nicu, I.; Vasile, E.; Mitran, R.A.; Matei, C.; Berger, D. Heteroatom modified MCM-41-silica carriers for Lomefloxacin delivery systems. Microporous Mesoporous Mater. 2019, 275, 214–222. [Google Scholar] [CrossRef]
  67. Jesus, R.A.; Rabelo, A.S.; Figueiredo, R.T.; Da Silva, L.C.; Codentino, I.C.; Fantini, M.C.D.A.; Araújo, G.L.B.D.; Araújo, A.A.S.; Mesquita, M.E. Synthesis and application of the MCM-41 and SBA-15 as matrices for in vitro efavirenz release study. J. Drug Deliv. Sci. Technol. 2016, 31, 153–159. [Google Scholar] [CrossRef]
  68. Liu, Y.; Li, C.; Peyravi, A.; Sun, Z.; Zhang, G.; Rahmani, K.; Zheng, S.; Hashisho, Z. Mesoporous MCM-41 derived from natural Opoka and its application for organic vapors removal. J. Hazard. Mater. 2021, 408, 124911. [Google Scholar] [CrossRef]
  69. Costa, J.A.S.; Paranhos, C.M. Mitigation of silica-rich wastes: An alternative to the synthesis eco-friendly silica-based mesoporous materials. Microporous Mesoporous Mater. 2020, 309, 110570. [Google Scholar] [CrossRef]
  70. Björk, E.M. Synthesizing and characterizing mesoporous silica SBA-15: A hands-on laboratory experiment for undergraduates using various instrumental techniques. J. Chem. Educ. 2017, 94, 91–94. [Google Scholar] [CrossRef] [Green Version]
  71. Martínez, M.L.; Ponte, M.V.; Beltramone, A.R.; Anunziata, O.A. Synthesis of ordered mesoporous SBA-3 materials using silica gel as silica source. Mater. Lett. 2014, 134, 95–98. [Google Scholar] [CrossRef]
  72. López-Mendoza, M.A.; Nava, R.; Peza-Ledesma, C.; Millán-Malo, B.; Huirache-Acuña, R.; Skewes, P.; Rivera-Muñoz, E.M. Characterization and catalytic performance of Co-Mo-W sulfide catalysts supported on SBA-15 and SBA-16 mechanically mixed. Catal. Today 2016, 271, 114–126. [Google Scholar] [CrossRef]
  73. Gonzalez, G.; Sagarzazu, A.; Cordova, A.; Gomes, M.E.; Salas, J.; Contreras, L.; Noris-Suarez, K.; Lascano, L. Comparative study of two silica mesoporous materials (SBA-16 and SBA-15) modified with a hydroxyapatite layer for clindamycin controlled delivery. Microporous Mesoporous Mater. 2018, 256, 251–265. [Google Scholar] [CrossRef]
  74. Wu, Q.; Li, Y.; Hou, Z.; Xin, J.; Meng, Q.; Han, L.; Xiao, C.; Hu, D.; Duan, A.; Xu, C. Synthesis and characterization of Beta-FDU-12 and the hydrodesulfurization performance of FCC gasoline and diesel. Fuel Process. Technol. 2018, 172, 55–64. [Google Scholar] [CrossRef]
  75. Sanaeishoar, H.; Sabbaghan, M.; Mohave, F.; Nazarpour, R. Disordered mesoporous KIT-1 synthesized by DABCO-based ionic liquid and its characterization. Microporous Mesoporous Mater. 2016, 228, 305–309. [Google Scholar] [CrossRef]
  76. Stöber, W.; Fink, A.; Bohn, E. Controlled growth of monodisperse silica spheres in micron size range. J. Colloid Interface Sci. 1968, 26, 62–69. [Google Scholar] [CrossRef]
  77. Choma, J.; Jamioła, D.; Augustynek, K.; Marszewski, M.; Gao, M.; Jaroniec, M. New opportunities in Stöber synthesis: Preparation of microporous and mesoporous carbon spheres. J. Mater. Chem. 2012, 22, 12636–12642. [Google Scholar] [CrossRef]
  78. Wu, S.H.; Mou, C.Y.; Lin, H.P. Synthesis of mesoporous silica nanoparticles. Chem. Soc. Rev. 2013, 42, 3862–3875. [Google Scholar] [CrossRef]
  79. Sakamoto, S.; Yoshikawa, M.; Ozawa, K.; Kuroda, Y.; Shimojima, A.; Kuroda, K. Formation of single-digit nanometer scale silica nanoparticles by evaporation-induced self-assembly. Langmuir 2018, 34, 1711–1717. [Google Scholar] [CrossRef]
  80. Sihler, S.; Nguyen, P.L.; Lindén, M.; Ziener, U. Green chemistry in red emulsion: Interface of dye stabilized emulsions as a powerful platform for the formation of sub-20-nm SiO2 nanoparticles. ACS Appl. Mater. Interfaces 2018, 10, 24310–24319. [Google Scholar] [CrossRef]
  81. Murray, E.; Born, P.; Weber, A.; Kraus, T. Synthesis of Monodisperse Silica Nanoparticles Dispersable in Non-Polar Solvents. Adv. Eng. Mater. 2010, 12, 374–378. [Google Scholar] [CrossRef]
  82. Mandal, M.; Kruk, M. Family of single-micelle-templated organosilica hollow nanospheres and nanotubes synthesized through adjustment of organosilica/surfactant ratio. Chem. Mater. 2012, 24, 123–132. [Google Scholar] [CrossRef]
  83. Savic, S.; Vojisavljevic, K.; Počuča-Nešić, M.; Zivojevic, K.; Mladenovic, M.; Knezevic, N. Hard Template Synthesis of Nanomaterials Based on Mesoporous Silica. Metall. Mater. Eng. 2018, 24, 225–241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhao, T.; Elzatahry, A.; Li, X.; Zhao, D. Single-micelle-directed synthesis of mesoporous materials. Nat. Rev. Mater. 2019, 4, 775–791. [Google Scholar] [CrossRef]
  85. Khan, A.H.; Ghosh, S.; Pradhan, B.; Dalui, A.; Shrestha, L.K.; Acharya, S.; Ariga, K. Two-dimensional (2D) nanomaterials towards electrochemical nanoarchitectonics in energy-related applications. Bull. Chem. Soc. Jpn. 2017, 90, 627–648. [Google Scholar] [CrossRef]
  86. Yamamoto, E.; Kuroda, K. Colloidal mesoporous silica nanoparticles. Bull. Chem. Soc. Jpn. 2016, 89, 501–539. [Google Scholar] [CrossRef]
  87. Kim, H.J.; Yang, H.C.; Chung, D.Y.; Yang, I.H.; Choi, Y.J.; Moon, J.K. Functionalized mesoporous silica membranes for CO2 separation applications. J. Chem. 2015, 2015, 202867. [Google Scholar] [CrossRef] [Green Version]
  88. Gunathilake, C.; Kalpage, C.; Kadanapitiye, M.; Dassanayake, R.S.; Manchanda, A.S.; Gangoda, M. Facile synthesis and surface characterization of titania-incorporated mesoporous organosilica materials. J. Compos. Sci. 2019, 3, 77. [Google Scholar] [CrossRef] [Green Version]
  89. Hao, P.; Peng, B.; Shan, B.Q.; Yang, T.Q.; Zhang, K. Comprehensive understanding of the synthesis and formation mechanism of dendritic mesoporous silica nanospheres. Nanoscale Adv. 2020, 2, 1792–1810. [Google Scholar] [CrossRef]
  90. Panek, R.; Wdowin, M.; Franus, W.; Czarna, D.; Stevens, L.A.; Deng, H.; Liu, J.; Sun, C.; Liu, H.; Snape, C.E. Fly ash-derived MCM-41 as a low-cost silica support for polyethyleneimine in post-combustion CO2 capture. J. CO2 Util. 2017, 22, 81–90. [Google Scholar] [CrossRef]
  91. Singh, B.; Polshettiwar, V. Solution-phase synthesis of two-dimensional silica nanosheets using soft templates and their applications in CO2 capture. Nanoscale 2019, 11, 5365–5376. [Google Scholar] [CrossRef]
  92. Li, Y.; Wang, X.; Cao, M. Three-dimensional porous carbon frameworks derived from mangosteen peel waste as promising materials for CO2 capture and supercapacitors. J. CO2 Util. 2018, 27, 204–216. [Google Scholar] [CrossRef]
  93. Ahmed, S.; Ramli, A.; Yusup, S. Development of polyethylenimine-functionalized mesoporous Si-MCM-41 for CO2 adsorption. Fuel Process. Technol. 2017, 167, 622–630. [Google Scholar] [CrossRef]
  94. Son, W.J.; Choi, J.S.; Ahn, W.S. Adsorptive removal of carbon dioxide using polyethyleneimine-loaded mesoporous silica materials. Microporous Mesoporous Mater. 2008, 113, 31–40. [Google Scholar] [CrossRef]
  95. Zeleňák, V.; Badaničová, M.; Halamova, D.; Čejka, J.; Zukal, A.; Murafa, N.; Goerigk, G. Amine-modified ordered mesoporous silica: Effect of pore size on carbon dioxide capture. Chem. Eng. J. 2008, 144, 336–342. [Google Scholar] [CrossRef]
  96. Lashaki, M.J.; Sayari, A. CO2 capture using triamine-grafted SBA-15: The impact of the support pore structure. Chem. Eng. J. 2018, 334, 1260–1269. [Google Scholar] [CrossRef]
  97. Kishor, R.; Ghoshal, A.K. APTES grafted ordered mesoporous silica KIT-6 for CO2 adsorption. Chem. Eng. J. 2015, 262, 882–890. [Google Scholar] [CrossRef]
  98. Ahmed, S.; Ramli, A.; Yusup, S. CO2 adsorption study on primary, secondary and tertiary amine functionalized Si-MCM-41. Int. J. Greenh. Gas Control 2016, 51, 230–238. [Google Scholar] [CrossRef]
  99. Nigar, H.; Garcia-Baños, B.; Peñaranda-Foix, F.L.; Catalá-Civera, J.M.; Mallada, R.; Santamaría, J. Amine-functionalized mesoporous silica: A material capable of CO2 adsorption and fast regeneration by microwave heating. AIChE J. 2016, 62, 547–555. [Google Scholar] [CrossRef]
  100. Heydari-Gorji, A.; Yang, Y.; Sayari, A. Effect of the pore length on CO2 adsorption over amine-modified mesoporous silicas. Energy Fuels 2011, 25, 4206–4210. [Google Scholar] [CrossRef]
  101. Gunathilake, C.; Manchanda, A.S.; Ghimire, P.; Kruk, M.; Jaroniec, M. Amine-modified silica nanotubes and nanospheres: Synthesis and CO2 sorption properties. Environ. Sci. 2016, 3, 806–817. [Google Scholar] [CrossRef]
  102. Cabriga, C.K.C.; Clarete, K.V.R.; Zhang, J.A.T.; Pacia, R.M.P.; Ko, Y.S.; Castro, J.C. Evaluation of biochar derived from the slow pyrolysis of rice straw as a potential adsorbent for carbon dioxide. Biomass Convers. Biorefin. 2021, 13, 7887–7894. [Google Scholar] [CrossRef]
  103. Wang, J.; Yuan, X.; Deng, S.; Zeng, X.; Yu, Z.; Li, S.; Lia, K. Waste polyethylene terephthalate (PET) plastics-derived activated carbon for CO2 capture: A route to a closed carbon loop. Green Chem. 2020, 22, 6836–6845. [Google Scholar] [CrossRef]
  104. Berger, A.H.; Bhown, A.S. Comparing physisorption and chemisorption solid sorbents for use separating CO2 from flue gas using temperature swing adsorption. Energy Proc. 2011, 4, 562–567. [Google Scholar] [CrossRef] [Green Version]
  105. Moni, P.; Chaves, W.F.; Wilhelm, M.; Rezwan, K. Polysiloxane microspheres encapsulated in carbon allotropes: A promising material for supercapacitor and carbon dioxide capture. J. Colloid Interface Sci. 2019, 542, 91–101. [Google Scholar] [CrossRef]
  106. Hahn, M.W.; Jelic, J.; Berger, E.; Reuter, K.; Jentys, A.; Lercher, J.A. Role of amine functionality for CO2 chemisorption on silica. J. Phys. Chem. B 2016, 120, 1988–1995. [Google Scholar] [CrossRef] [PubMed]
  107. Chen, C.; Yang, S.T.; Ahn, W.S.; Ryoo, R. Amine-impregnated silica monolith with a hierarchical pore structure: Enhancement of CO2 capture capacity. Chem. Commun. 2009, 120, 627–3629. [Google Scholar] [CrossRef]
  108. Chen, C.; Son, W.J.; You, K.S.; Ahn, J.W.; Ahn, W.S. Carbon dioxide capture using amine-impregnated HMS having textural mesoporosity. Chem. Eng. J. 2010, 161, 46–52. [Google Scholar] [CrossRef]
  109. Chen, C.; Zhang, S.; Row, K.H.; Ahn, W.S. Amine–silica composites for CO2 capture: A short review. J. Energy Chem. 2017, 26, 868–880. [Google Scholar] [CrossRef] [Green Version]
  110. Wei, J.; Liao, L.; Xiao, Y.; Zhang, P.; Shi, Y. Capture of carbon dioxide by amine-impregnated as-synthesized MCM-41. J. Environ. Sci. 2010, 22, 1558–1563. [Google Scholar] [CrossRef]
  111. Chew, T.L.; Ahmad, A.L.; Bhatia, S. Ordered mesoporous silica (OMS) as an adsorbent and membrane for separation of carbon dioxide (CO2). Adv. Colloid Interface Sci. 2010, 153, 43–57. [Google Scholar] [CrossRef]
  112. Li, W.; Choi, S.; Drese, J.H.; Hornbostel, M.; Krishnan, G.; Eisenberger, P.M.; Jones, C.W. Steam-stripping for regeneration of supported amine-based CO2 adsorbents. ChemSusChem 2010, 3, 899–903. [Google Scholar] [CrossRef] [PubMed]
  113. Bollini, P.; Didas, S.A.; Jones, C.W. Amine-oxide hybrid materials for acid gas separations. J. Mater. Chem. 2011, 21, 15100–15120. [Google Scholar] [CrossRef]
  114. Li, Y.; Sun, N.; Li, L.; Zhao, N.; Xiao, F.; Wei, W.; Sun, Y.; Huang, W. Grafting of amines on ethanol-extracted SBA-15 for CO2 adsorption. Materials 2013, 6, 981–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Lopez-Aranguren, P.; Builes, S.; Fraile, J.; Vega, L.F.; Domingo, C. Understanding the performance of new amine-functionalized mesoporous silica materials for CO2 adsorption. Ind. Eng. Chem. Res. 2014, 53, 15611–15619. [Google Scholar] [CrossRef]
  116. Sanz, R.; Calleja, G.; Arencibia, A.; Sanz-Perez, E.S. CO2 capture with pore-expanded MCM-41 silica modified with amino groups by double functionalization. Microporous Mesoporous Mater. 2015, 209, 165–171. [Google Scholar] [CrossRef]
  117. Wang, X.; Guo, Q.; Zhao, J.; Chen, L. Mixed amine-modified MCM-41 sorbents for CO2 capture. Int. J. Greenh. Gas Control 2015, 37, 90–98. [Google Scholar] [CrossRef]
  118. Jung, H.; Lee, C.H.; Jeon, S.; Jo, D.H.; Huh, J.; Kim, S.H. Effect of amine double-functionalization on CO2 adsorption behaviors of silica gel-supported adsorbents. Adsorption 2016, 22, 1137–1146. [Google Scholar] [CrossRef]
  119. Fillerup, E.; Zhang, Z.; Peduzzi, E.; Wang, D.; Guo, J.; Ma, X.; Wang, X.; Song, C. CO Capture from Flue Gas Using Solid Molecular Basket Sorbents; The Pennsylvania State University: State College, PA, USA, 2012. [Google Scholar]
  120. Gargiulo, N.; Peluso, A.; Aprea, P.; Pepe, F.; Caputo, D. CO2 adsorption on polyethylenimine-functionalized SBA-15 mesoporous silica: Isotherms and modeling. J. Chem. Eng. Data 2014, 59, 896–902. [Google Scholar] [CrossRef]
  121. Heydari-Gorji, A.; Sayari, A. Thermal, oxidative, and CO2-induced degradation of supported polyethylenimine adsorbents. Ind. Eng. Chem. Res. 2012, 51, 6887–6894. [Google Scholar] [CrossRef]
  122. Kuwahara, Y.; Kang, D.Y.; Copeland, J.R.; Brunelli, N.A.; Didas, S.A.; Bollini, P.; Sievers, C.; Kamegawa, T.; Yamashita, H.; Jones, C.W. Dramatic enhancement of CO2 uptake by poly (ethyleneimine) using zirconosilicate supports. J. Am. Chem. Soc. 2012, 134, 10757–10760. [Google Scholar] [CrossRef]
  123. Kishor, R.; Ghoshal, A.K. Amine-modified mesoporous silica for CO2 adsorption: The role of structural parameters. Ind. Eng. Chem. Res. 2017, 56, 6078–6087. [Google Scholar] [CrossRef]
  124. Ko, Y.G.; Lee, H.J.; Oh, H.C.; Choi, U.S. Amines immobilized double-walled silica nanotubes for CO2 capture. J. Hazard. Mater. 2013, 250, 53–60. [Google Scholar] [CrossRef] [PubMed]
  125. Choi, S.; Drese, J.H.; Eisenberger, P.M.; Jones, C.W. Application of amine-tethered solid sorbents for direct CO2 capture from the ambient air. Environ. Sci. Technol. 2011, 45, 2420–2427. [Google Scholar] [CrossRef] [PubMed]
  126. Guo, X.; Ding, L.; Kanamori, K.; Nakanishi, K.; Yang, H. Functionalization of hierarchically porous silica monoliths with polyethyleneimine (PEI) for CO2 adsorption. Microporous Mesoporous Mater. 2017, 245, 51–57. [Google Scholar] [CrossRef]
  127. Liu, Z.; Teng, Y.; Zhang, K.; Chen, H.; Yang, Y. CO2 adsorption performance of different amine-based siliceous MCM-41 materials. J. Energy Chem. 2015, 24, 322–330. [Google Scholar] [CrossRef]
  128. de Carvalho, L.S.; Silva, E.; Andrade, J.C.; Silva, J.A.; Urbina, M.; Nascimento, P.F.; Carvalho, F.; Ruiz, J.A. Low-cost mesoporous adsorbents amines-impregnated for CO2 capture. Adsorption 2015, 21, 597–609. [Google Scholar] [CrossRef]
  129. López-Aranguren, P.; Builes, S.; Fraile, J.; López-Periago, A.; Vega, L.F.; Domingo, C. Hybrid aminopolymer–silica materials for efficient CO2 adsorption. RSC Adv. 2015, 5, 104943–104953. [Google Scholar] [CrossRef]
  130. Mello, M.R.; Phanon, D.; Silveira, G.Q.; Llewellyn, P.L.; Ronconi, C.M. Amine-modified MCM-41 mesoporous silica for carbon dioxide capture. Microporous Mesoporous Mater. 2011, 143, 174–179. [Google Scholar] [CrossRef]
  131. Ahmed, S.; Ramli, A.; Yusup, S.; Farooq, M. Adsorption behavior of tetraethylenepentamine-functionalized Si-MCM-41 for CO2 adsorption. Chem. Eng. Res. Des. 2017, 122, 33–42. [Google Scholar] [CrossRef]
  132. Niu, M.; Yang, H.; Zhang, X.; Wang, Y.; Tang, A. Amine-impregnated mesoporous silica nanotube as an emerging nanocomposite for CO2 capture. ACS Appl. Mater. Interfaces 2016, 8, 17312–17320. [Google Scholar] [CrossRef]
  133. Ullah, R.; Atilhan, M.; Aparicio, S.; Canlier, A.; Yavuz, C.T. Insights of CO2 adsorption performance of amine impregnated mesoporous silica (SBA-15) at wide range pressure and temperature conditions. Int. J. Greenh. Gas Control 2015, 43, 22–32. [Google Scholar] [CrossRef]
  134. Sánchez-Vicente, Y.; Stevens, L.A.; Pando, C.; Torralvo, M.J.; Snape, C.E.; Drage, T.C.; Cabañas, A. A new sustainable route in supercritical CO2 to functionalize silica SBA-15 with 3-aminopropyltrimethoxysilane as material for carbon capture. J. Chem. Eng. 2015, 264, 886–898. [Google Scholar] [CrossRef]
  135. Zhao, A.; Samanta, A.; Sarkar, P.; Gupta, R. Carbon dioxide adsorption on amine-impregnated mesoporous SBA-15 sorbents: Experimental and kinetics study. Ind. Eng. Chem. Res. 2013, 52, 6480–6491. [Google Scholar] [CrossRef]
  136. Gholami, M.; Talaie, M.R.; Aghamiri, S.F. Direct synthesis of bi-modal porous structure MCM-41 and its application in CO2 capturing through amine-grafting. Korean J. Chem. Eng. 2014, 31, 322–326. [Google Scholar] [CrossRef]
  137. Loganathan, S.; Tikmani, M.; Ghoshal, A.K. Novel pore-expanded MCM-41 for CO2 capture: Synthesis and characterization. Langmuir 2013, 29, 3491–3499. [Google Scholar] [CrossRef]
  138. Kassab, H.; Maksoud, M.; Aguado, S.; Pera-Titus, M.; Albela, B.; Bonneviot, L. Polyethylenimine covalently grafted on mesostructured porous silica for CO2 capture. RSC Adv. 2012, 2, 2508–2516. [Google Scholar] [CrossRef]
  139. Teng, Y.; Li, L.; Xu, G.; Zhang, K.; Li, K. Promoting effect of inorganic alkali on carbon dioxide adsorption in amine-modified MCM-41. Energies 2016, 9, 667. [Google Scholar] [CrossRef] [Green Version]
  140. Sharma, P.; Seong, J.K.; Jung, Y.H.; Choi, S.H.; Park, S.D.; Yoon, Y.I.; Baek, I.H. Amine modified and pelletized mesoporous materials: Synthesis, textural–mechanical characterization and application in adsorptive separation of carbondioxide. Powder Technol. 2012, 219, 86–98. [Google Scholar] [CrossRef]
  141. Yan, X.; Komarneni, S.; Yan, Z. CO2 adsorption on Santa Barbara Amorphous-15 (SBA-15) and amine-modified Santa Barbara Amorphous-15 (SBA-15) with and without controlled microporosity. J. Colloid Interface Sci. 2013, 390, 217–224. [Google Scholar] [CrossRef]
  142. Sanz, R.; Calleja, G.; Arencibia, A.; Sanz-Pérez, E.S. Amino functionalized mesostructured SBA-15 silica for CO2 capture: Exploring the relation between the adsorption capacity and the distribution of amino groups by TEM. Microporous Mesoporous Mater. 2012, 158, 309–317. [Google Scholar] [CrossRef]
  143. Rao, N.; Wang, M.; Shang, Z.; Hou, Y.; Fan, G.; Li, J. CO2 adsorption by amine-functionalized MCM-41: A comparison between impregnation and grafting modification methods. Energy Fuels 2018, 32, 670–677. [Google Scholar] [CrossRef]
  144. Sim, K.; Lee, N.; Kim, J.; Cho, E.B.; Gunathilake, C.; Jaroniec, M. CO2 adsorption on amine-functionalized periodic mesoporous benzenesilicas. ACS Appl. Mater. Interfaces 2015, 7, 6792–6802. [Google Scholar] [CrossRef] [PubMed]
  145. Dassanayake, R.S.; Gunathilake, C.; Dassanayake, A.C.; Abidi, N.; Jaroniec, M. Amidoxime-functionalized nanocrystalline cellulose–mesoporous silica composites for carbon dioxide sorption at ambient and elevated temperatures. J. Mater. Chem. A 2017, 5, 7462–7473. [Google Scholar] [CrossRef]
  146. Gunathilake, C.; Dassanayake, R.S.; Abidi, N.; Jaroniec, M. Amidoxime-functionalized microcrystalline cellulose–mesoporous silica composites for carbon dioxide sorption at elevated temperatures. J. Mater. Chem. A 2016, 4, 4808–4819. [Google Scholar] [CrossRef]
  147. Gunathilake, C.; Jaroniec, M. Mesoporous calcium oxide–silica and magnesium oxide–silica composites for CO2 capture at ambient and elevated temperatures. J. Mater. Chem. A 2016, 4, 10914–10924. [Google Scholar] [CrossRef]
  148. Gunathilake, C.; Dassanayake, R.S.; Kalpage, C.S.; Jaroniec, M. Development of Alumina–Mesoporous Organosilica Hybrid Materials for Carbon Dioxide Adsorption at 25 °C. Materials 2018, 11, 2301. [Google Scholar] [CrossRef] [Green Version]
  149. Gunathilake, C.; Gangoda, M.; Jaroniec, M. Mesoporous isocyanurate-containing organosilica–alumina composites and their thermal treatment in nitrogen for carbon dioxide sorption at elevated temperatures. J. Mater. Chem. A 2013, 1, 8244–8252. [Google Scholar] [CrossRef]
  150. Gunathilake, C.; Jaroniec, M. Mesoporous alumina–zirconia–organosilica composites for CO2 capture at ambient and elevated temperatures. J. Mater. Chem. A 2015, 3, 2707–2716. [Google Scholar] [CrossRef]
  151. Choi, W.; Min, K.; Kim, C.; Ko, Y.S.; Jeon, J.W.; Seo, H.; Park, Y.K.; Choi, M. Epoxide-functionalization of polyethyleneimine for synthesis of stable carbon dioxide adsorbent in temperature swing adsorption. Nat. Commun. 2016, 7, 12640. [Google Scholar] [CrossRef] [Green Version]
  152. Hu, Y.; Liu, W.; Yang, Y.; Qu, M.; Li, H. CO2 capture by Li4SiO4 sorbents and their applications: Current developments and new trends. J. Chem. Eng. 2019, 359, 604–625. [Google Scholar] [CrossRef]
  153. Wang, L.; Yang, R.T. Increasing selective CO2 adsorption on amine-grafted SBA-15 by increasing silanol density. J. Phys. Chem. C 2011, 115, 21264–21272. [Google Scholar] [CrossRef]
  154. Rafigh, S.M.; Heydarinasab, A. Mesoporous chitosan–SiO2 nanoparticles: Synthesis, characterization, and CO2 adsorption capacity. ACS Sustain. Chem. Eng. 2017, 5, 10379–10386. [Google Scholar] [CrossRef]
  155. Yang, J.; Li, J.; Wang, W.; Li, L.; Li, J. Adsorption of CO2, CH4, and N2 on 8-, 10-, and 12-membered ring hydrophobic microporous high-silica zeolites: DDR, silicalite-1, and beta. Ind. Eng. Chem. Res. 2013, 52, 17856–17864. [Google Scholar] [CrossRef]
  156. Zohdi, S.; Anbia, M.; Salehi, S. Improved CO2 adsorption capacity and CO2/CH4 and CO2/N2 selectivity in novel hollow silica particles by modification with multi-walled carbon nanotubes containing amine groups. Polyhedron 2019, 166, 175–185. [Google Scholar] [CrossRef]
  157. Fernandes, J.; Fernandes, A.C.; Echeverría, J.C.; Moriones, P.; Garrido, J.J.; Pires, J. Adsorption of gases and vapours in silica based xerogels. Colloids Surf. A Physicochem. Eng. Asp. 2019, 561, 128–135. [Google Scholar] [CrossRef]
  158. Liu, X.; Gao, F.; Xu, J.; Zhou, L.; Liu, H.; Hu, J. Zeolite@ Mesoporous silica-supported-amine hybrids for the capture of CO2 in the presence of water. Microporous Mesoporous Mater. 2016, 222, 113–119. [Google Scholar] [CrossRef]
  159. Wei, L.; Jing, Y.; Gao, Z.; Wang, Y. Development of a pentaethylenehexamine-modified solid support adsorbent for CO2 capture from model flue gas. Chin. J. Chem. Eng. 2015, 23, 366–371. [Google Scholar] [CrossRef]
  160. Yan, X.; Zhang, L.; Zhang, Y.; Yang, G.; Yan, Z. Amine-modified SBA-15: Effect of pore structure on the performance for CO2 capture. Ind. Eng. Chem. Res. 2011, 50, 3220–3226. [Google Scholar] [CrossRef]
  161. Zhang, X.; Qin, H.; Zheng, X.; Wu, W. Development of efficient amine-modified mesoporous silica SBA-15 for CO2 capture. Mater. Res. Bull. 2013, 48, 3981–3986. [Google Scholar] [CrossRef]
  162. Subagyono, D.J.; Marshall, M.; Knowles, G.P.; Chaffee, A.L. CO2 adsorption by amine modified siliceous mesostructured cellular foam (MCF) in humidified gas. Microporous Mesoporous Mater. 2014, 186, 84–93. [Google Scholar] [CrossRef]
  163. Ma, J.; Liu, Q.; Chen, D.; Wen, S.; Wang, T. CO2 adsorption on amine-modified mesoporous silicas. J. Porous Mater. 2014, 21, 859–867. [Google Scholar] [CrossRef]
  164. Zhang, H.; Goeppert, A.; Prakash, G.S.; Olah, G. Applicability of linear polyethylenimine supported on nano-silica for the adsorption of CO2 from various sources including dry air. RSC Adv. 2015, 5, 52550–52562. [Google Scholar] [CrossRef]
  165. Wang, X.; Chen, L.; Guo, Q. Development of hybrid amine-functionalized MCM-41 sorbents for CO2 capture. Chem. Eng. J. 2015, 260, 573–581. [Google Scholar] [CrossRef]
  166. Liu, J.L.; Lin, R.B. Structural properties and reactivities of amino-modified silica fume solid sorbents for low-temperature CO2 capture. Powder Technol. 2013, 241, 188–195. [Google Scholar] [CrossRef]
  167. Li, K.; Jiang, J.; Tian, S.; Yan, F.; Chen, X. Polyethyleneimine–nano silica composites: A low-cost and promising adsorbent for CO2 capture. J. Mater. Chem. A 2015, 3, 2166–2175. [Google Scholar] [CrossRef]
  168. Li, K.; Jiang, J.; Yan, F.; Tian, S.; Chen, X. The influence of polyethyleneimine type and molecular weight on the CO2 capture performance of PEI-nano silica adsorbents. Appl. Energy 2014, 136, 750–755. [Google Scholar] [CrossRef]
  169. Quang, D.V.; Hatton, T.A.; Abu-Zahra, M.R. Thermally stable amine-grafted adsorbent prepared by impregnating 3-aminopropyltriethoxysilane on mesoporous silica for CO2 capture. Ind. Eng. Chem. Res. 2016, 55, 7842–7852. [Google Scholar] [CrossRef]
  170. Zeng, W.; Bai, H. High-performance CO2 capture on amine-functionalized hierarchically porous silica nanoparticles prepared by a simple template-free method. Adsorption 2016, 22, 117–127. [Google Scholar] [CrossRef]
  171. Linneen, N.N.; Pfeffer, R.; Lin, Y.S. CO2 adsorption performance for amine grafted particulate silica aerogels. Chem. Eng. J. 2014, 254, 190–197. [Google Scholar] [CrossRef]
  172. Le, Y.; Guo, D.; Cheng, B.; Yu, J. Amine-functionalized monodispersed porous silica microspheres with enhanced CO2 adsorption performance and good cyclic stability. J. Colloid Interface Sci. 2013, 408, 173–180. [Google Scholar] [CrossRef]
  173. Singh, B.; Polshettiwar, V. Design of CO2 sorbents using functionalized fibrous nanosilica (KCC-1): Insights into the effect of the silica morphology (KCC-1 vs. MCM-41). J. Mater. Chem. A 2016, 4, 7005–7019. [Google Scholar] [CrossRef]
  174. Zhang, L.; Zhan, N.; Jin, Q.; Liu, H.; Hu, J. Impregnation of polyethylenimine in mesoporous multilamellar silica vesicles for CO2 capture: A kinetic study. Ind. Eng. Chem. Res. 2016, 55, 5885–5891. [Google Scholar] [CrossRef]
  175. Zhou, L.; Fan, J.; Cui, G.; Shang, X.; Tang, Q.; Wang, J.; Fan, M. Highly efficient and reversible CO2 adsorption by amine-grafted platelet SBA-15 with expanded pore diameters and short mesochannels. Green Chem. 2014, 16, 4009–4016. [Google Scholar] [CrossRef]
  176. Mittal, N.; Samanta, A.; Sarkar, P.; Gupta, R. Postcombustion CO2 capture using N-(3-trimethoxysilylpropyl) diethylenetriamine-grafted solid adsorbent. Energy Sci. Eng. 2015, 3, 207–220. [Google Scholar] [CrossRef]
  177. Yao, M.; Dong, Y.; Feng, X.; Hu, X.; Jia, A.; Xie, G.; Hu, G.; Lu, J.; Luo, M.; Fan, M. The effect of post-processing conditions on aminosilane functionalizaiton of mesocellular silica foam for post-combustion CO2 capture. Fuel 2014, 123, 66–72. [Google Scholar] [CrossRef]
  178. Ko, Y.G.; Lee, H.J.; Kim, J.Y.; Choi, U.S. Hierarchically porous aminosilica monolith as a CO2 adsorbent. ACS Appl. Mater. Interfaces 2014, 6, 12988–12996. [Google Scholar] [CrossRef]
  179. Shi, X.; Xiao, H.; Azarabadi, H.; Song, J.; Wu, X.; Chen, X.; Lackner, K.S. Sorbents for the direct capture of CO2 from ambient air. Angew. Chem. Int. Ed. 2020, 59, 6984–7006. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of post-combustion technology (Reprinted with permission from Osman et al. [1]).
Figure 1. Schematic representation of post-combustion technology (Reprinted with permission from Osman et al. [1]).
Nanomaterials 13 02050 g001
Figure 2. Reaction mechanism of CO2 capture into MEA solution (Reprinted with permission from Lv et al. [31]).
Figure 2. Reaction mechanism of CO2 capture into MEA solution (Reprinted with permission from Lv et al. [31]).
Nanomaterials 13 02050 g002
Figure 3. Schematic representation of (a) physisorption and (b) chemisorption.
Figure 3. Schematic representation of (a) physisorption and (b) chemisorption.
Nanomaterials 13 02050 g003
Figure 4. The different types of adsorption processes.
Figure 4. The different types of adsorption processes.
Nanomaterials 13 02050 g004
Figure 6. Mechanism for the synthesis of mesoporous silica using block copolymer (Re-printed with permission from Gunathilake et al. [88]).
Figure 6. Mechanism for the synthesis of mesoporous silica using block copolymer (Re-printed with permission from Gunathilake et al. [88]).
Nanomaterials 13 02050 g006
Figure 7. Amino silane- and polymer-containing amino groups used in the functionalization of mesoporous silicas.
Figure 7. Amino silane- and polymer-containing amino groups used in the functionalization of mesoporous silicas.
Nanomaterials 13 02050 g007aNanomaterials 13 02050 g007b
Figure 9. Schematic representation of the covalent bonding through the alkyl-silyl linkages and formation of carbamates (Reprinted with permission from Nigar et al. [99]).
Figure 9. Schematic representation of the covalent bonding through the alkyl-silyl linkages and formation of carbamates (Reprinted with permission from Nigar et al. [99]).
Nanomaterials 13 02050 g009
Table 1. Different approaches used in different countries in order to reduce the CO2 emissions [4].
Table 1. Different approaches used in different countries in order to reduce the CO2 emissions [4].
Type of ApproachDetails
Improve energy efficiency and promote energy conservation
  • This approach is mainly used in commercial and industrial buildings
  • It shows mainly 10–20% of energy saving
  • It shows extensive capital investment for installation
Increase in usage of low carbon or clean fuels such as natural gas, hydrogen or nuclear power; Substitution for Power generation
  • Natural gas emits 40–50% less CO2 than coal
  • Main advantages of this method are high efficiency and cleaner exhaust gas
  • Main disadvantage is the high cost
Deploy renewable energy
  • The renewable energy sources include solar, wind, hydropower, geothermal, oceanic energy and bioenergy
  • This method emits low green house and toxic gases
  • The main limitation is high cost and geographic distribution of the available resources
CO2 capture and storage
  • This method is applicable for large CO2 point emission sources
  • It can reduce vast amount of CO2 with capture efficiency of 48%
Table 2. Comparison of different post-combustion capture technologies for CO2 capture.
Table 2. Comparison of different post-combustion capture technologies for CO2 capture.
TechnologyTypesExamplesEfficiency (%)AdvantagesDisadvantagesRef.
AbsorptionChemicalAmines
Caustics
>90
  • Ability to regenerate
  • Established method
  • Very flexible
  • Reacts rapidly
  • High absorption capacities
  • High energy requirement for regeneration
  • Environmental problems
  • High boiling point
  • Equipment corrosion
[21,22]
PhysicalSelexol
Rectisol
fluorinated
solvents
AdsorptionChemicalMetal Oxides
Si based materials
>85
  • Recyclable
  • Cost effective
  • High stability
  • Adjustable catalytic site and pore sizes
  • Low energy consumption
  • Suitable for separating CO2 from dilute streams
  • High energy cost
  • Limited to process feed rates
  • Loss of material and pressure drop
  • Decreased catalytic efficiency
  • Low adsorption capacities
[6,21]
PhysicalCarbons
Zeolites
Si based materials
Membrane-based technologiesOrganic
Cellulose derivatives
Polyamides
Polyphenyleneoxide,
Polydimethylsiloxane
>80
  • Simple device
  • Easy production process and process flow scheme
  • Low energy consumption
  • No phase changes
  • Capable of maintaining the membrane structure
  • Requires a high-cost module and support materials
  • Not suitable for large volumes of emission gases
  • Reduced selectivity and separation
  • Pressure drops across the membrane
  • Less durability
[6,21]
InorganicMetallic
Ceramics
Cryogenic distillation
  • Low capital investment
  • High reliability
  • Recovery with high purity of CO2
  • Liquid CO2 production
  • Not requiring solvents or other components
  • Easily scalable to industrial-scale applications
  • High energy consumption
[6,21,23]
Table 4. Advantage and disadvantages of non-carbonaceous adsorbents.
Table 4. Advantage and disadvantages of non-carbonaceous adsorbents.
Material TypesExamplesAdvantagesDisadvantages
Pours silica
materials
M41S
SBA-n
AMS
  • High specific surface area, Pore volume, and good thermal and mechanical properties
  • High molecular diffusion resistance
  • Decreased adsorption capacity at high temperature [42]
ZeolitesNaY
13X
  • Low production cost
  • Large micropores/mesopores
  • Medium CO2 adsorption capacity at room temperature
  • Low CO2 adsorption capacity
  • Moisture-sensitivity
  • High energy consumption [6,43]
Metal organic
frameworks (MOFs)
M-MOF-74
IRMOF-6
USO-2-Ni
Zn4O (BDC)3
(MOF-5)
USO-1-Al (MIL-53)
  • Large specific surface area
  • Ease of controlling pore sizes
  • High selectivity of CO2
  • Low CO2 adsorption capacity at the partial pressure
  • High production cost
  • Complicated synthesis process
  • Moisture-sensitivity
  • Unstable at high temperature [6]
Alkali-based dry
adsorbents
  • Possible adsorption and desorption at a low temperature and wet conditions
  • Low adsorption capability (3–11 wt%)
  • High-temperature reactions
  • Requires high temperatures during desorption
  • Complicated operation [6]
Metal oxides-based
adsorbents
CaO, MgO
  • Dry chemical adsorbents
  • Adsorption/desorption at medium to high temperatures
  • High energy consumption
  • High cost for regeneration
  • Complicated process [6]
Table 5. Comparison between chemisorption and physisorption.
Table 5. Comparison between chemisorption and physisorption.
ChemisorptionPhysisorption
Description
  • Chemical reaction occurs between the solid sorbents and CO2
  • Depends on the physical properties of CO2 and the ability to engage in noncovalent interactions with the solid sorbent
Chemical Bonding
  • Covalent Bonding-Occur between functional groups and CO2 in the surface
  • Week Vander-walls forces-London and Dispersion forces, Occur inside pore walls
Advantages
  • High selectivity
  • Low recycling energy requirements
  • High working capacity
  • High selectivity even in wet environments
  • Fast
Disadvantages
  • High energy required for recycling and the breakage of the chemical bonds
  • Slow reactivity
  • Poor selectivity in binary or mixed gas applications
References[55,56][25,57,58]
Table 8. Summary of gas selectivity values of previous studies performed for porous SiO2.
Table 8. Summary of gas selectivity values of previous studies performed for porous SiO2.
Porous SiO2 MaterialGas MixtureSelectivity ValuePressure (Bar)Temperature
(°C)
Reference
PEI-MCM-41CO2, N2 and H225.561100[93]
SBA-15CO2/N2123125[154]
SBA-15 (calcination)CO2/N255 125[154]
Mesoporous chitosan−SiO2 nanoparticles-15.46125[155]
hydrophobic microporous high-silica zeolitesCH4:N2 = 50%:50%36.5125[156]
Hollow silica spherical particles (HSSP)CO2/N28.5425[157]
microporous silica xerogelCO2/CH460625[158]
Silica based xerogelsC2H4/C2H620625[158]
Table 9. Summary of stability of silica-based adsorbent studied in past performance capacity.
Table 9. Summary of stability of silica-based adsorbent studied in past performance capacity.
Synthesis MethodType of Silica-Based SorbentAmine TypeRegeneration Condition Stability PerformanceReferences
Temperature (°C)Types of Gas FlowNo. of Cycles (Cyclic Runs)Capacity
Loss (%)
ImpregnatedMCM-41 PEHA100N2 15 Less than 1 [159]
MCM-41 TEPA + AMP100N2 for 60 min 15 4.32 [117]
SBA-15 PEI-linear 100Ar 1213.5[160]
SBA-15 Acrylonitrile-modified TEPA100N2121.1 [161]
HMS PEI-linear75N2 for 100 min 41.6[110]
MCFPEI-branched115Ar for 20 min 1032[162]
MCFPEI100H2105 [163]
MCFGuanidinylated poly(allylamine) 120He517[52]
Fumed silica PEI-linear55N2 for 15 min180Stable[164]
MCM-41 TEPA100N2103.43[165]
Silica fumeDiisopropanolamine50N2107[166]
Nano-SiO2 PEI-branched120N23010.5[167]
Nano-SiO2 PEI-branched120N23019.4[168]
Mesoporous-SiO2APTS120 Air for 30 min114.3[169]
Porous SiO2PEI100N2 for 30 min205[170]
Silica aerogel TEPA75Ar for 20 min103.9[171]
Porous SiO2TEPA75 He for 20 min102[172]
SNTPEI 110N2 for 40 min103.3[132]
KCC-1-SiO2TEPA 110N2211.2[173]
Mesoporous
multilamellar SiO2
PEI110N2103.7[174]
Silica aerogel TEPA80Ar for 30 min10012[173]
Mesoporous
SiO2
DEA90N21012[169]
GraftingSBA-15 AP90Vacuum101[175]
SBA-15DEAPTMS120 N2 for 10 min1007.2[176]
MCM-482-[2-(3-trimethoxysilyl propylamino)
ethylamino] ethylamine
-N220Stable[98]
KIT-6 APTES120He10Stable[97]
MCFTRI150 N2 for 30 min51.9[177]
HMSAPTS110 N2 for 180 min3Less than 1[178]
MCM-41APTS105 N2 for 90 min10Stable[115]
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

Amaraweera, S.M.; Gunathilake, C.A.; Gunawardene, O.H.P.; Dassanayake, R.S.; Cho, E.-B.; Du, Y. Carbon Capture Using Porous Silica Materials. Nanomaterials 2023, 13, 2050. https://doi.org/10.3390/nano13142050

AMA Style

Amaraweera SM, Gunathilake CA, Gunawardene OHP, Dassanayake RS, Cho E-B, Du Y. Carbon Capture Using Porous Silica Materials. Nanomaterials. 2023; 13(14):2050. https://doi.org/10.3390/nano13142050

Chicago/Turabian Style

Amaraweera, Sumedha M., Chamila A. Gunathilake, Oneesha H. P. Gunawardene, Rohan S. Dassanayake, Eun-Bum Cho, and Yanhai Du. 2023. "Carbon Capture Using Porous Silica Materials" Nanomaterials 13, no. 14: 2050. https://doi.org/10.3390/nano13142050

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

Article Metrics

Back to TopTop