- 1.1 Introduction
- 1.2 Porous Carbons
- 1.2.1 Chemical Activation
- 1.2.2 Physical Activation
- 1.2.3 Metal Ion Activation
- 1.2.4 Templating Method
- 1.2.5 Combined Method of Templating and Activation
- 1.3 Graphene-based Porous Materials
- 1.3.1 Graphene-based Adsorbents by Chemical Activation
- 1.3.2 Graphene-based Adsorbents by Physical Activation
- 1.3.3 Graphene-based Adsorbents by Other Techniques
- 1.4 Carbon Nanotubes
- 1.5 Carbon-based Hybrid Adsorbents
- 1.5.1 Carbon–Organic Hybrid Adsorbents
- 1.5.2 Carbon–Inorganic Hybrid Adsorbents
- 1.6 Effect of Carbon Structure on CO2 Adsorption
- 1.6.1 Pore Size Effect
- 1.6.2 Surface Chemistry Effect on CO2 Adsorption
- 1.7 Summary and Outlook
- References
CHAPTER 1: Carbon-based CO2 Adsorbents
-
Published:22 Oct 2018
-
Series: Inorganic Materials Series
J. Zhou, X. Wang, and W. Xing, in Post-combustion Carbon Dioxide Capture Materials, ed. Q. Wang, The Royal Society of Chemistry, 2018, pp. 1-75.
Download citation file:
Carbon materials have been considered to be one of the most promising candidates for CO2 capture due to their specific features such as low price, high specific surface area, hydrophobic surface, excellent thermal and chemical stability, and low energy requirements for regeneration. This chapter aims to summarize the recent research progress made in developing carbon-based adsorbents for post-combustion CO2 capture. Specifically, this chapter provides overviews of (1) porous carbons, (2) graphene-based porous materials, (3) carbon nanotubes, (4) carbon-based hybrid sorbents, and (5) important factors (pore size and surface chemistry) influencing CO2 uptake over carbon adsorbents. Further, the future prospects of carbon-based adsorbents are briefly discussed.
1.1 Introduction
Carbon dioxide (CO2) has been recognized to be the biggest driver of global warming, which is one of the most serious problems that our world is facing.1–3 Furthermore, CO2 is an important source of C1 chemical engineering, and could be converted into high-value chemical products via chemical,2,3 photochemical,4,5 or electrochemical processes.6,7 However, the efficient capture of CO2 is essential for these processes. So, there is an urgent need to develop CO2 capture and storage (CCS) technologies. The basic concept of CCS is to capture CO2 from emissions without releasing it into the atmosphere. CCS can be classified as post-combustion, pre-combustion, and oxy-fuel combustion technologies. Among the current CCS technologies, post-combustion capture, a technology for capturing CO2 from post-combustion emission gases (e.g., flue gas from power plants) is the most easily applied technology for existing emission sources.
In general, post-combustion capture technologies include chemical absorption, dry adsorption, membrane-based technologies, and cryogenic technologies. Currently, chemical absorption is the most applicable technology for CO2 capture in power plants, but this technology suffers from several drawbacks. The biggest challenge in applying a chemical absorption process for post-combustion is how to reduce the heat of regeneration. Another problem is the release of hazardous byproducts. For these reasons, dry adsorption using solid adsorbents is considered to be promising for the capture of post-combustion CO2.
The dry adsorption technique is a process of selective adsorption of CO2 from post-combustion gases using solid adsorbents, which has advantages such as a simple device, easy operation, it is environmentally friendly and has a high energy efficiency. When evaluating solid adsorbents, it is important to consider their surface area, apparent density, pore size and volume, feasibility of regeneration, stability, abundance and sustainability. CO2 capture by solid adsorbents mainly relies on the mechanism of physical adsorption that is also interfered with by some weak interactions between CO2 and the adsorbent's surface (i.e. hydrogen bonding or electric quadrupole interactions). Due to the main contribution of van der Waals forces to the physical adsorption, materials that possess a developed microporous texture are preferred for CO2 capture. Nowadays, many kinds of solid adsorbent materials with porous textures, such as porous carbonaceous materials,8,9 zeolites,10 zeolitic imidazolate frameworks (ZIFs),11 metal–organic frameworks (MOFs),12 covalent organic frameworks (COFs),13 and porous coordination polymers (PCPs),14 have been investigated, and show excellent CO2 capture performances.
Among these solid adsorbents, porous carbonaceous materials have been studied intensively because of their desirable physical and chemical properties, such as low cost, variety of form (powder, fibers, aerogels, composites, sheets, monoliths, tubes, etc.), ease of processability, controllable porosity (adjustable pore size and its distribution, high specific surface area and pore volume), and tailored surface chemistry (O, N, S, P, F or other heteroatom doping). They also possess some other advantages, particularly for adsorption applications: (1) carbon materials have excellent stability especially in hot and humid environments; (2) gas sorption on carbon materials is not moisture-sensitive because the surface is usually hydrophobic; (3) the energy consumption of regeneration is low due to the nature of physical adsorption; (4) the adsorption/desorption temperatures are always below 373 K; (5) these materials can be used at atmospheric pressure.
In this chapter, we summarize the recent research progress made in developing carbon-based sorbents for post-combustion CO2 capture. Specifically, this chapter will provide overviews of (1) porous carbons, (2) graphene-based porous materials, (3) carbon nanotubes, (4) carbon-based hybrid sorbents, and (5) important factors influencing CO2 uptake over carbon adsorbents.
1.2 Porous Carbons
Porous carbons have been extensively studied in the field of CO2 capture. In order to control the pore structure in carbon materials, a variety of preparation methods have been developed and certain successes have been achieved. Herein, we summarize typical preparation methods for porous carbon adsorbents, including chemical activation, physical activation, metal ion activation, templating methods, and the combined method of templating and activation. In each section, synthesis principles, carbon precursors, pore structures, as well as their CO2 adsorption performance, are discussed.
1.2.1 Chemical Activation
Activated carbon is the oldest and most widely used carbon material. Generally, the production routes of activated carbons are divided into physical activation and chemical activation. In chemical activation, the carbon precursor is mixed/impregnated with an activating agent (such as KOH, H3PO4, ZnCl2, K2CO3, etc.), then the precursor is simultaneously carbonized and activated at an elevated temperature (from 400 to 900 °C) and under an inert atmosphere (usually N2 or Ar). In physical activation, the carbon precursor is usually pre-carbonized at temperatures over 500 °C in an inert atmosphere to remove non-carbon species, followed by etching by an oxidizing gas (such as CO2, steam, and air) at a higher temperature (from 700 to 1200 °C). Comparatively, chemical activation needs a lower temperature and shorter activation time, and generally results in a higher specific surface area and more uniform pore size distribution (PSD), while physical activation is simple and does not require chemical agents and repeated washing procedures to remove the inorganic residues after activation. The structure of activated carbons, containing the surface area, pore size and its distribution, and surface chemistry, etc., strongly depends on the activation conditions, activating agents as well as the carbon precursors used.
1.2.1.1 KOH as an Activating Agent
KOH is the most common activating agent in chemical activation. Jaroniec et al. treated a commercial carbon sorbent (Ambersorb 563) with five of the most commonly-used activating agents, including CO2, H2O, NH3, KOH, and ZnCl2, and compared their activating power for the evolution of microporosity responsible for CO2 adsorption.15 N2 adsorption analysis showed that the investigated activating agents enlarged microporosity and consequently surface area and pore volume of the carbons in the following order: KOH > CO2 > NH3 > H2O > ZnCl2. It was shown that KOH activation yielded the highest volume of micropores and small micropores. Besides, the CO2 uptake for the KOH-activated sample was the highest, indicating that KOH activation appears to be the most effective to obtain carbon adsorbents for CO2 capture.
However, the mechanism of KOH activation has not been totally understood due to the complexity of this process. In a previous review about KOH activated carbon materials for energy storage, Wang et al. suggested that KOH activation is a synergistic process of chemical activation, physical activation, and carbon lattice expansion by metallic K intercalation.16 Firstly, the potassium species serving as chemical activating reagents vigorously etch the carbon framework by the redox reactions shown in eqn (1.1)–(1.3). Secondly, H2O (from dehydration of carbon precursors or eqn (1.4)) and CO2 (eqn (1.5) and (1.6)) produced in situ in the activation system further develop the porosity through the gasification of carbon, namely physical activation (eqn (1.7) and (1.8)). Meanwhile, the produced metallic K intercalates into the lattices of the carbon matrix, responsible for both stabilization and widening of the interlayer spacing (Figure 1.1). After removal of the intercalated metallic K and other K compounds by washing, the expanded carbon lattices cannot return to their previous non-porous structure and thus create a narrow microporosity and large specific surface areas.
Activation mechanism by the penetration of metallic K into the carbon lattices. (a) Carbon lattices, (b) metallic K intercalated in the carbon lattices, (c) activated carbon.
Activation mechanism by the penetration of metallic K into the carbon lattices. (a) Carbon lattices, (b) metallic K intercalated in the carbon lattices, (c) activated carbon.
Generally, the raw materials of KOH activation could be classified into non-renewable fossil-based materials and renewable biomass resources. As shown in Table 1.1, various fossil-based precursors, such as petrol coke,17,18 pitch,19 and synthetic polymers,20–22 have been used as precursors for the preparation of porous carbon adsorbents. Wahby et al. prepared a series of carbon molecular sieves (CMS) from petroleum pitch using KOH as the activating agent. Depending on the type of petroleum residue and the conforming step applied, the prepared CMS possessed a well-defined pore size (0.35–0.7 nm), together with a very large surface area up to 3100 m2 g−1, thus exhibited a high CO2 adsorption capacity up to 4.09 mmol g−1 at 1 bar and 25 °C.23 After further optimizing the process parameters, including the nature of the petroleum residue, the KOH–pitch ratio, the mesophase content, the temperature and time of activation, the CO2 uptake could increase to 5.23 mmol g−1 at 1 bar and 25 °C.24 The activated carbons by the KOH activation of petroleum coke possessed high surface areas, over 3000 m2 g−1, and a high CO2 capacity, over 15 wt% at 1 bar.25 High-resolution analysis of N2 sorption isotherms concluded that the micropores smaller than 1 nm played a critical role in CO2 capture under ambient conditions due to the high-density filling of CO2 in these small pores.25 Using petroleum coke as the precursor, Yang et al. prepared N-doped porous carbons by combining ammoxidation with KOH activation.18 The sample prepared under mild conditions (a low temperature of 650 °C and a low KOH–precursor ratio of 2) showed the highest CO2 uptake of 4.57 mmol g−1 at 25 °C and 1 bar, while the CO2/N2 selectivity and CO2 heats of adsorption of the sorbent were 22 and 37 kJ mol−1, respectively. The high CO2 capture capacity was attributed to the synergetic effect of N-doping and high narrow microporosity, while the latter was suggested to contribute more.18
Porous carbons prepared by KOH activation of fossil-based resources for CO2 capture
Precursor . | Surface area (m2 g−1) . | Pore volume (cm3 g−1) . | Surface chemistry . | CO2 uptake (25 °C, 1 bar, mmol g−1) . | CO2–N2 selectivity . | Reference . |
---|---|---|---|---|---|---|
Petroleum pitch | 3100 | 1.40 | — | 4.09 | 2.8a | 23 |
Petroleum pitch | 2895 | 1.42 | — | 5.23 | — | 24 |
Coal–pitch mixture | 1044 | 0.50 | — | 4.00 | 5.94a | 26 |
Petroleum coke | 1445 | 0.58 | N-doped | 4.57 | 22c | 18 |
Petroleum coke | 1745 | — | — | 3.45 | — | 25 |
Phenolic resin | 2400 | 1.07 | — | 4.60 | — | 27 |
Waste ion-exchange resin | 828 | 0.42 | — | 1.85 | — | 28 |
Vinylidene chloride | 2151 | 0.90 | — | 3.97 | 12.59c | 29 |
Styrene divinylbenzene resin | 3870 | 2.07 | — | 4.75 | — | 30 |
Polypyrrole | 1700 | 0.88 | N-doped | 3.90 | — | 20 |
Polypyrrole–graphene composite | 1360 | 0.59 | N-doped | 4.30 | 16a | 31 |
Polyaniline | 1091 | 0.61 | N-doped | 4.30 | 8a | 21 |
Polyacrylonitrile | 2231 | 1.16 | N-doped | 4.50 | — | 22 |
Polyurethane | 1516 | 0.64 | N-doped | 4.33 | 12a | 32 |
Urea furfural resin | 1013 | 0.53 | N-doped | 4.70 | — | 33 |
1,3-Bis(cyanomethyl imidazolium) chloride | 1317 | 0.59 | N-doped | 5.39 | 62b | 34 |
Polyimine | 1561 | 0.75 | N-doped | 3.10 | 47a | 35,36 |
Polythiophene–graphene composite | 1567 | — | S-doped | 4.50 | 51c | 37 |
Melamine-doped phenolic resins | 1286 | 0.54 | N-doped | 4.04 | 52.9b | 38 |
Precursor . | Surface area (m2 g−1) . | Pore volume (cm3 g−1) . | Surface chemistry . | CO2 uptake (25 °C, 1 bar, mmol g−1) . | CO2–N2 selectivity . | Reference . |
---|---|---|---|---|---|---|
Petroleum pitch | 3100 | 1.40 | — | 4.09 | 2.8a | 23 |
Petroleum pitch | 2895 | 1.42 | — | 5.23 | — | 24 |
Coal–pitch mixture | 1044 | 0.50 | — | 4.00 | 5.94a | 26 |
Petroleum coke | 1445 | 0.58 | N-doped | 4.57 | 22c | 18 |
Petroleum coke | 1745 | — | — | 3.45 | — | 25 |
Phenolic resin | 2400 | 1.07 | — | 4.60 | — | 27 |
Waste ion-exchange resin | 828 | 0.42 | — | 1.85 | — | 28 |
Vinylidene chloride | 2151 | 0.90 | — | 3.97 | 12.59c | 29 |
Styrene divinylbenzene resin | 3870 | 2.07 | — | 4.75 | — | 30 |
Polypyrrole | 1700 | 0.88 | N-doped | 3.90 | — | 20 |
Polypyrrole–graphene composite | 1360 | 0.59 | N-doped | 4.30 | 16a | 31 |
Polyaniline | 1091 | 0.61 | N-doped | 4.30 | 8a | 21 |
Polyacrylonitrile | 2231 | 1.16 | N-doped | 4.50 | — | 22 |
Polyurethane | 1516 | 0.64 | N-doped | 4.33 | 12a | 32 |
Urea furfural resin | 1013 | 0.53 | N-doped | 4.70 | — | 33 |
1,3-Bis(cyanomethyl imidazolium) chloride | 1317 | 0.59 | N-doped | 5.39 | 62b | 34 |
Polyimine | 1561 | 0.75 | N-doped | 3.10 | 47a | 35,36 |
Polythiophene–graphene composite | 1567 | — | S-doped | 4.50 | 51c | 37 |
Melamine-doped phenolic resins | 1286 | 0.54 | N-doped | 4.04 | 52.9b | 38 |
The selectivity calculated by the ratio of CO2–N2 sorption capacity.
The selectivity based on ideal adsorbed solution theory (IAST).
The selectivity of Henry's law.
Besides petroleum coke and pitch, synthetic polymers, such as phenolic resins,27 poly(vinylidene chloride),29 styrene-divinylbenzene resin,30 polypyrrole,20 polyaniline,21 polyurethane,32 urea furfural resins,33 polyacrylonitrile,22 and polyimine,37,38 have also been widely used as precursors for the preparation of porous carbon adsorbents. For instance, Jaroniec et al. prepared activated carbon spheres by direct KOH activation of phenolic resin spheres obtained by a modified Stöber method.27 Due to the small micropore (<0.8 nm) and large specific surface areas, the prepared porous carbon spheres exhibited superior CO2 uptakes reaching 4.6 and 8.9 mmol g−1 at 23 and 0 °C under 1 bar, respectively.
Many studies have reported that N-doping could significantly improve the CO2 capture performance of porous carbons, especial CO2 uptake at low partial pressure and the selectivity of CO2-to-N2 (Table 1.1). Activation/carbonization of nitrogen-containing synthetic polymers is an important approach to preparing N-doped porous carbons. In 2010, Lu et al. reported a facile and rapid preparation of N-doped porous carbon monoliths using a basic amino acid as both the catalyst and nitrogen source.39 This monolithic carbon directly pyrolyzed at 500 °C exhibited a CO2 adsorption capacity of 3.13 mmol g−1 at room temperature. KOH activation of nitrogen-containing polymers could result in highly developed microporosity at the same time maintaining the N-doping. Sevilla et al. prepared highly porous N-doped carbon by using polypyrrole (PPy) as the carbon precursor and KOH as the activating agent.20 The mildly activated carbon (KOH–PPy = 2) showed a larger CO2 uptake than the severely activated ones (KOH–PPy = 4) as a result of two important characteristics for the mildly activated samples: (a) a high-level N-doping (up to 10.1 wt% N) identified as main pyridonic-N and a small proportion of pyridinic-N groups, and (b) a narrower micropore size. Moreover, these polypyrrole-based carbons showed a high adsorption rate for the capture of CO2, more than 95% of the CO2 being adsorbed in 2 min with a high kinetic CO2–N2 selectivity of 12. Other nitrogen-containing polymers, such as polyacrylonitrile,22 polyaniline,21 polyurethane,32 urea furfural resin,33 and polyimine,35,36 have also been used as precursors and the corresponding N-doped porous carbons showed CO2 uptakes of 4.50, 4.30, 4.33, 4.70, and 3.10 mmol g−1 at 25 °C and 1 bar, respectively.
From Table 1.1, we find that the CO2 uptakes for most of the activated carbons are lower than 5.0 mmol g−1. Recently, Sethia and Syari prepared a series of strictly microporous N-doped activated porous carbons by using a nitrogen-containing ionic liquid of 1,3-bis(cyanomethyl imidazolium) chloride as a precursor and KOH as the activating agent.34 The optimized material exhibited an extraordinary CO2 uptake of 5.39 mmol g−1 at 25 °C and 1 bar, which is the highest uptake reported so far for activated carbons. They concluded that both nitrogen content and ultramicropores played important roles in the CO2 capture, with the latter being predominant.34
Besides the N-doped porous carbons, S-doped porous carbons have also been prepared by the KOH activation of sulfur-containing synthetic polymers and are applied in CO2 capture. For instance, S-doped microporous carbon materials can be prepared by the KOH activation of polythiophene–graphene composite.37 This material displayed a high CO2 uptake of 4.5 mmol g−1 at 25 °C and 1 bar, as well as an impressive CO2 adsorption selectivity over N2 (51), CH4 (12), and H2 (214) based on the initial slopes method of Henry's law.40–42
Fossil-derived materials, such as petroleum coke, pitch, coal, and synthesized polymers are non-renewable and expensive. Considering the demand for a huge amount of solid CO2 sorbents, the development of low-cost carbon materials from renewable and sustainable sources is urgent.43 Biomass materials are renewable, easily available and very cheap, which are promising raw materials for the preparation of carbon adsorbents. As shown in Table 1.2, there have been over 40 kinds of biomass employed as precursors for porous carbon adsorbents, and most of them are obtained via chemical or physical activation, especially KOH activation. The specific surface areas and pore volumes of the biomass-based carbons were tuned with controlled carbonization and/or activation processes. The biomass used previously could be roughly divided into three categories: (1) polysaccharides, such as sucrose,44 cellulose,45 chitosan,46 starch,47,48 etc., (2) raw biomass or waste biomass, such as nutshells,49,50 wood residues,51 Jujun grass,52 Enteromorpha prolifera,53 sugar cane bagasse,54 granular bamboo,55 leaves,56 and others, and (3) microorganisms, including fungi57 and yeast.58
Biomass-based porous carbons prepared by KOH activation for CO2 capture
Biomass . | Activating agents . | Surface area (m2 g−1) . | Pore volume (cm3 g−1) . | Doping . | CO2 capacity at 25 °C and 1 bar (mmol g−1) . | Reference . |
---|---|---|---|---|---|---|
Sugarcane bagasse | None | 388 | 1.67 | 54 | ||
Palm shell | None | 753 | — | 2.3 | 60 | |
Chitosan–polybenzoxazine | None | 679 | 0.37 | 5.72 | 66 | |
Carboxymethylcellulose | Hydrothermal | 1063 | 0.70 | N-doped | 2.84 | 67 |
Palm shell | KOH | 1890 | 0.82 | 4.40 | 61 | |
Peanut shell | KOH | 1790 | 0.77 | N-doped | 1.54a | 50 |
Sawdust hydrochar | KOH | 2370 | 1.15 | 4.80 | 48 | |
Celtuce leaves | KOH | 3404 | 1.88 | 4.36 | 56 | |
Bamboo | KOH | 2332 | 1.00 | 4.50 | 55 | |
Enteromorpha prolifera | KOH | 481 | 0.37 | N-doped | 1.40b | 53 |
Coffee ground | KOH | 2785 | 1.36 | 2.81 | 68 | |
Chicken feather | KOH | 1610 | 0.86 | N-doped | 6.50c | 69 |
Coconut shell | KOH | 1535 | 0.60 | N-doped | 5.00 | 49 |
Jujun grass | KOH | 1512 | 0.74 | 5.00 | 52 | |
Gelatin | KOH | 1636 | 0.51 | N-doped | 3.80 | 47 |
Wheat flour | KOH | 1438 | 0.65 | 3.48 | 59 | |
Pine cone shell | KOH | 1033 | 4.80 | 70 | ||
Popcorn | KOH | 1489 | 0.71 | 4.60 | 71 | |
Microalgae | KOH | 1745 | 0.89 | N-doped | 4.30 | 64 |
Broussonetia papyrifera bark | KOH | 1759 | 0.92 | 4.45 | 72 | |
Coconut shell | KOH | 1593 | 0.68 | N-doped | 4.47 | 62 |
Burnt wood | KOH | 1554 | 0.7 | 5.00 | 73 | |
Oil palm shell | KOH | 1630 | 0.92 | 2.70 | 74 | |
Black locust | KOH | 2511 | 1.16 | N-doped | 5.05 | 75 |
Fish scales | KOH | 3206 | 2.29 | N-doped | 3.89 | 76 |
Fungi | KOH | 2264 | 0.92 | 3.30 | 57 | |
Yeast | KOH | 1348 | 0.67 | N-doped | 4.77 | 58 |
Sucrose | KOH | 1745 | 0.9 | N-doped | 4.30 | 44 |
Chitosan | KOH | 970 | 0.46 | N-doped | 5.00 | 77 |
Granular bamboo | KOH | 2641 | 1.25 | 4.5 | 55 | |
Cellulose | NaOH | 615 | 0.64 | N-doped | 4.99 | 45 |
Chitosan | K2CO3 | 1381 | 0.57 | N-doped | 3.86 | 46 |
Corncobs | NH3 | 1154 | 0.57 | N-doped | 2.81 | 78 |
Palm shell | NH3 | 889 | 0.47 | N-doped | 1.70 | 79 |
Banana peel | CO2 | 1426 | 0.83 | N-doped | 2.70 | 80 |
Coconut shell | CO2 | 1327 | 0.65 | 3.90 | 81 | |
Olive stones | CO2 | 1215 | 0.51 | 3.10 | 82 | |
Grass cuttings | CO2 | 841 | 0.38 | 1.45d | 83 | |
Cotton stalk | Steam | 439 | 0.19 | N-doped | 2.25 | 84 |
Biomass . | Activating agents . | Surface area (m2 g−1) . | Pore volume (cm3 g−1) . | Doping . | CO2 capacity at 25 °C and 1 bar (mmol g−1) . | Reference . |
---|---|---|---|---|---|---|
Sugarcane bagasse | None | 388 | 1.67 | 54 | ||
Palm shell | None | 753 | — | 2.3 | 60 | |
Chitosan–polybenzoxazine | None | 679 | 0.37 | 5.72 | 66 | |
Carboxymethylcellulose | Hydrothermal | 1063 | 0.70 | N-doped | 2.84 | 67 |
Palm shell | KOH | 1890 | 0.82 | 4.40 | 61 | |
Peanut shell | KOH | 1790 | 0.77 | N-doped | 1.54a | 50 |
Sawdust hydrochar | KOH | 2370 | 1.15 | 4.80 | 48 | |
Celtuce leaves | KOH | 3404 | 1.88 | 4.36 | 56 | |
Bamboo | KOH | 2332 | 1.00 | 4.50 | 55 | |
Enteromorpha prolifera | KOH | 481 | 0.37 | N-doped | 1.40b | 53 |
Coffee ground | KOH | 2785 | 1.36 | 2.81 | 68 | |
Chicken feather | KOH | 1610 | 0.86 | N-doped | 6.50c | 69 |
Coconut shell | KOH | 1535 | 0.60 | N-doped | 5.00 | 49 |
Jujun grass | KOH | 1512 | 0.74 | 5.00 | 52 | |
Gelatin | KOH | 1636 | 0.51 | N-doped | 3.80 | 47 |
Wheat flour | KOH | 1438 | 0.65 | 3.48 | 59 | |
Pine cone shell | KOH | 1033 | 4.80 | 70 | ||
Popcorn | KOH | 1489 | 0.71 | 4.60 | 71 | |
Microalgae | KOH | 1745 | 0.89 | N-doped | 4.30 | 64 |
Broussonetia papyrifera bark | KOH | 1759 | 0.92 | 4.45 | 72 | |
Coconut shell | KOH | 1593 | 0.68 | N-doped | 4.47 | 62 |
Burnt wood | KOH | 1554 | 0.7 | 5.00 | 73 | |
Oil palm shell | KOH | 1630 | 0.92 | 2.70 | 74 | |
Black locust | KOH | 2511 | 1.16 | N-doped | 5.05 | 75 |
Fish scales | KOH | 3206 | 2.29 | N-doped | 3.89 | 76 |
Fungi | KOH | 2264 | 0.92 | 3.30 | 57 | |
Yeast | KOH | 1348 | 0.67 | N-doped | 4.77 | 58 |
Sucrose | KOH | 1745 | 0.9 | N-doped | 4.30 | 44 |
Chitosan | KOH | 970 | 0.46 | N-doped | 5.00 | 77 |
Granular bamboo | KOH | 2641 | 1.25 | 4.5 | 55 | |
Cellulose | NaOH | 615 | 0.64 | N-doped | 4.99 | 45 |
Chitosan | K2CO3 | 1381 | 0.57 | N-doped | 3.86 | 46 |
Corncobs | NH3 | 1154 | 0.57 | N-doped | 2.81 | 78 |
Palm shell | NH3 | 889 | 0.47 | N-doped | 1.70 | 79 |
Banana peel | CO2 | 1426 | 0.83 | N-doped | 2.70 | 80 |
Coconut shell | CO2 | 1327 | 0.65 | 3.90 | 81 | |
Olive stones | CO2 | 1215 | 0.51 | 3.10 | 82 | |
Grass cuttings | CO2 | 841 | 0.38 | 1.45d | 83 | |
Cotton stalk | Steam | 439 | 0.19 | N-doped | 2.25 | 84 |
At 25 °C and 0.15 bar.
At 2 °C, N2–CO2 = 85 : 15.
At 0 °C and 1 bar.
At 0 °C and 0.1 bar.
Polysaccharides could be converted into carbonaceous materials via a hydrothermal treatment approach using water as the carbonization medium under a self-generated pressure. In 2011, Sevilla and Fuertes reported a series of porous carbons prepared by the KOH activation of hydrothermally carbonized polysaccharides (starch and cellulose) and biomass (sawdust).48 The prepared porous carbons were used as CO2 adsorbents. Although the porous carbons prepared at a ratio of KOH–precursor = 2 have a considerably low-level pore development than those obtained at a ratio of KOH–precursor = 4, they exhibited significantly better CO2 capture capacities. This is mainly due to the presence of a larger number of narrow micropores (<1 nm) in the mildly activated carbons. The sawdust-based carbon activated at 600 °C using KOH–precursor = 2 gave the highest CO2 uptake of 4.8 mmol g−1 (21.2 wt%) at 25 °C and 1 atm, which is still among the highest ever reported for activated carbons now.
Microporous carbons were prepared from wheat flour (mainly starch) via a combined process of pre-carbonization and post-KOH activation.59 The pore texture of the wheat-flour-based carbons was significantly improved by the post-chemical activation of KOH and varied with the KOH–carbon ratio. By increasing the KOH–carbon ratio up to 3, narrow micropores smaller than 0.8 nm were developed primarily, and the corresponding activated carbon possessed a moderate surface area but the highest volume of pores smaller than 0.8 nm, thus delivering the highest CO2 adsorption capacities of 5.70 and 3.48 mmol g−1 at 0 and 25 °C, respectively. The experimental results confirmed that CO2 adsorption uptake at ambient conditions was significantly dependent on the volume of narrow micropores with a pore size of less than 0.8 nm rather than the total volume or specific surface area.
Biomass waste is not only abundant, and sustainably renewable, but is also more cost-effective as carbon precursors. A wide range of biomass waste has been used to prepare porous carbon materials and showed excellent CO2 sorption capacities. Shell is one of them. As early as 2004, Tan and Ani reported carbon molecular sieves by direct carbonization of palm shell waste at 600–1000 °C.60 When the material was pyrolyzed at 600 °C, the micropore surface area was 753 m2 g−1; it showed CO2 adsorption near 2.3 mmol g−1 at 25 °C. When the pyrolysis temperature increased to 1000 °C, the micropore surface area and the pore volume decreased; however, the selectivity toward CO2 increased.60 After further KOH activation at 600 °C, the CO2 uptake of palm-shell-based porous carbon significantly increased from 2.7 mmol g−1 to 4.4 mmol g−1 at 25 °C and 1 bar, due to the much more developed microporosity of the activated carbons compared to the non-activated ones.61 Hu and his co-workers prepared two series of nitrogen-doped porous carbons from coconut shells.50,62 One type was prepared by urea modification and KOH activation, and the other was prepared by a successive process of pre-oxidization by H2O2, ammoxidation, and KOH activation. These carbons were found to exhibit very high CO2 uptakes at 1 bar, almost 5.0 mmol g−1 and 4.47 mmol g−1 at 25 °C, 1 bar, respectively. The pre-oxidization by H2O2 increased oxygen-containing surface groups, which were expected to increase the amount of nitrogen incorporated into the carbon product in the subsequent ammoxidation process. The resulting carbons possessed a much higher nitrogen content and a narrower microporosity than the control sample without H2O2 pre-treatment. These results made coconut shell an attractive precursor for carbon adsorbents.
Shahkarami et al. performed a comprehensive study on the effects of the nature of precursors and carbonization conditions on the physical and chemical properties of activated carbon and the influence of these factors on CO2 adsorption capacity, selectivity, and stability of the produced activated carbon in a fixed-bed reactor and mixed feed stream of N2/CO2/O2.63 In that work, three types of abundant feedstocks were used: agricultural waste (wheat straw and flax straw), forest residue (sawdust and willow ring), and animal manure (poultry litter). The selected precursors were carbonized via both fast and slow pyrolysis processes and were converted to porous carbons after KOH activation. Slow-pyrolysis-based activated carbon had a lower surface area and total pore volume but higher CO2 adsorption capacity in the presence of N2. Sawdust-based activated carbon synthesized from the slow pyrolysis possessed the largest ultramicropore volume of 0.36 cm3 g−1, and the highest CO2 adsorption capacity (78.1 mg g−1) in N2 but low selectivity (2.8) over O2 because of the oxygen functional groups on the surface. The ultramicropores and surface chemistry of adsorbents were found to be much more important than particle size, total pore volume, and internal surface area of the adsorbents.
Besides terrestrial raw biomass, marine raw biomass, such as microalgae64 and Enteromorpha prolifera,53 were also converted into carbons. For instance, a sugar-rich microalgae (Chlorococcum sp.) was used to prepare N-doped activated carbons through hydrothermal carbonization and a subsequent activation of KOH or NH3.64 Although the NH3-activated carbon possessed a much higher nitrogen content than the KOH-activated ones, the latter exhibited higher CO2 sorption capacities than the former, indicating the main contribution of narrow microporosity on the CO2 capacities. Ma et al. have successfully prepared highly porous N-doped carbon monoliths by using binary H3PO4–HNO3 mixed acid as a co-activating agent and sodium alginate, a marine biopolymer, as a carbon precursor.65 The SA-2N–P carbon prepared at a volume ratio of HNO3–H3PO4 = 2 showed high CO2 adsorption capacities of 8.99 mmol g−1 at 0 °C and 4.57 mmol g−1 at 25 °C, along with a high CO2 capacity of 1.51 mmol g−1 at 25 °C and 0.15 bar.
Microorganisms are another promising kind of carbon precursor. Wang et al. reported a series of porous carbons with narrow microporosities by KOH activation of pre-carbonized fungi (Agaricus).57 A moderate CO2 uptake of 5.5 mmol g−1 and a high CO2–N2 selectivity of 27.3 at 0 °C, 1 bar were obtained. Similarly, Shen et al. found that yeast-based carbon was a promising adsorbent.58 Hierarchical microporous carbon with a specific surface area of 1348 m2 g−1 and a pore volume of 0.67 cm3 g−1 was prepared by KOH activation of yeast. This type of carbon material showed a high CO2 uptake of 4.77 mmol g−1 and a fast adsorption rate with an equilibrium time less than 10 min at 25 °C.
1.2.1.2 Other Chemicals as Activating Agents
Other activating agents, such as ZnCl2,90,91 NaOH,85 K2CO3,38,86 H3PO4,92 NaNH2,94 etc. have also been used to prepare porous carbon for CO2 capture. ZnCl2 is the second most commonly used activating agent in chemical activation.96–98 However, there are only a few reports on ZnCl2 activated carbons for CO2 capture, as shown in Table 1.3, which may be due to its relatively undeveloped porosity as a result of the low activating power of ZnCl2.90,91 For example, ZnCl2-activated polypyrrole gave a specific surface area of 1283 m2 g−1 and pore volume of 0.46 cm3 g−1, much lower than those of KOH-activated polypyrrole (1700 m2 g−1 and 0.88 cm3 g−1).20,91 The as-prepared N-doped carbon exhibited a moderate CO2 uptake of 3.80 mmol g−1 at 25 °C and 1 bar when the activation temperature was 600 °C.91 Although the ZnCl2 activated carbons showed lower CO2 uptakes compared to the KOH activated carbons, the ZnCl2 activation usually gave a higher carbon yield and higher carbon density thus could exhibit some advantage in volumetric CO2 capacities.
Activated carbons prepared by other chemical activations
Precursor . | Activating agents . | SBET (m2 g−1) . | Pore volume (cm3 g−1) . | CO2 uptakea (mmol g−1) . | CO2–N2 selectivity . | Reference . |
---|---|---|---|---|---|---|
Polypyrrole | NaOH | 1005 | 0.53 | 3.73 | — | 85 |
Polyaniline | K2CO3 | 3670 | 1.70 | 7.60d | — | 86 |
Resorcinol resin | Potassium oxalate | 2130 | 1.10 | 4.7 | — | 87 |
m-Aminophenol–formaldehyde resin | K+ | 664 | — | 3.90 | 50c | 88 |
Carboxylic phenolic resins | Cs+ | 1312 | 0.67 | 5.20 | 18c | 89 |
Polyfuran | ZnCl2 | 1123 | 0.64 | 3.46 | 27.8b | 90 |
Polypyrrole | ZnCl2 | 1283 | 0.46 | 3.80 | — | 91 |
Eucalyptus wood | H3PO4 | 2079 | 1.29 | 3.22 | — | 92 |
Pine cone | H3PO4 | 1470 | — | 1.91 | — | 93 |
Mesoporous carbon | NaNH2 | 1917 | 1 | 3.66 | 16.2c | 94 |
RF polymer | NH3 | 1458 | 0.65 | 4.54 | — | 95 |
Precursor . | Activating agents . | SBET (m2 g−1) . | Pore volume (cm3 g−1) . | CO2 uptakea (mmol g−1) . | CO2–N2 selectivity . | Reference . |
---|---|---|---|---|---|---|
Polypyrrole | NaOH | 1005 | 0.53 | 3.73 | — | 85 |
Polyaniline | K2CO3 | 3670 | 1.70 | 7.60d | — | 86 |
Resorcinol resin | Potassium oxalate | 2130 | 1.10 | 4.7 | — | 87 |
m-Aminophenol–formaldehyde resin | K+ | 664 | — | 3.90 | 50c | 88 |
Carboxylic phenolic resins | Cs+ | 1312 | 0.67 | 5.20 | 18c | 89 |
Polyfuran | ZnCl2 | 1123 | 0.64 | 3.46 | 27.8b | 90 |
Polypyrrole | ZnCl2 | 1283 | 0.46 | 3.80 | — | 91 |
Eucalyptus wood | H3PO4 | 2079 | 1.29 | 3.22 | — | 92 |
Pine cone | H3PO4 | 1470 | — | 1.91 | — | 93 |
Mesoporous carbon | NaNH2 | 1917 | 1 | 3.66 | 16.2c | 94 |
RF polymer | NH3 | 1458 | 0.65 | 4.54 | — | 95 |
CO2 capacities at 25 °C and 1 bar.
Selectivity based on the ideal adsorbed solution theory (IAST).
The selectivity of Henry's law.
CO2 capacity at 0 °C and 1 bar.
Using K2CO3 as an activating agent, Fan et al. activated chitosan into N-doped microporous carbon.46 By changing the weight ratio of K2CO3–chitosan and the activation temperature, the porosity and nitrogen content of the prepared carbons could be tuned in the range of 1180–2567 m2 g−1 and 1.29–6.02 wt%, respectively. The sample prepared at 635 °C with a K2CO3-chitosan ratio of 2 showed a CO2 uptake of 3.86 mmol g−1 at 25 °C, 1 atm, five consecutive recyclabilities, and a good CO2–N2 selectivity of ca. 21. Silvestre-Albero prepared a series of activated carbons by pre-carbonization of PANI at different temperatures and post-activation of KOH or K2CO3.86 They studied the activating effects of KOH and K2CO3. It was found that carbonization temperature significantly influenced the porosity of the prepared carbons when using KOH as an activating agent, while K2CO3 mainly produced microporosity, independent of the carbonization temperature. The highest CO2 uptake of these carbons was 7.60 mmol g−1 at 1 bar and 0 °C.
In the last decade, ammonia (NH3) treatment has been used to introduce nitrogen surface groups into the carbon framework. Several authors have reported ammonia-treated porous carbons for CO2 capture.79,99 Pevida observed that nitrogen groups were successfully introduced into carbon frameworks after a NH3 treatment, and the nitrogen content was proportional to the oxygen content of the pristine porous carbons.99 Due to the incorporated basic nitrogen groups, the ammoxidized carbons showed enhanced CO2 uptakes at high adsorption temperatures. However, the low reaction efficiency between the NH3 and the carbon resulted in a relatively low N-doping level. Recently, Geng et al. reported a NH3-assisted activation process in which NH3 played the roles of activating agent and nitrogen source at the same time.78 When carbonizing a corncob in a NH3 atmosphere, the nitrogen could be easily incorporated into a carbon framework. The N content increased as the activation temperature increased, reaching a high level of ∼12 wt% at 800 °C, along with an increase in the specific surface area and pore volume. Especially, the nitrogen was mainly incorporated in the form of phenyl amine and pyridinic N groups, which were very efficient for CO2 capture. These carbons showed a moderate CO2 uptake of 2.81 mmol g−1, but superior IAST CO2–N2 selectivity up to 82. Similarly, Hu et al. prepared a N-doped hierarchical porous carbon by carbonization of a cellulose aerogel under a NH3 atmosphere.45 This N-doped carbon aerogel exhibited a N-doping of 4.62 wt%, and a high CO2 adsorption capacity of 4.99 mmol g−1 at 25 °C and 1 atm.
Most of the other activating agents usually show a lower activating ability than KOH. For example, KOH-activated polypyrrole carbons show a much higher specific surface area than the NaOH-activated ones under the same activation conditions (e.g., 2940 m2 g−1 vs. 1453 m2 g−1, activation conditions: the weight ratio activating reagent–polypyrrole = 2, and activation temperature = 700 °C).20,85 Interestingly, NaNH2, a strong base commonly used in organic synthesis, was reported to exhibit a more powerful activation ability compared to NaOH and KOH. At an activating reagent–carbon weight ratio of 2 and an activation temperature of 550 °C, NaNH2 could activate a much higher microporosity, especially more small micropores, than KOH and NaOH, resulting in a higher CO2 uptake of 3.66 mmol g−1 at 25 °C and 1 bar, verifying the superiority of NaNH2 in the activation under relatively moderate conditions.94
1.2.2 Physical Activation
The so-called physical activation is the partial gasification of the carbon framework with CO2, steam, and air, or a combination of these, at high temperatures (from 700 to 1200 °C) as shown in eqn (1.9) to (1.12). In view of porosity development, the most important variables in the gasification process are the activating agents, the final burn-off ratio, and the inorganic impurities. Rodríguez-Reinoso and Molina-Sabio comprehensively studied the evolution in the porosity of several series of activated carbons prepared by physical activation of lignocellulosic materials (uncatalyzed and iron-catalyzed) in CO2 or in a water–nitrogen mixture.100 They found that at the initial stage of CO2 activation, the micropore and macropore volume increase coincided with the proceeding burn-off. Further burn-off with CO2 opened and enlarged the micropores of the char with even a shift to meso- and macropores, the ablation of the exterior of the particle being very important at high burn-offs over 50%.
The final activated carbon had a well-developed micro- and macroporosity, with a relatively small portion of mesoporosity. The Fe-catalyzed CO2 gasification was initially very fast even at a lower gasification temperature and declined at an increasing burn-off due to the deactivation of the Fe catalyst. Steam activation produced a more selective attack of the carbon framework and a more uniform widening of porosity. It developed a large number of micropores, but fewer macropores, compared to CO2 activation.
1.2.2.1 CO2 as an Activating Agent
Mietek Jaroniec et al. prepared a series of porous carbon spheres by a combination of pre-carbonization and subsequent CO2 activation of phenolic resin spheres obtained by a modified Stöber method.101 The activated carbon spheres had diameters from 200 to 420 nm, a high surface area (from 730 to 2930 m2 g−1), abundant narrow micropores (<1 nm, pore volume of 0.28 to 1.12 cm3 g−1), thus exhibited very high CO2 adsorption capacities: 8.05 and 4.55 mmol g−1 at 1 bar and two temperatures, 0 and 25 °C, respectively.101 They further reported a series of N-doped activated carbon spheres.102 These materials were prepared by one-pot hydrothermal synthesis in the presence of resorcinol-formaldehyde resin as a carbon precursor and ethylenediamine (EDA) as both a base catalyst and nitrogen source, followed by carbonization in a N2 atmosphere and CO2 activation. The N-doping level and the particle size can be tuned by varying the EDA amount in the range of 2.80–7.20% and 50–1200 nm, respectively. These activated carbon spheres exhibited a high surface area and volume of micropores smaller than 1 nm, reaching up to 1224 m2 g−1 and 0.36 cm3 g−1, and could capture 6.2 and 4.1 mmol g−1 of CO2 at 0 °C and 25 °C. Obviously, the CO2 uptake of the N-doped CO2-activated carbon sphere was lower than those of non-doped or KOH-activated ones101 due its relative lower surface area and ultramicropores, indicating the primary importance of small micropores.
Two biomass materials – olive stones and almond shells – have been directly activated by CO2 to prepare porous carbon adsorbents for CO2 capture.82 The effect of holding time on overall yield, pore texture, and CO2 uptake, was studied. As the holding time was increased, the overall yield decreased and more porosity was developed, thus resulting in an increase of CO2 uptake. The olive-stone-based carbons showed a higher specific surface area and pore volume than the almond-shell-based ones, indicating that the olive stone was more easily activated by CO2. These carbons showed a CO2 uptake up to 3.10 mmol g−1 at 120 kPa and 25 °C, and 0.6–1.1 mmol g−1 at 15 kPa and 25–50 °C.
1.2.2.2 Steam as an Activating Agent
Steam was also used to activate both synthetic polymers and biomass into highly porous carbons. N-doped activated carbons were prepared by steam activation of melamine-modified phenol-formaldehyde resins.103 As the activation temperature increased, the specific surface area significantly increased from 382 to 1439 m2 g−1. The carbon activated at 850 °C could adsorb 6.71 mmol g−1 CO2 at 0 °C and 1 atm, and released 93.6% of the CO2 adsorbed when the temperature increased to 50 °C. Young-Jung Heo and Soo-Jin Park prepared cellulose-fiber-based ultramicroporous carbons by steam activation.104 The steam activation was found to have a strong influence on the development of new pores and the expansion of pore sizes and to be effective in developing optimal micropores for CO2 adsorption. The porous carbon prepared by pre-carbonization at 800 °C and further activation with steam exhibited a high CO2 adsorption capacity of 3.78 mmol g−1 at 25 °C and 1 bar, as well as an impressive Henry's law selectivity of 47.1 for CO2 over N2.
In short, researchers have synthesized a large number of porous carbon materials from a wide range of raw materials and activating agents for CO2 capture. The porosity and chemical surface properties varied with factors including raw materials, the pre-treatment process, activating agents, and activation temperature, etc. Particularly, many biomass materials, classified as polysaccharides, raw biomass and microorganisms here, are employed to successfully fabricate porous carbon adsorbents. Biomass-based carbon materials have shown high performances in CO2 capture and will play a more important role in post-combustion CO2 capture in the future. However, considering the huge emissions of carbon dioxide, there will be an enormous demand for solid CO2 sorbents, and biomass materials that are easy to mass-produce and are widely-collected should be the focus of studies. KOH is still the most common activating agent, but it is highly corrosive. Additionally, the KOH dosage in the current literature is too high (e.g., 200 wt% of the carbon precursor). Therefore, the development of new activating methods is also needed.
1.2.3 Metal Ion Activation
It has been widely accepted that the small micropores in carbon frameworks, especially ultramicropores smaller than 0.8 nm, are the most efficient for CO2 capture.105–107 Chemical or physical activation methods are still widely used in the preparation of highly microporous carbon materials. In order to get a high capture capacity of CO2, severe activation conditions, like excess activating agents and high burn-offs, are usually employed to increase the ultramicroporosity. In this way, an inhomogeneous activation is inevitable, resulting in a wide pore size distribution (PSD), low carbon yield and enhanced cost. Most activated carbon materials show a wide pore size distribution (PSD) and contain substantial supermicropores (larger than 1.0 nm), resulting in low CO2 uptake, especially at a low CO2 pressure (e.g., CO2 capture from flue gas). For instance, Zhao et al. have reported that porous carbon prepared by the direct carbonization of poly(vinylidene chloride) gave a slightly higher CO2 uptake than the further KOH-activated ones at 25 °C and 1 bar, because KOH activation generated some large micropores or small mesopores in the carbon framework, and then resulted in the reduction of interactions between CO2 molecules and the pore walls.29
Ultramicroporosity is desired for a high CO2 capture performance. Recently, we reported a new activating strategy for strictly ultramicroporous carbons by using phenolic resins as platforms, that could be called “single-ion activation”.49,50,108 In this process, mono-dispersed alkali ions are introduced via an acid–base reaction between an alkali hydroxide and the acidic groups (–OH or –COOH) of phenolic resins (Figure 1.2). After a direct carbonization of the prepared phenolic salts of resin, strictly ultramicroporous carbons are obtained. The mono-dispersed ions in the form of –O−M+ or –COO−M+ serve as activating agents, and produce a homogeneous activation effect, resulting in highly developed and narrowly-distributed ultramicropores.
This strategy was first used to prepare N-doped microporous carbons (NMCs).88 As shown in Figure 1.2, potassium phenolate of meta-aminophenol-formaldehyde resin was firstly prepared by the aqueous poly-condensation of meta-aminophenol and formaldehyde using equimolar KOH as both a catalyst precursor and a base. During the polycondensation reaction, the K+ ion can exchange with H+ in the hydroxyl groups, leading to the single dispersion of K+ ions in the bulk of the phenolic resin. After a simple carbonization process, N-doped ultramicroporous carbons were obtained.
From Figure 1.3, it could be seen that all the NMC samples exhibited a standard sorption isotherm of type I with a very narrow knee at a very low relative pressure (p/p0 < 0.02) and an absence of apparent adsorption increment in the relative pressures over 0.02, indicating that the porous carbons had a narrow microporosity and that no mesoporosity was present. The specific surface area of NMCs gradually increased from 272 m2 g−1 to 664 m2 g−1 as the activation temperature increased from 600 °C to 900 °C. Dubinin–Radushkevich (D–R) plots of CO2 sorption on NMCs exhibited well-defined linear shapes, showing that the porosity of NMCs was strictly made up of uniform ultramicropres.107,109 The average sizes of these ultramicropores calculated according to the slope of the D–R plots gradually increased from 0.50 to 0.58 nm as the carbonization temperature increased from 600 °C to 900 °C.
(a) Nitrogen sorption isotherms, (b) QSDFT pore size distributions derived from N2 sorption, (c) NLDFT pore size distributions derived from CO2 sorption, (d) D–R plots for CO2 sorption by NMC-600. Reproduced from ref. 88 with permission from the Royal Society of Chemistry.
(a) Nitrogen sorption isotherms, (b) QSDFT pore size distributions derived from N2 sorption, (c) NLDFT pore size distributions derived from CO2 sorption, (d) D–R plots for CO2 sorption by NMC-600. Reproduced from ref. 88 with permission from the Royal Society of Chemistry.
The carbon prepared at 600 °C (NMC-600) exhibited a high CO2 uptake of 3.90 mmol g−1 at 25 °C and 1 bar, although this carbon possessed a very low specific surface area of 272 m2 g−1. Besides, this carbon gave an outstanding CO2 uptake of 1.67 mmol g−1 mmol g−1 at 25 °C and 0.15 bar, and an impressive CO2-over-N2 selectivity of 50 based on Henry's law. At 75 °C and 1 bar, the carbon prepared at 800 °C (NMC-800) still showed a very high CO2 uptake of 2.0 mmol g−1. The excellent performance of NMCs for CO2 capture indicated that these carbons preferably interact with CO2 molecules based on the combined contributions of the extremely small micropores ca. 0.5 nm and the polar surface caused by the N, O-doping.
The strategy of single-ion activation could be applied to finely tune the sizes of ultramicropores by using different alkali metal ions (Li+, Na+, K+, Rb+, Cs+) as activating agents (Figure 1.4).89 All the samples presented a standard type I N2 adsorption isotherm, showing a narrow micropore size distribution. The pore volume and the apparent surface areas gradually increased from 0.07 to 0.70 cm3 g−1 and 111 to 1312 m2 g−1, respectively, while varying the activating ion from Li+ to Cs+, thereby revealing that the activation power significantly increased from Li+ to Cs+. All the samples, except Li+-activated carbon (LiAC), exhibited well-defined linear D–R plots (correlation coefficient R2 > 0.995), indicating the existence of strictly uniform ultramicropores (Figure 1.5). Interestingly, a systematic widening of micropores took place; the pore size gradually increased from 0.60 nm up to 0.76 nm as the activation ions varied from Li+ to Cs+. In other words, the ultramicropore size was finely tuned at sub-angstrom level by simply varying the activating ions.
Schematic diagram of the synthesis of MAC. Reproduced from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Schematic diagram of the synthesis of MAC. Reproduced from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
(a) Nitrogen adsorption isotherms; analysis of the narrow microporosity by Dubinin–Radushkevich equations for: (b) LiAC, (c) NaAC, (d) KAC, (e) RbAC and (f) CsAC. Adapted from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
(a) Nitrogen adsorption isotherms; analysis of the narrow microporosity by Dubinin–Radushkevich equations for: (b) LiAC, (c) NaAC, (d) KAC, (e) RbAC and (f) CsAC. Adapted from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
When applying for CO2 capture, the Cs+-activated carbon (CsAC) showed a superior CO2 adsorption capacity of 5.20 mmol g−1 at 25 °C and 1.0 bar, high CO2-to-N2 selectivity and full regenerability for four consecutive cycles (Figure 1.6). Furthermore, the resulting carbons present ultrahigh capacitances of up to 223 F g−1 or 205 F cm−3 in an ionic liquid electrolyte. The outstanding performance of these carbons is due to their uniform pore size, which is finely tuned to adapt to the dimensions of the CO2 molecule and the ions of ionic liquids. The CO2 capacities and specific capacitances reported here were higher than for most traditional activated carbons. More importantly, the dosage of activation agents used herein was only 30 wt% of the carbon precursors as counted in KOH, much lower than that used in traditional chemical activation,89 confirming the high efficiency of metal ion activation.
(a) CO2 adsorption isotherms at 0 °C and 1.0 bar; (b) CO2 adsorption isotherms at 25 °C and 1.0 bar. Adapted from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
(a) CO2 adsorption isotherms at 0 °C and 1.0 bar; (b) CO2 adsorption isotherms at 25 °C and 1.0 bar. Adapted from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
The mechanism of metal ion activation was systematically studied by thermogravimetry-mass spectrometry (TG-MS) analysis and in-situ X-ray diffraction (XRD) (Figure 1.7). The pyrolysis process of the potassium salt of carboxylic phenolic resin (PR-COOK) was much more complicated than that of the phenolic resin carboxylic acid (PR-COOH) due to the activation of metal ions. Three gases, including H2O (m/z 18), CO (m/z 28), and CO2 (m/z 44), were detected by mass spectrometry. In-situ XRD patterns of PR-COOK indicated that the PR-COOK began to decompose into K2CO3 (PDF card 71-1466) at around 200 °C, and then transformed into K2O at 600 °C, along with two steps of weight loss and the release of CO2 and CO. When the carbonization temperature increased to 800 °C, new peaks at 10.4° and 16.5° appeared, which is attributed to K (100) (PDF card 01-0500) and KC8 (100) (PDF card 04-0221), and large amounts of CO were detected (Figure 1.7). These facts are strong evidence that metallic potassium was produced by the reduction of K2O and was further intercalated into the carbon layers to form graphite intercalation-like compounds.
(a) TG curves, MS responses of evolved gases in TG-MS analysis, (b) H2O, (c) CO2, and (d) CO, (e) in-situ XRD patterns (CuKa) during the carbonization of PR-COOK, and (f) possible activation mechanism. Reprinted from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
(a) TG curves, MS responses of evolved gases in TG-MS analysis, (b) H2O, (c) CO2, and (d) CO, (e) in-situ XRD patterns (CuKa) during the carbonization of PR-COOK, and (f) possible activation mechanism. Reprinted from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
The mechanism of single-ion activation was suggested as: first, alkali salts of phenolic resins decomposed into alkali carbonate (M2CO3), H2O and CO2 below 400 °C. Second, alkali oxides (M2O) were produced by the thermal decomposition of M2CO3 or the redox reaction between M2CO3 and C. Third, framework carbon atoms were etched by M2O via the vigorous redox reaction of M2O + C → 2M + CO, then the produced metallic alkali intercalated into the lattices of the carbon matrix. The interlayer spacing will be primarily determined by the size of the metal intercalate, and this spacing (i.e., pore size) will be systematically widened with increasing metal ion size from Li+ to Cs+. In a word, the activation is similar to traditional KOH activation, but much more efficient.
This strategy was further applied for the synthesis of microporous carbon spheres for the following reasons: (1) the developed ultramicroporosity offered a large space for CO2 adsorption; (2) the small size of the carbon spheres ensured a short diffusion distance of the CO2 molecule; (3) the gaps between carbon spheres allowed a rapid flow of flue gas.110 In that work, potassium salts of resorcinol-formaldehyde (RF) resin spheres were prepared by the reaction of KOH and pristine RF resin spheres, and followed by carbonization into microporous carbon materials (denoted as CS-x, where x stands for the weight ratio of KOH–RF) (Figure 1.8). When the prepared resins were mixed with KOH, the –OH groups of the RF resins immediately reacted with KOH via a fast acid–base reaction and were changed into mono-dispersed K+ in the form of –OK groups. Based on a simplified molecular structure of RF resins, the KOH–RF resins' weight ratios of 0.50, 0.75, 1, 1.5 and 2 are roughly equivalent to the KOH–OH molar ratios of 0.50, 0.75, 1, 1.5 and 2. In the cases of CS-0.5, CS-0.75 and CS-0.75, the activating agents exclusively existed as –OK groups, while in the cases of CS-1.5 and CS-2, the KOH was in excess (Figure 1.8).
Illustration for preparation of the porous carbon spheres. Reprinted from X. Wang, J. Zhou, W. Xing, B. Liu, J. Zhang, H. Lin, H. Cui and S. Zhuo, Resorcinol–formaldehyde resin-based porous carbon spheres with high CO2 capture capacities, J. Energy Chem., 26, 1007–1013, Copyright 2017, with permission from Elsevier.
Illustration for preparation of the porous carbon spheres. Reprinted from X. Wang, J. Zhou, W. Xing, B. Liu, J. Zhang, H. Lin, H. Cui and S. Zhuo, Resorcinol–formaldehyde resin-based porous carbon spheres with high CO2 capture capacities, J. Energy Chem., 26, 1007–1013, Copyright 2017, with permission from Elsevier.
After activation, the majority of the carbon particles retained, roughly, their spherical shape even when the weight ratio of KOH to RF resins was up to 2 (Figure 1.9). The detailed microscopy morphology and the porosity of the prepared carbon materials were closely related to KOH dosage (Figures 1.9 and 1.10). The CS-0.5, CS-0.75 and CS-1 samples showed regular spherical shapes, a smooth carbon surface, and an exclusively narrow microporosity (<0.8 nm), indicating that the K+ activation almost occurred in the interiors of the carbon mono-spheres. In comparison, some defects and roughness were observed on the surface of the CS-1.5 and CS-2 carbon particles, and a widening of PSD took place, indicating a non-homogeneous activation. As the KOH dosage increases, the surface areas and volumes of the micropores first increase and then decrease, and CS-1 possesses the largest specific surface area, micropore surface area and volume, up to 1235 m2 g−1, 1084 m2 g−1 and 0.57 cm3 g−1, respectively.
Microscopic morphology of CS-x. SEM images: (a, b) RF resin spheres, (c, d) CS-0.5, (e) CS-0.75, (f) CS-1, (g) CS-1.5 and (h) CS-2. Reprinted from X. Wang, J. Zhou, W. Xing, B. Liu, J. Zhang, H. Lin, H. Cui and S. Zhuo, Resorcinol–formaldehyde resin-based porous carbon spheres with high CO2 capture capacities, J. Energy Chem., 26, 1007–1013, Copyright 2017, with permission from Elsevier.
Microscopic morphology of CS-x. SEM images: (a, b) RF resin spheres, (c, d) CS-0.5, (e) CS-0.75, (f) CS-1, (g) CS-1.5 and (h) CS-2. Reprinted from X. Wang, J. Zhou, W. Xing, B. Liu, J. Zhang, H. Lin, H. Cui and S. Zhuo, Resorcinol–formaldehyde resin-based porous carbon spheres with high CO2 capture capacities, J. Energy Chem., 26, 1007–1013, Copyright 2017, with permission from Elsevier.
N2 sorption measurements for CS-x materials (a, b) N2 sorption isotherms, (c, d) PSDs plots calculated by QSDFT model. Reprinted from X. Wang, J. Zhou, W. Xing, B. Liu, J. Zhang, H. Lin, H. Cui and S. Zhuo, Resorcinol–formaldehyde resin-based porous carbon spheres with high CO2 capture capacities, J. Energy Chem., 26, 1007–1013, Copyright 2017, with permission from Elsevier.
N2 sorption measurements for CS-x materials (a, b) N2 sorption isotherms, (c, d) PSDs plots calculated by QSDFT model. Reprinted from X. Wang, J. Zhou, W. Xing, B. Liu, J. Zhang, H. Lin, H. Cui and S. Zhuo, Resorcinol–formaldehyde resin-based porous carbon spheres with high CO2 capture capacities, J. Energy Chem., 26, 1007–1013, Copyright 2017, with permission from Elsevier.
The as-prepared carbon spheres exhibited excellent CO2 capture capacities ranging from 6.55–7.34 mmol g−1 at 0 °C and 4.27–4.83 mmol g−1 at 25 °C and 1 bar, and the highest values belonged to CS-1, which is attributed to its largest micropore volume with narrow micropore size. The value of 4.83 mmol g−1 is higher than those for lots of phenolic-resin-based porous carbon materials under the same conditions, including porous carbon spheres prepared by traditional chemical or physical activation, such as CO2-activated carbon spheres,101,107 N-rich microporous carbon spheres,102 and KOH-activated carbon spheres.27
Overall, single-ion activation has been used to develop and adjust ultramicroporosity. This strategy has exhibited many merits, such as a low demand for activating agents, a high activating efficiency, easy operation, and it reaches high CO2 sorption capacities; it is worthy of further study.
1.2.4 Templating Method
Historically, the templating method was first reported by Knox and co-workers in 1986, who demonstrated the synthesis of graphitic porous carbons for liquid chromatography separation by impregnation of spherical porous silica gel particles with phenolic resin and subsequent carbonization and removal of silica.111 Since then, this method has received extensive attention and various types of template carbons are synthesized. The resulting carbon synthesized by the templating method, called a templated carbon, possesses a relatively narrow PSD and controlled architecture. The templated carbonization method permits one to control the carbon structure in terms of various aspects, such as pore structure, specific surface area, microscopic morphology and graphitizability, which makes this method very attractive.112,113
Hard templates and soft templates are the main kinds of templates used as molds to form templated carbons. The hard template method includes the following steps: (a) synthesis of a suitable porous template; (b) introduction of a suitable carbon precursor into the template pores using the method of wet impregnation, chemical vapor deposition or a combination of both methods; (c) polymerization and carbonization of the carbon precursor; and (d) removal of the inorganic template.114,115 Following these steps, porous carbon with a specific pore structure is formed. Compared with the hard template, the soft template is a kind of surfactant, which has a strong interaction with the carbon source, and mesoporous carbons with different structures can be obtained through the soft template method. This method possesses good controllability and operability; as a result, it has very good application prospects. The mechanisms of the soft template method include a liquid crystal template mechanism, a synergistic assembly mechanism, a “rod micellar” mechanism and so on; these mechanisms have been widely recognized.116
Since the porous carbon replicates the morphology of the template, selecting a suitable template to synthesize the carbon with a specific porous structure is the most important step. Template carbons with different pore structures, such as ordered porous carbon, disordered porous carbon and hierarchical porous carbons, can be prepared by different templates, such as porous silica,117–120 zeolites,121 metal–organic frameworks (MOFs),122 nanoparticles (e.g., MgO123 and SiO2 124 ), and surfactant micelles.125–131 These templates will be described in detail in the following sections.
1.2.4.1 Porous Silica as a Hard Template
Mesoporous molecular sieve SBA-15 has a large surface area, uniform pore diameter distribution, adjustable pore size and wall thickness, and high hydrothermal stability. Therefore, SBA-15 is an excellent template to prepare ordered porous carbons.120,132 Wang et al. designed a new method of using a one-step “surfactant-assisted” nanocasting route to synthesize highly ordered mesoporous graphitic carbon. They selected ordered mesoporous silica SBA-15 maintaining its triblock copolymer surfactant P123 (EO20PO70EO20) as the hard template and natural organic soybean oil (SBO) as a carbon precursor. Finally, the ordered mesoporous graphitic carbon was synthesized and named as MGC-1 (Figure 1.11).133 The main advantage of this method is that the carbon precursor SBO can easily infiltrate the mesopore channels of the silica template with the help of P123 to reach a high filling level, resulting in an enhanced yield of graphitic carbon materials. Because of the improved structural ordering, the mesoporous carbon, after amine modification, could adsorb more CO2 (57.2 mg g−1) compared with amine-functionalized carbon prepared without the assistance of surfactant (31.3 mg g−1).133
Schematic illustration of the synthesis process of ordered mesoporous graphitic carbon by a “surfactant-assisted” nanocasting route. Reproduced from ref. 133, http://dx.doi.org/10.1038/srep26673, © The Authors. Published under the terms of the CC BY 4.0 licence, https://creativecommons.org/licenses/by/4.0/.
Schematic illustration of the synthesis process of ordered mesoporous graphitic carbon by a “surfactant-assisted” nanocasting route. Reproduced from ref. 133, http://dx.doi.org/10.1038/srep26673, © The Authors. Published under the terms of the CC BY 4.0 licence, https://creativecommons.org/licenses/by/4.0/.
In order to clarify the influence of porosity and surface chemistry on carbon dioxide capture, Sanchez–Sanchez et al. synthesized ordered mesoporous carbons with high surface areas and pore volumes by the hard template method, in which 3-aminobenzoic acid acted as both carbon precursor and nitrogen source, and SBA-15 acted as the hard template. The synthesis was accomplished in the presence of different H3PO4 concentrations. The surface chemistry and the porous texture of the carbons could be easily modulated by varying the H3PO4 concentration and carbonization temperature. Finally, they found that with H3PO4 concentrations higher than 50 wt%, the pyrolysis mechanism of the thermally polymerized precursor changed and both the structural order and mesopore arrangement declined. On the contrary, low H3PO4 concentrations led to mesoporous carbon materials with a very narrow PSD and favored nitrogen retention at the carbon surface. By comparing the amount of CO2 adsorption of different samples under different conditions, they also proved that the CO2 adsorption capacity depends on the narrow micropore volume and is almost unaffected by the surface chemistry at 0 °C and 1 bar; the largest CO2 uptake of 5.04 mmol g−1 was obtained for the carbon with the most developed ultramicroporous structure. On the other hand, the nitrogen functional groups exerted a beneficial influence on CO2 capture, while oxygen and phosphorus functionalities exerted a negative influence at 25 or 50 °C and 1 bar. Moreover, among the N-containing groups, the pyrrolic N groups provoked the largest improvements in CO2 adsorption.134
In addition to SBA-15, there are also other porous silica templates used to prepare ordered mesoporous carbons. For instance, Tiwari prepared oxygen enriched carbon adsorbents with epoxy resin as the precursor and MCM-41 as a template.135 The sample prepared at 700 °C shows the highest capacity of CO2 adsorption of 0.65 mmol g−1 at 30 °C at a CO2 concentration of 12.5%. Zhao et al. reported nitrogen-doped mesoporous carbons, which were prepared by a nanocasting route using tri-continuous mesoporous silica IBN-9 as a hard template and p-diaminobenzene as a nitrogen-containing precursor.136 One of the carbons possessed a very high nitrogen doping concentration (13 wt%) and exhibited an excellent performance for CO2 adsorption at about 4.50 mmol g−1 at 298 K and 1 bar.
1.2.4.2 Zeolite as a Hard Template
Zeolites are highly crystalline aluminosilicate materials that possess uniform sub-nanometer-sized pores. The pore channel apertures vary between 0.3 and 1.5 nm. As a result, zeolite-templated carbons (ZTCs) possess highly ordered microporosity and a large surface area, which is helpful in enhancing the capabilities of CO2 capture. With these advantages, ZTCs are especially attractive and many groups have prepared different types of ordered microporous carbons using various zeolites as templates to study the capabilities of CO2 capture.137,138
Xia et al. synthesized zeolite-templated, high-surface-area, microporous, N-doped carbons, which exhibit CO2 uptake capacities of up to 6.9 mmol g−1 at 273 K at 1 bar and 4.4 mmol g−1 at 298 K at 1 bar, along with CO2 adsorption energies of 36 kJ mol−1 to 20 kJ mol−1.139 Combined with their ease of preparation, regeneration stability, excellent recyclability and high selectivity for CO2, the N-doped zeolite-templated carbons are suitable for CO2 capture. They further successfully synthesized structurally well-ordered sulfur-doped microporous carbon materials for the first time.121 The sulfur-modified ordered microporous carbons were obtained by infiltration of the carbon precursors (2-thiophenemethanol) into the pores of zeolite EMC-2 via impregnation combined with a CVD method. They found that S-doped microporous carbons with well-resolved XRD patterns arising from a high level of zeolite-type structural ordering can be nanocasted from zeolite EMC-2. The carbon materials possessed a high surface area with varied sulfur content depending on the preparation conditions. They found that both the presence of S functional groups and the textural properties play important roles in the capacity of CO2 adsorption. These carbons also showed a very high CO2 adsorption energy up to 59 kJ mol−1.121
1.2.4.3 Porous Organic Frameworks as Self-templates
Porous organic frameworks (POFs), such as metal–organic frameworks (MOFs), zeolite imidazole frameworks (ZIFs) and covalent organic frameworks (COFs), have been intensively investigated for CO2 capture application recently.42,140,141 MOFs generally consist of two main components: organic ligands and metal ions. The structures of MOFs change with the types of inorganic metal ions and organic ligands and can be functionally modified by a variety of methods. ZIFs are a newly emerging class of porous crystals with extended three-dimensional structures constructed from tetrahedral metal ions (e.g., Zn, Co) bridged by imidazolate. Another kind of porous material that has emerged recently, COFs possess a long-range ordered network composed of covalent connections by the light elements (C, N, O, B). Recently, these porous organic frameworks have been used to prepare porous carbons and related nanostructured functional materials.142 Those carbons have also been quickly developed and used in CO2 capture.
Generally, there are two routes to fabricate porous carbons from these porous organic frameworks. One is using the porous organic frameworks as sacrificial templates with the incorporation of an additional carbon source. Xu and his co-workers used the classical MOF-5 as a sacrificial template and precursor with furfuryl alcohol as the additional carbon source to prepare porous carbons, which is the first report of MOF-derived porous carbons.143 The resultant porous carbon obtained at 1000 °C exhibited a high specific surface area up to 2872 m2 g−1. The as-synthesized porous carbon exhibited a high H2 uptake, reaching 2.6 wt% at 77 K and 1 bar, much higher than that of raw MOF-5 under the same conditions. Moreover, this material gave a good electrochemical performance as an electrode material for supercapacitors with a capacitance of 258 F g−1 at a current density of 0.05 A g−1.143 They further investigated the effect of carbonization temperature on the porosity of the resultant porous carbons.144 The specific surface area for MOF-5 templated porous carbons obtained at carbonization temperatures from 530 to 1000 °C is in the range from 1140 to 3040 m2 g−1, and the dependence of surface area on carbonization temperature shows a “V” shape.
Wang prepared N-doped porous carbons using ZIF-70 with a high porosity and large pore size as the model system and polyethyleneimine (PEI) molecules as the nitrogen source. Many PEIs can be readily incorporated into the framework pores of ZIF-70. Moreover, the ZIF-70 framework has many N elements, which can be partly doped into a carbon network during the carbonization process. Finally, a series of N-rich porous carbon materials was successfully fabricated by carbonization of polyamine-incorporated MOFs, which had an excellent CO2 capture capacity of 4.86 mmol g−1 at 1 bar and 25 °C.145 Pachfule et al. reported a series of porous carbon materials by using isoreticular zeolitic imidazolate frameworks (IRZIFs).122 A series of carbon materials named as C-68, C-69 and C-70 was prepared using ZIF-68, ZIF-69 and ZIF-70 as templates and furfuryl alcohol as the additional carbon source under a carbonization temperature of 1000 °C. Among these carbons, the resultant C-70 synthesized using ZIF-70 as the template shows the most developed porosity with the highest specific surface area of 1510 m2 g−1 due to the more porous ZIF-70 template. The porous carbons C-70, C-68 and C-69 show CO2 uptake capacities (273 K and 1 atm) of 5.45, 4.98 and 4.54 mmol g−1 for C-70, C-68 and C-69, respectively. Also, the C-70 carbon shows the highest H2 uptake capacity (77 K and 1 atm) of 2.37 wt% among these carbons. Zhang et al. reported a new strategy of pre-introducing an extra carbon source (furfuryl alcohol) into a PAF (porous aromatic framework) followed by carbonization, which afforded a new microporous carbon material with a small pore size of 0.54 nm, thus facilitating a high CO2 uptake capacity of about 4.1 mmol g−1 at 295 K and 1 bar.146
The other route is direct carbonization in which the POFs serve as both self-templates and carbon precursors at the same time without any additives. Srinivas et al. reported a new type of hierarchical porous carbon from the controlled carbonization of Zn-MOF-5, Zn-MOF-74, and Al-MIL-53 (Figure 1.12).147 The authors found that an increase in the size of the MOF-5 precursor could effectively enhance the specific surface area of the resultant porous carbons. The simultaneously enhanced surface area and pore volume in HPC5b2 were attributed to the large, millimeter-sized crystallites of the MOF precursor, that delivered more void space due to the extended pores within the crystals after carbonization, compared with the smaller crystals in HPC5b1. HPC5b2-1100, obtained by carbonization of millimeter-sized MOF-5 at 1100 °C showed a very high specific surface area and total pore volume, up to 2734 m2 g−1 and 5.23 cm3 g−1, respectively. Through simultaneous thermogravimetric analysis and mass spectroscopy (H2, H2O, CO, O2, CO2, and carbonaceous gases), the carbonization process and mechanism for all of the MOFs were investigated. Taking Zn-MOF-74 as an example, as shown in Figure 1.12, the first mass-loss, below 150 °C, was due to the evaporation of adsorbed/terminal water molecules within the pores and the second mass-loss between 400 °C and 600 °C was attributed to framework decomposition leading to a major release of carbon containing gaseous products (mostly CO2, CO, C6H6 and a small amount of H2 and CxHy hydrocarbon mixtures) and formation of ZnO occurs. The third mass-loss starting at 900 °C was due to a further release of CO2 and CO with Zn through the reduction of ZnO by carbon via ZnO + C → Zn(g) + CO. In most cases, the CO2 uptakes in these MOFs-derived carbons were higher than in their MOF precursor. A high CO2 uptake, over 27 mmol g−1 (119 wt%) at 30 bar and 27 °C, is obtained, which is one of the largest reported in the literature for porous carbons. However, the CO2 capacity of these carbons at ambient pressure was only about 3 mmol g−1 due to their relatively low ultramicroporosity.
Schematic representation of MOF-5, MOF-74 and MIL-53 framework structures. Bottom: mechanism involved in the carbonization process of Zn-MOF-74. The plot represents the MOF mass change vs. carbonization temperature. Reproduced from ref. 147 with permission from the Royal Society of Chemistry.
Schematic representation of MOF-5, MOF-74 and MIL-53 framework structures. Bottom: mechanism involved in the carbonization process of Zn-MOF-74. The plot represents the MOF mass change vs. carbonization temperature. Reproduced from ref. 147 with permission from the Royal Society of Chemistry.
Ding et al. reported a series of N-doped porous carbon monoliths by direct carbonization of IRMOF-3, an isoreticular MOF-3.148 The ligand (2-aminobenzene-1,4-dicarboxylic acid) served as the nitrogen source. Meanwhile, the ZnO nanoclusters formed during the carbonization process acted as the in-situ activator and self-template, which was the key to the transformation of the microporous structure to the meso-macroporous structure. With an increase of the carbonization temperature, the specific surface area, micropore surface area and volume calculated using the t-plot method simultaneously increased. Compared to the pristine IRMOF-3, IRMOF-3/800 carbonized at 800 °C exhibited a much higher CO2 capacity (3.99 vs. 2.32 mmol g−1 at 1 bar and 1.64 vs. 0.70 mmol g−1 at 0.15 bar) and CO2–N2 selectivity (22.7 vs. 7.4 at 0 °C and 38.8 vs. 8.6 at 25 °C).
Ma et al. synthesized a N-doped porous carbon using zeolitic imidazolate framework-8 (ZIF-8) both as a solid template and precursor. The resultant carbons were further modified with NH3. It was confirmed that absorption of CO2 was enhanced by ammonia treatment. This is attributed to the presence of increased N-containing functional groups and a specific surface area during modifications.149 Similar to that, Zhao reported the design and synthesis of a series of new B, N-containing cross-linked polymers with well-defined chemical structures and high thermal stabilities. Further, they were efficiently converted into B/N co-doped porous carbons in high yields under pyrolysis treatment. The prepared carbons showed excellent CO2 uptakes (3.25 mmol g−1 at 273 K and 1 atm) with respect to their relatively low surface areas, and high selectivity of CO2–CH4 with a ratio of more than 5 : 1 at 298 K.150
Unambiguously, MOFs, ZIFs and COFs have been emerging as very promising templates and/or precursors to prepare highly porous carbons that show great potential in gas storage and electrical applications. However, the CO2 capacities of MOFs (ZIF, COFs)-derived carbons are at a moderate level among the porous carbons. Further exploration is needed, such as how to introduce high microporosity and heteroatom-doping, the development of low-cost MOFs (ZIFs, COFs) and their large-scale synthesis methods, etc.
1.2.4.4 Carbide Lattices as Self-templates
Carbide-derived carbons (CDCs) are synthesized by selective extraction of a metal and metalloid from carbide precursors. High-temperature chlorination is the most widely used method for CDC synthesis (eqn (1.13), Figure 1.13).151,152 In general, the starting carbide precursor templates the initial pore size of the CDCs, and then any range of average pore sizes from less than 1 nm to greater than 10 nm can be achieved by changing the chlorination temperature.153 From this perspective, this preparation could be considered as a crystal lattice templating process. The most well-known application of CDCs is as a supercapacitor, which was led by Prof. Gogotsi.154,155 These carbons have also shown great promise in H2 storage,156–158 and CH4 storage,159 as well as CO2 capture.106,159–163
(a) Unit cell of Ti3SiC2. (b) Schematic of the lattice structure of Ti3SiC2 with and without Ti and Si atoms shown. (c) Proposed model for generation of cracks between the grains. Contraction of CDC along the [0 0 0 1] direction of the former Ti3SiC2 is suggested. Reprinted from G. N. Yushin, E. N. Hoffman, A. Nikitin, H. Ye, M. W. Barsoum and Y. Gogotsi, Synthesis of nanoporous carbide-derived carbon by chlorination of titanium silicon carbide, Carbon, 43, 2075–2082, Copyright 2005, with permission from Elsevier.
(a) Unit cell of Ti3SiC2. (b) Schematic of the lattice structure of Ti3SiC2 with and without Ti and Si atoms shown. (c) Proposed model for generation of cracks between the grains. Contraction of CDC along the [0 0 0 1] direction of the former Ti3SiC2 is suggested. Reprinted from G. N. Yushin, E. N. Hoffman, A. Nikitin, H. Ye, M. W. Barsoum and Y. Gogotsi, Synthesis of nanoporous carbide-derived carbon by chlorination of titanium silicon carbide, Carbon, 43, 2075–2082, Copyright 2005, with permission from Elsevier.
The TiC-CDC produced at a chlorination temperature of 800 °C showed a good H2 capacity of 2.55 wt%, which was further improved to 3.0 wt% after a post-treatment of H2 annealing.157 Gogotsi et al. prepared a large number of CDCs with various pore sizes by using different starting carbides (B4C, ZrC, TiC and SiC) and chlorination temperatures to explore the effects of pore size and volume on the H2 sorption and heat of adsorption.156,157 They found that H2 storage was dominated by small pores, and thus not directly connected to total specific surface area. A clear dependence on pore size was also suggested – the smaller pores increased both the heat of adsorption and the H2 uptake.156
As far as we know, the first study of CDCs for CO2 adsorption was reported by Presser and his co-workers in 2011.106 The Micro-TiC-CDC sample prepared under the combined conditions of chlorination at 700 °C and post-H2 annealing at 600 °C showed a mean pore size of 0.66 nm and exhibited a very high sorption capacity for CO2 of up to 7.1 mmol g−1 at 0 °C and 1 bar. The highest CO2 capacity at 0.1 bar and 0 °C is 1.62 mmol g−1 delivered by the sample prepared at 600 °C. More importantly, they comprehensively studied the effect of pore size on CO2 capacities at different sorption pressures. This will be detailed in Section 1.6.1. Bhatia and Nguyen studied the potential of carbon derived from SiC (SiC-CDC) for CO2 capture from flue gas.161 They found that N2 was very weakly adsorbed onto SiC-CDC, and H2O adsorption on SiC-CDC was highly kinetically restricted due to the strong hydrophobicity of this material inherited from the diamond-like structure (sp3 bonding) of the SiC precursor. This feature is very promising for the selective adsorption of CO2 from real flue gas since the co-adsorption of H2O can significantly degrade the performance of any adsorbents for CO2 capture.161
1.2.4.5 Triblock Copolymer as a Soft Template
Wei et al. synthesized nitrogen-doped ordered mesoporous carbon materials by using a self-assembly process with dicyandiamide as a nitrogen source, soluble resol as a carbon source and a triblock copolymer (F127) as a soft template. Figure 1.14 shows the synthesis process of this ordered N-doped mesoporous carbon. In this synthesis, resol molecules can bridge the template F127 and dicyandiamide via hydrogen bonding and electrostatic interactions. The obtained N-doped ordered mesoporous carbons possess tunable mesostructures and pore sizes (3.1–17.6 nm), high surface areas (494–586 m2 g−1), and high N contents (up to 13.1 wt%). Ascribed to the unique feature of a large surface area and high N content, the carbon materials show a good CO2 adsorption capacity of 2.8–3.2 mmol g−1 at 298 K and 1.0 bar.131
The formation process of ordered N-doped mesoporous carbon from a one-pot assembly method using dicyandiamide (DCDA) as a nitrogen source. Reproduced from ref. 131 with permission from John Wiley and Sons, Copyright 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
The formation process of ordered N-doped mesoporous carbon from a one-pot assembly method using dicyandiamide (DCDA) as a nitrogen source. Reproduced from ref. 131 with permission from John Wiley and Sons, Copyright 2012 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Zhang et al. reported a facile one-pot melting-assisted solvent-free method to develop ordered mesoporous carbons (Figure 1.15).127 This method only involved a simple thermal treatment procedure and avoided the time-consuming process of evaporation-induced self-assembly (EISA). Using this method, non-doped and N-doped mesoporous carbons can be synthesized. The non-doped carbon possessed a well-developed hierarchical pore texture with a high surface area of 748 m−2 g−1 and exhibited a CO2 capacity of 2.73 mmol g−1 at 75 °C and 1 bar. Due to the large amounts of nitrogen in the carbon matrix (11.67%), the N-doped mesoporous carbon exhibited excellent CO2 capture properties, with a CO2 capacity of 1.66 mmol g−1 as well as an excellent IAST selectivity of CO2-to-N2, up to 240.7 at 75 °C and 1 bar. Wu et al. reported the consecutive incorporation of carbon nitride and nitrogen-containing functionalities into FDU-type ordered mesoporous carbon materials by using a simple post-synthetic route with melamine as a nitrogen source and F127 as the soft template. The nitrogen-enriched mesoporous carbons deliver promising properties for CO2 capture with the high heats of adsorption and capacity (6–9 mmol g−1 at 10 bar), well retained or even promoted at lower pressures (<3.5 mbar at 303 K).128
Schematic of the one-pot facile synthesis of ordered mesoporous carbons. Reproduced from ref. 127 with permission from the Royal Society of Chemistry.
Schematic of the one-pot facile synthesis of ordered mesoporous carbons. Reproduced from ref. 127 with permission from the Royal Society of Chemistry.
In summary, the templating method allows us to effectively fabricate porous carbons with specific porous structures and surface chemical properties, which are suitable for CO2 adsorption. However, there are also some weaknesses. For instance, hard templates are usually prepared by a complicated production process, but are sacrificed after carbonization. Although the soft template method avoids the procedure of removing the templates after carbonization, only a few soft templates are successfully developed due to the requirement of thermal stability. Moreover, the templated carbons usually lack microporosity, which is the really efficient type of porosity for CO2 capture.
1.2.5 Combined Method of Templating and Activation
Recently, hierarchical porous carbons (HPCs) have attracted much attention. Hierarchical porous carbons possess complex network structures formed by two or more levels of pores, which include mainly micro-mesopores and micro-meso-macropores and so on. These carbons have shown a better performance in many application areas than carbons with a single porosity, on account of the improved mass transport facilitated by the large-sized pores and the high surface area and pore volume from the micro-/mesopores. For CO2 capture, the combination of micropores and mesopores can simultaneously promote the adsorption and diffusion of CO2 molecules.
To date, several approaches have been proposed for the synthesis of hierarchical porous carbons. One of the main techniques is a dual templating strategy, in which two templates with dimensions at different length scales are combined to generate multimodal pores. Either two hard templates164 or a combination of hard and soft templates165,166 can be employed in this technique. However, the biggest shortcoming of this strategy is its low efficiency in developing microporosity, especially ultramicroporosity, which is crucial for gas adsorption. Another commonly-used method to synthesize hierarchical porous carbons is a combination of a templating method and activation, in which mesopores and micropores are generated by the two different mechanisms respectively.167 This combined technique could be further classified into a two-step process and a one-step process.
1.2.5.1 Two-step Process
As mentioned above, a combination of a templating method and activation is one of the most commonly-used routes to prepare HPCs. This method usually contains two steps: firstly, the construction of mesopores by a template method; then, physical activation or chemical activation of the templated carbons to generate micropores. Many groups have synthesized various hierarchical porous carbons by choosing appropriate templates, carbon precursors and activation conditions.22,168,169
Marta Sevilla et al. designed and prepared hierarchical porous carbons through the KOH activation of two ordered mesoporous carbons of CMK-3 and CMK-8, which were respectively obtained by using hexagonal SBA-15 and cubic KIT-6 ordered mesostructured silica as hard templates.98 The process of KOH activation was carried out at different temperatures in the 600–800 °C range. Textural characterization of these activated carbons showed that they had a dual porosity made up of mesopores derived from the templated carbons and micropores generated during the chemical activation step. As a result of the activation process, there was an increase in the surface area and pore volume from 1020 m2 g−1 and 0.91 cm3 g−1 for the CMK-8 carbon to a maximum of 2660 m2 g−1 and 1.38 cm3 g−1 for the sample activated at 800 °C (KOH–CMK-8 mass ratio of 4). By comparison, various samples exhibited similar CO2 uptake capacities of around 3.2 mmol g−1 at 298 K no matter what type of templated carbon was used as a precursor. Through analysis of the CO2 adsorption quantities of different porous carbons, they found that CO2 capture capacity depends on the presence of narrow micropores (<1 nm) rather than on the surface area or pore volume of the porous carbons. Furthermore, it was found that these porous carbons exhibited a high CO2 adsorption rate, a good selectivity for CO2–N2 separation and they can be easily regenerated.98
Luis et al. demonstrated a novel strategy to synthesize hierarchical porous carbons with large pore volumes, high surface areas, and more importantly, tunable micro/meso/macro porosities. The approach is based on a combination of ice templating, colloidal silica and CO2 activation for generating interconnected macro-, meso-, and microporosity, respectively (Figure 1.16). During the synthesis, the range, size and extent of porosity can be easily controlled with different conditions. The synthesis was simple, green, reproducible, and used widely available and inexpensive starting materials, all of which made the process highly scalable. It produced a good CO2 capture capacity of 4.2 mmol g−1. More importantly, for dilute CO2 (10% CO2–90% N2 gas mixture), the sorption capacity measured under dry and moist conditions was the same, which was hardly affected by the presence of moisture.167
Schematic showing the combination of ice templating, colloidal silica and CO2 activation. Reproduced from ref. 167 with permission from the Royal Society of Chemistry.
Schematic showing the combination of ice templating, colloidal silica and CO2 activation. Reproduced from ref. 167 with permission from the Royal Society of Chemistry.
Many studies have proved that nitrogen-doping is an effective strategy to enhance the CO2 adsorption capacities of porous carbon adsorbents. However, to achieve a high doping level of nitrogen and a significant porosity in the carbon material simultaneously remains a challenge. Li et al. explored a facile approach to construct three-dimensional macroporous nitrogen-doped carbons with periodic ordered macropores and further etched micropores on the frameworks by KOH activation.169 These hierarchical porous nitrogen-doped carbons were partly graphitic and possessed relatively high nitrogen contents of about 14.45%. The CO2 adsorption capacities of these samples were about 2.69 mmol g−1 at 298 K and 3.82 mmol g−1 at 273 K at 1 bar with an extraordinarily high CO2–N2 selectivity of 134 calculated from the IAST method. Such an unprecedented CO2–N2 selectivity was largely associated with the unusually high N content and the partially graphitic framework of this material.
1.2.5.2 One-step Process
The one-step method is simple and practicable, and is an important way to synthesize hierarchical porous carbons. Hao et al. designed a series of poly(benzoxazine-co-resol)-based porous carbon monoliths with multiple-length-scale porosity (macro-, meso-, and micropores), a nitrogen-containing framework (polar surface), and remarkable mechanical strength (Figure 1.17).170 Interconnected mesopores (cubic Im3m symmetry) were formed by the assembly of poly(benzoxazine-co-resol) under the direction of surfactant Pluronic F127, while co-continuous macropores were formed due to polymerization-induced phase separation. Herein, organic amines were used as both catalyst and nitrogen source. The nature of the organic amines strongly influences the assembly of the mesostructure. It is reported that protic organic bases, such as DMA (dimethylamine), EDA (ethylenediamine), and DAH (1,6-diaminohexane), favored the formation of ordered mesopores in carbon monoliths. In contrast, aprotic organic bases (dimethylamine and triethylamine) resulted in the formation of microporous carbons due to the lack of hydrogen bonds between these amines and F127 molecules. HCM-DAH-1 carbon gave a CO2 adsorption capacity of 3.3 mmol g−1 at 0 °C and 780 mm Hg, a high CO2–N2 selectivity of 28, and high isosteric heats of adsorption of 21.1–35.9 kJ mol−1, indicating that the prepared carbon preferentially interacted with CO2. Further CO2 activation at 900 °C significantly increased the specific surface area and newly generated micropores with a size of 0.9–1.3 nm; as a result, the CO2 capacities increased to 4.9 mmol g−1. However, the selectivities of CO2-to-N2 and isosteric heats of adsorption decreased after the additional activation, which may be due to the widening of the microporosity.
Poly(benzoxazine-co-resol)-based porous carbon monoliths for CO2 capture. Reprinted with permission from G.-P. Hao, W.-C. Li, D. Qian, G.-H. Wang, W.-P. Zhang, T. Zhang, A.-Q. Wang, F. Schueth, H.-J. Bongard and A.-H. Lu, J. Am. Chem. Soc., 2011, 133, 11378, Copyright 2011 American Chemical Society.
Poly(benzoxazine-co-resol)-based porous carbon monoliths for CO2 capture. Reprinted with permission from G.-P. Hao, W.-C. Li, D. Qian, G.-H. Wang, W.-P. Zhang, T. Zhang, A.-Q. Wang, F. Schueth, H.-J. Bongard and A.-H. Lu, J. Am. Chem. Soc., 2011, 133, 11378, Copyright 2011 American Chemical Society.
Liu et al. reported the one-step synthesis of nitrogen-doped graphene-like meso-macroporous carbons as highly efficient and selective adsorbents for CO2 capture. Graphene-like meso-macroporous carbons (GMCs) with high nitrogen contents and controllable nitrogen sites were synthesized by one-step carbonization of dicyandiamide or melamine with glucose. The unique layered structure and the presence of abundant meso-macropores in the GMCs largely enhance the degree of exposure and accessibility of anchored nitrogen sites to CO2, which provides the GMCs with an excellent performance for the selective capture of CO2 (about 101 calculated according to the IAST).171
Tian et al. reported a method of direct carbonization of algae biomass to generate hierarchical porous carbons. Algae biomass is a fantastic precursor for fabricating porous carbon due to its natural composition. Enteromorpha prolifera, which was selected as a carbon source, was converted into carbons with evident hierarchical micro-mesoporous structures through a simple direct pyrolysis of freeze-dried algae. The formation process for hierarchical pores was preliminary discussed. The surface areas of the obtained carbons were only close to 450 m2 g−1 with 70% areas inherited from mesopores, resulting in high density materials with the major pore sizes at around 1.5 and 5.0 nm. Moreover, the carbons show a CO2 uptake of 6.48 mmol g−1 at 20 bar and 25 °C.172
In this section, we have shown that researchers have fabricated an enormous number of porous carbons as CO2 adsorbents. As we have reviewed, the porosity (e.g., single- or multi-level, narrow-distributed or wide-distributed, ordered or disordered, microporous or mesoporous, etc.) and surface chemistries (pristine or doped) could be engineered well. In general, their synthesis strategies could be roughly categorized into three types: activation, templating, and a combination of both. The CO2 sorption capacities of these materials could be comparable to those of other well-known porous solid adsorbents, such as MOFs, ZIFs, zeolites, and others. Porous carbons have been one of the most promising candidates for CO2 adsorbents, and will be more and more important in the field of CO2 capture.
1.3 Graphene-based Porous Materials
Graphene, the world's first two-dimensional (2D) material, has captured the attention of scientists, researchers, and industry worldwide. In simple terms, graphene is a thin layer of an sp2-bonded carbon sheet. This material was first isolated and characterized in 2004 by Andre Geim and Konstantin Novoselov at the University of Manchester. In 2010, they received the Nobel prize in physics for their work.
As we know, graphene has many advantages, such as a high specific surface area, superior electron mobility, easy self-assembly into three-dimensional (3D) macroscopic materials with controlled microstructures, high mechanical, chemical, thermal, and electrochemical stabilities, etc. With these amazingly attractive properties, graphene already exhibits tremendous promise in many applications: nanoelectronics,173 supercapacitors,174 fuel cells,175 batteries,176 solar cells,177 water filtration,178 gas sorption, separation and storage,179,180 and sensing,181 etc. In recent years, there has also been an enormous effort to explore graphene and its oxygen-containing derivative, graphene oxide (GO), for CO2 capture applications. The latest progress with both the experimental and theoretical research into graphene-based CO2 adsorbents has been summarized in some recent reviews.182,183 In this section, we mainly focus on the experimental reports of graphene-based porous materials for CO2 capture.
As gas adsorption is highly dependent on the pore texture of the adsorbents, it is highly desirable that the 2D graphene sheets are engineered further in specific ways to obtain a tailored porosity with a high surface area and pore volume. Accordingly, several techniques have been adapted to develop graphene-based porous adsorbents for CO2 capture.
1.3.1 Graphene-based Adsorbents by Chemical Activation
Chemical activation with KOH is an efficient technique to introduce pore texture into graphene. In 2011, Zhu et al. first reported the KOH activation of microwave exfoliated GO (MEGO).174 The KOH activation generated a much developed porosity in the product carbon. The specific surface area of the activated MEGO could be controlled by the ratio of KOH–MEGO. The optimized activated MEGO had a well-defined micro-mesopore-sized distribution with a very high BET surface area and pore volume, up to 3100 m2 g−1 and 2.14 cm3 g−1, respectively. This activated MEGO carbon could provide a large and accessible surface area for charge accommodation, and, therefore, exhibited an outstanding capacitive performance in organic and ionic electrolytes to achieve high gravimetric energy densities.
In 2012, Srinivas et al. reported the synthesis of a range of high surface area graphene-oxide-derived carbons (denoted as GODCs) and their applications toward carbon capture and methane storage.184 GODCs are prepared by chemical activation with KOH from thermally exfoliated GO and solvothermally reduced GO precursors. The authors explored the effects of activation temperature and KOH concentration upon the development of porosity in the resultant carbons. At low activation temperatures and KOH concentrations, the activation mainly generated pore development in the microporous region, while increasing the activation temperature and KOH concentration resulted in a large proportion of pore development in both the microporous and mesoporous regions (>2 nm). The optimum activation conditions were found to be 800 °C and a GO–KOH ratio of 1 : 9 to obtain a porous carbon (GODCsol-800) with the maximum specific surface area and total pore volume of 1900 m2 g−1 and 1.65 cm3 g−1, respectively. The results of CO2 sorption displayed that the high-pressure CO2 adsorption trend is more or less linearly dependent on the BET surface area, total pore volume, and average pore size, while the low pressure (1 bar) adsorption behavior is almost independent of the surface area, but is closely related to the narrow pore size distribution in the micropore region. The optimized sample of GODCsol-800 exhibited a high CO2 adsorption capacity of 72.1 wt% at 300 K and 20 bar, as well as a high methane adsorption capacity of 17.5 wt% at 300 K and 35 bar.
1.3.2 Graphene-based Adsorbents by Physical Activation
Physical activation with steam and CO2 has also been employed to fabricate porous graphene-based materials for CO2 capture. Sui et al. developed porous graphene-based carbons through physical activation of graphene aerogels using steam as the activating agent.185 The activation temperature plays a critical role in determining the BET surface area and pore volume of the resultant carbons, and the optimal activation temperature to obtain the highest BET surface area and pore volume is 850 °C. A low activation temperature (750 °C) was less efficient at enhancing the porosity due to the slow reaction rate between graphene aerogel particles and steam. Meanwhile, above 850 °C, the samples also showed a decreased specific surface area and pore volume due to the high burn-off of graphene sheets, thus resulting in the destruction of the porous structure and a lower yield. The steam-activated graphene aerogel exhibited a high specific surface area (830–1230 m2 g−1), an abundant large pore volume (2.2–3.6 cm3 g−1), and excellent thermal stability. The optimized SAGA-850 showed a CO2 adsorption capacity of 2.45 mmol g−1 at 1 bar and 273 K, much higher than that of the non-activated GA sample (1.45 mmol g−1) under the same conditions.
Chowdhury and Balasubramanian prepared graphene-based porous carbons though CO2 activation using reduced graphene oxide as a precursor (Figure 1.18).186 By increasing the activation temperature of CO2, the specific surface area, micropore pore volume and total pore volume increase, resulting in an improvement of CO2 uptake. Specifically, the adsorbent material obtained at an activation temperature of 950 °C (i.e., a-RGO-950) exhibited the largest specific surface area (above 1300 m2 g−1), the highest pore volume (over 1 cm3 g−1), and a well-defined bimodal micro-mesoporous structure. This adsorbent material displayed a good gravimetric CO2 uptake (3.36 and 2.45 mmol g−1 at 0 °C, 1 bar and 25 °C, 1 bar, respectively), rapid adsorption kinetics, as well as stable and readily reversible adsorption–desorption cycling behavior at room temperature. Moreover, a-RGO-950 exhibited excellent Henry law CO2–N2 selectivities of 162 and 253 under conditions pertinent to CO2 capture from the dry flue gas steam of a coal-fired (75% N2 and 15% CO2) and natural-gas-fired (80% N2 and 5% CO2) power plant, respectively.
Illustration of the CO2 activated reduced graphene oxide (rGO). Reprinted with permission from S. Chowdhury and R. Balasubramanian, Ind. Eng. Chem. Res., 2016, 55, 7906, Copyright 2016 American Chemical Society.
Illustration of the CO2 activated reduced graphene oxide (rGO). Reprinted with permission from S. Chowdhury and R. Balasubramanian, Ind. Eng. Chem. Res., 2016, 55, 7906, Copyright 2016 American Chemical Society.
Similarly, Xia et al. prepared a series of hierarchical porous graphene-based carbons (HPGCs) by CO2 activation of graphite oxide.187 HPGC-850, which was prepared by 2 hours of CO2 activation at 850 °C, possessed the highest specific surface area and micropore volume, thus exhibited the highest CO2 sorption capacity of 1.76 mmol g−1 at 274 K and 1 bar, as well as the highest H2 sorption capacity of 3.76 mmol g−1 at 77 K and 1 bar.
1.3.3 Graphene-based Adsorbents by Other Techniques
Other techniques, like hydrogen-induced exfoliation,188 and covalent functionalization,189 are also used to introduce pore structures into graphene materials for CO2 capture. For instance, porous graphene materials prepared via hydrogen-induced exfoliation of graphite oxide exhibited a maximum sorption capacity of 21.6 mmol g−1 at 11 bar and 25 °C.188 The physical adsorption nature of CO2 in the prepared graphene material was confirmed using a Fourier transform infrared spectroscopy (FTIR) study. Beyond the normal hydroxyl (3435 cm−1), carboxyl (1726 cm−1) and carbonyl (1173 cm−1) peaks, a new peak was observed at 2324 cm−1 in the IR spectra. This peak corresponds to the asymmetric stretching of CO2, implying physisorption of CO2 onto graphene sheets.
Chowdhury reported the CO2 capture performance of thermally treated graphene oxides.190 As the thermal treatment temperature increased, the pore texture of the graphene sheets became more developed. GPN-800 treated at 800 °C processed the highest specific surface area (484 m2 g−1) and micropore volume (0.094 cm3 g−1), resulting in the highest CO2 uptake of 2.9 mmol g−1 at 0 °C and 1 bar. Kumar designed two pillared porous graphene frameworks (PGFs) by linking reduced graphene oxide layers with 1,4-diethynylbenzene (PGF-1) and 4,4′-diethynylbiphenyl (PGF-2) via a C–C coupling reaction.189 Both frameworks show high CO2 uptakes of 112 wt% for PGF-1 and 60 wt% for PGF-2 at 195 K and 0.85 atm.
Besides, graphene has been widely used to fabricate composite materials as a carrier support due to its 2D structure with a high surface area. Lu's group reported the preparation of porous carbon nanosheets (PCNs) with precisely tunable thicknesses in which GO played the role of shape-directing agent. The resorcinol-formaldehyde resins grew in situ on the GO sheets due to the bridging effect of asparagine and were converted into porous carbon sheets. The thickness of the carbon sheets was tuned from 20 to 200 nm according to the mass ratio of the resin–GO. At 25 °C and 1 bar, the maximum CO2 uptakes of PCN-9.9, PCN-17, and PCN-71 with thicknesses of 9.9, 17, and 71 nm, respectively, were 2.02, 2.36, and 2.88 mmol g−1. Moreover, these porous carbon sheets showed a good ability to separate CO2 from simulated flue gas (a water-saturated CO2/N2 stream) under dynamic conditions; the CO2 capacity reached 0.28 mmol g−1 at a CO2 concentration of 4 v%. PCN-17 could stably work for 200 cycles in total under a CO2–N2 gas stream of 14 : 86 v% (Figure 1.19). Kim and colleagues reported various N- or S-doped porous carbons by using graphene/polypyrrole,31 rGO/polyaniline,192 rGO/polyindole,193 and rGO/polythiophene37 as carbon precursors and KOH as an activating agent. The graphene in the carbon precursor is believed to increase the contact area between the KOH activator and carbon precursors. As a result, the synthesized carbons featured a high surface area, a large pore volume, and developed microporosity, and captured a large amount of CO2 under ambient conditions (>4 mmol g−1), a value much higher than those of the porous graphene materials discussed earlier.
Schematic of the formation of the PCNs. (a) The negatively charged GO sheet; the picture (right) shows its Tyndall phenomenon. (b) Positively charged amino acids, equally dispersed at the molecular level on both surfaces of the GO. (c) In situ co-polymerization of pre-adsorbed asparagine, resorcinol and formaldehyde. (d) The polymer layer transformed into a microporous carbon layer during pyrolysis in Ar. Reproduced from ref. 191 with permission from The Royal Society of Chemistry.
Schematic of the formation of the PCNs. (a) The negatively charged GO sheet; the picture (right) shows its Tyndall phenomenon. (b) Positively charged amino acids, equally dispersed at the molecular level on both surfaces of the GO. (c) In situ co-polymerization of pre-adsorbed asparagine, resorcinol and formaldehyde. (d) The polymer layer transformed into a microporous carbon layer during pyrolysis in Ar. Reproduced from ref. 191 with permission from The Royal Society of Chemistry.
Hybrid CO2 adsorbents, like PEI/graphene,194 polyaniline/graphene,195 LDH/graphene,196 etc., have also been reported to exhibit excellent CO2 capture performances, in which the supporting effect of the graphene ensures the high loading content of the active materials, and improves the stability and CO2 capacities. These hybrid materials will be briefly discussed in Section 1.5. Clearly, various graphene-based materials have been investigated in the CO2 capture field. The present studies have demonstrated the promise of graphene-based materials for selectively capturing and isolating CO2 from flue gas. However, this investigation is still in its infancy, and further research is needed.
1.4 Carbon Nanotubes
Carbon nanotubes (CNTs) are allotropes of carbon with a cylindrical nanostructure. These materials were first discovered by Ijima in 1991 as minority by-products of fullerene synthesis. According to the number of carbon walls, CNTs are categorized as single-walled nanotubes (SWNTs) and multi-walled nanotubes (MWNTs). The inner diameter of CNTs can vary from approximately 1 nm for SWCNTs to over 10 nm for MWCNTs. Considering the uniform inner diameter and almost defect-free wall of CNTs, the interaction of gas molecules with the inner pores of the CNTs can be described by a smooth potential energy surface. Recent computational simulations have proved that the smoothness of the inner pores make gas molecules diffuse and transport rapidly through the CNTs. For example, the transport rates of light gases, such as H2, CH4, and N2, in CNTs are orders of magnitude faster than in microporous materials with comparable pore sizes.197,198 This feature makes CNTs ideal candidates for the selective sorption and separation of gases, such as CO2 capture from flue gas.
Lu et al. carried out a comparative study of CO2 capture by CNTs with inner diameters <10 nm, activated carbons, and zeolites.199 Raw CNTs showed a CO2 sorption amount of 22.7 mg g−1 under an influent CO2 concentration of 10% at 25 °C, which is slightly less than activated carbons (24.9 mg g−1) but larger than zeolites (19.0 mg g−1) under the same conditions. After modification with 3-aminopropyltriethoxysilane, the modified CNTs showed an enhanced adsorption capacity of CO2 of up to 40.9 mg g−1 and 96.3 mg g−1 at CO2 concentrations of 10% and 50%, respectively. The mechanism of CO2 adsorption on these adsorbents appears to be mainly attributable to physical force.
Most studies of pure CNTs for CO2 capture are simulations, including quantum calculations (i.e., density functional theory (DFT) and ab initio simulations),200–202 grand canonical Monte Carlo (GCMC) simulations,203,204 and molecular dynamics (MD) simulations.205 DFT calculations predicted that zigzag SWNTs possessed stronger bonds with CO2 molecules compared to armchair and chiral ones, and N-doping could increase the binding energies between zigzag SWNTs and CO2, suggesting that zigzag SWNTs show greater promise as a means of CO2 capture.200 Based on Monte Carlo simulations, Liu and Bhatia found that increasing the nanotube diameter from 1.36 nm (10, 10) to 2.03 (15, 15) led to enhanced CO2 capacity at 5 bar due to the increase of accessible volume, while the change in chirality had a negligible effect.205 Furthermore, the efficiency of CO2 capture from flue gas can be significantly improved by pre-adsorbing small clusters of water in carbon nanotubes.205 Rahimi et al. simulated adding a positive charge to the carbon nanotubes, which caused a significant increase in adsorption (up to 35% at a pressure of 1.88 bar), while the gas adsorption decreased by up to 15% on negatively charged carbon nanotubes.206 The increase or decrease of adsorption upon charging was attributed to the change in potential energy for the interactions between the individual CO2 molecules and the nanotubes.
Owing to their extraordinary thermal conductivity, mechanical, and electrical properties, carbon nanotubes find applications as additives to various structural materials. Jin et al. fabricated carbon composite monoliths by CO2 activation of a composite of a commercial phenolic resin and carbon nanotubes.207 CNTs in the composite monoliths played an important role in improving pore structures and the CO2 capacity of CNT-modified CCMs. CNT-modified carbon monolith samples were much more reactive with CO2 during activation; they exhibited significant burn-off within a much shorter activation time, thus achieving a well-developed porosity, especially a narrow microporosity. CPD-30 with a burn-off of 25.5 wt% gave the highest CO2 uptake of 3.5 mmol g−1 at 298 K and 1 atm, while CPD-15 with a burn-off of 16.7 wt% had the highest CO2 uptake of at 1.1 mmol g−1 at 298 K and 0.15 atm, more than twice that of the resin alone, Res-60 (only 0.55 mmol g−1). Mixed matrix membranes (MMMs) based on CNTs dispersed inside a polymer matrix have been well-developed and characterized for CO2 separation measurements. For instance, Ahmad et al. fabricated MMMs from cellulose acetate with β-cyclodextrin-functionalized MWCNTs (MWCNTs-F) using a wet phase inversion technique.208 The results showed that MMMs with 0.1 wt% loadings of functionalized MWCNTs demonstrated enhanced permeance and selectivity towards the separation of CO2/nitrogen (N2), as well as outstanding mechanical properties with a high tensile strength of 16 MPa. This behavior results from the enhanced compatibility of MWCNTs-F within the polymer chain segments, which increased the sufficient interlayer spacing.
As discussed, CNTs have attracted much attention in the field of CO2 capture. These materials could directly serve as adsorbents, as well as supporters/additives for the fabrication of hybrid adsorbents. However, the studies reported have mainly focused on simulations; more experimental studies should be performed in the future.
1.5 Carbon-based Hybrid Adsorbents
CO2 sorption on pristine carbon materials is essentially “physisorption”, which is very sensitive to temperature and pressure, and is usually low-selective. Generally, the sorption capacities of CO2 on carbon materials will significantly decrease as the sorption temperature increases or the sorption pressure decreases. For example, the CO2 uptakes of the previously reported porous carbons at 0.15 bar and 25 °C (typical CO2 partial pressure in flue gas) are usually lower than 1.5 mmol g−1. Reducing the pore size to extremely small micropores (<0.5 nm) and introducing polar surface chemistry can elevate the CO2 uptake to about 2 mmol g−1 at low pressure or high temperature.88 However, this will lead to a serious obstruction to gas diffusion kinetics. Synthesis of carbon-based hybrid sorbents is an effective strategy to enhance CO2 capture performance. Mesoporous carbons, carbon nanotubes, and graphene are commonly used because these materials possess larger pore sizes, an opened pore texture and a high surface area, ensuring a high mass loading and good dispersion of the incorporated components. Herein, carbon-based hybrid sorbents are classified into carbon–organic hybrid sorbents and carbon–inorganic ones according to the nature of the incorporated component.
1.5.1 Carbon–Organic Hybrid Adsorbents
Organic amines are the most commonly used functional components for carbon–organic hybrid sorbents. Various amines, such as monoamines, diamines, triamines, polyamines and amine-containing polymers, have been studied for this purpose. In the last few years, several excellent review papers have been published and provide a good perspective towards amine-functionalized solid adsorbents for CO2 capture and separation.209–211 Generally, there are two ways to incorporate the amines: (1) impregnation, and (2) grafting.
With impregnation, the amine compounds are dispersed inside the pores of the carbon support, and thereby, they produce an enhanced CO2 capture performance relative to the mass loading of the bulk amines. In theory, the more amines that are loaded, the higher the CO2 capture capacities of the amine-impregnated carbon materials. The dispersion of the incorporated amines is also important. The excess loading of amines may result in a decline of the amine utilization ratio due to a lengthening of the diffusion distance and the agglomeration of the amine molecules.212,213
Polyethyleneimine (PEI) is most commonly used due to its high CO2 uptake/release capacity and good stability. For example, Wang et al. developed highly efficient CO2 sorbents by loading PEI on mesoporous carbons.212 As the PEI loading increased, the utilization ratio of the PEI first increased and then decreased, and the optimal PEI loading is determined to be 65 wt% with a CO2 sorption capacity of 4.82 mmol g−1 in 15% CO2/N2 at 75 °C.212 These sorbents could work over a wide CO2 concentration range from 5% to 80%. Furthermore, moisture was found to have a promoting effect on the sorption separation of CO2. This group also reported a sorbent of PEI-impregnated millimeter-sized mesoporous carbon spheres for post-combustion CO2 capture (Figure 1.20).214 Under the optimal conditions of using a high molecular weight PEI and polyethylene glycol (PEG) loading of 20 wt%, the adsorbent exhibited a high equilibrium adsorption capacity of 187.5 mg g−1 (4.26 mmol g−1) for 15% CO2 at 75 °C at a relative humidity of 60%. Noticeably, these adsorbents could be regenerated by a novel electric swing adsorption (ESA) operation due to the good electrical conductivity of the carbon support.
Schematic of the adsorption column for electric swing adsorption. Reprinted from M. Wang, L. Yao, J. Wang, Z. Zhang, W. Qiao, D. Long and L. Ling, Adsorption and regeneration study of polyethylenimine-impregnated millimeter-sized mesoporous carbon spheres for post-combustion CO2 capture, Appl. Energy, 168, 282–290, Copyright 2016, with permission from Elsevier.
Schematic of the adsorption column for electric swing adsorption. Reprinted from M. Wang, L. Yao, J. Wang, Z. Zhang, W. Qiao, D. Long and L. Ling, Adsorption and regeneration study of polyethylenimine-impregnated millimeter-sized mesoporous carbon spheres for post-combustion CO2 capture, Appl. Energy, 168, 282–290, Copyright 2016, with permission from Elsevier.
With grafting, the amine groups are chemically bound to the carbon supports. In 2011, Houshmand et al. reviewed this strategy well.209 In their review, the methods used for grafting are roughly categorized into amination, silylation with aminosilanes, nitration followed by reduction, anchoring diamines/polyamines, anchoring halogenated amines, surface-initiated polymerization of ethyleneimine and its derivatives, and plasma treatment etc. To compare, the interactions between the grafted amines and carbon supports are strong covalent bonds, while the interactions between the impregnated amines and carbon supports are mainly weak interactions, such as dipole–dipole interactions, van der Waals forces, or hydrogen bonding. Therefore, amine groups incorporated by grafting are more stable and do not desorb during regeneration.209 Nevertheless, grafting suffers the shortcoming of a complicated synthesis process. In addition, with both strategies, the porosities of the carbon supports should be tailored to be suitable for the amine compounds selected.
Besides the commonly used mesoporous carbon materials, graphene and carbon nanotubes have also been used as supports. For instance, Liu et al. prepared hybrid sorbents by covalent grafting of PEI on hydroxylated 3D graphene (HG).196 The optimal HG-PEI hybrid sorbent exhibited a high CO2 adsorption capacity up to 4.13 mmol g−1 at 1 atm of dry CO2. PEI-impregnated graphene-silica sheets also showed an excellent CO2 adsorption capacity (171 mg g−1 at 75 °C/100 kPa) and good cycle stability (>20 cycles).194 Lu et al. grafted 3-aminopropyltriethoxysilane (APTS) onto MWCNTs, and studied the CO2 capture behaviors of the amine-loaded MWCNTs.215,216 The adsorption capacity of CNT(APTS) was significantly influenced by the presence of water vapor and reached a maximum of 2.45 mmol g−1 at 2.2% water vapor.215 In addition to the high surface area and opened pore texture, which promotes diffusion of CO2 to the active adsorption sites, the high thermal conductivity of graphene and CNTs allowed a fast transfer of heat and avoided degradation of the organic amines.
Other organics, such as ionic liquids and quinones, are also used to fabricate carbon-based hybrid sorbents. Tamilarasan reported that polymerized ionic liquid functionalized graphene showed an adsorption capacity that was 22% higher than graphene, while ionic liquid functionalization improved it only by 2%.217 Wang et al. developed a hydroquinone (quinone)-functionalized hierarchical porous carbon sorbent via the Friedel–Crafts reaction.218 The hydroquinone-grafted carbon gave a CO2 adsorption capacity of 5.41 mmol g−1 at 1 bar and a CO2–N2 selectivity of 26.5, higher than those of pristine porous carbon.
1.5.2 Carbon–Inorganic Hybrid Adsorbents
Fabrication of carbon–inorganic hybrid materials is another approach to enhance CO2 uptake. Various carbon materials, such as porous carbon materials,219 carbon nanotubes,220 and graphene and its derivative,221 are used as solid supports. The inorganic sorbents are mostly metal compounds, including metal coordination polymers (MOFs,222,223 ZIFs224 ), metal oxides (MgO,221 CaO,225 ZnO226 ), alkali metal carbonates,227 layered double hydroxides (LDHs),228 etc. According to a recent review by Prof. Wang, these carbon–inorganic hybrid sorbents could be classified into intermediate-temperature (200–400 °C) and high-temperature (>400 °C) solid CO2 sorbents, and have been summarized well.229
In summary, a large number of carbon-based hybrid sorbents have been studied. Due to the strong affinity of incorporated components to CO2, these sorbents have shown high CO2 capture capacities at high sorption temperatures or low CO2 concentrations. Carbon supports have been proved to have significant promoting effects on CO2 capture, which could be mainly because: (1) the high surface area and pore volume of carbon supports improve the dispersion of active materials while ensuring a high mass loading; (2) the opened pore texture facilitates the diffusion of CO2 molecules to the active adsorption sites; (3) the high dispersion also guarantees a high adsorption efficiency and good cyclic stability.
1.6 Effect of Carbon Structure on CO2 Adsorption
A large number of studies have shown that there are two main strategies to improve the CO2 adsorption capacity of carbons. One simple strategy is to introduce polarity by chemical doping in the carbon framework to enhance its affinity with CO2. Another approach that has been investigated to enhance CO2 capture capacities is to tailor pore structures to make them suitable for CO2 adsorption. In this section, we pay attention to studies of the effect of carbon structure on CO2 adsorption, which are of use for the development of high-performance carbon-based adsorbents.
1.6.1 Pore Size Effect
1.6.1.1 Analysis of Porosity
The IUPAC (International Union for Pure and Applied Chemistry) has proposed a classification of pores based on pore sizes. Pores are generally classified into micropores (<2 nm), mesopores (2–50 nm), and macropores (>50 nm). Micropores are further divided into ultramicropores (<0.7 nm) and supermicropores (0.7–2 nm). Since the porosity of carbons is responsible for their adsorption properties, the analysis of the porosity, especially the microporosity, is very important to foresee the behavior of these porous solids in CO2 capture applications. Pores in carbon materials could be identified by several techniques depending on their type and size, such as gas adsorption (opened pores < 100 nm), mercury porosimetry (opened macropores), small angle scattering (SAS) (latent pores, including closed pores), transmission and scanning electron microscopy (TEM and SEM) (extrinsic pores on the surface of carbon materials), etc.
Physical adsorption of various gases is, undoubtedly, the most widely used technique. N2 is the most widely used adsorptive. However, N2 adsorption at 77 K is not suitable for the characterization of narrow microporosity (size < 0.7 nm), because diffusion kinetics of N2 molecule is very slow in ultramicropores at a liquid N2 temperature of 77 K. To overcome this problem, the use of CO2 sorption at 273 K has been proposed. The higher temperature of CO2 adsorption results in a larger kinetic energy of the molecules, which are able to enter the narrow pores. However, CO2 adsorption analysis of the whole porosity should be performed at high pressures due to its high saturated vapor pressure (P0) at 273 K (34.7 bar). In practice, the combination of N2 and CO2 adsorption allows us to obtain comprehensive information about the porosity.
In order to evaluate the specific surface area and PSD, various methods have been applied to gas isotherms, including, but not limited to, the BET (Brunauer–Emmett–Teller) method, D–R (Dubinin–Radushkevich) plot, BJH (Barret–Joyner–Halender) method, t plot, HK (Horvath–Kawazoe) method, and the DFT (density functional theory) method. Porous carbon materials, especially activated carbons, typically possess wide PSDs ranging from ultramicropores, supermicropores and mesopores to macropores. The NLDFT (non-linear density functional theory) method has been widely applied and featured in a standard by ISO-15901-3.230 The standard NLDFT model assumes an idealized flat graphitic-like surface. Due to the heterogeneity in the chemistry and structure of the carbon materials, this model may not describe all the subtleties of a real carbon surface. The QSDFT (quenched solid-state density functional theory) model takes into account carbon surface roughness features. This model provides a better representation of the surface properties, including surface roughness features,231,232 and has been recommended in recent years.106,233 However, QSDFT models are not yet available for some of the measurement instruments (e.g., micromeritics instruments) and sorption conditions (e.g., CO2, 0 °C). The NLDFT method is still widely used to evaluate porosity from N2 sorption measurements.
Micropores, especially, feature in interactions between gas adsorbate molecules and adsorbents, and there are still no definitive methods to determine the PSD in the micropore region. For the analysis of microporosity, Dubinin and Radushkevitch have developed an equation:
where V (cm3 g−1) is the volume filled at a temperature T and a relative pressure (P/P0), V0 (cm3 g−1) is the micropore volume, and k and β are constants for the pore structure and the affinity coefficient (e.g., β = 0.35 for CO2). This equation, based on the postulation that the mechanism for adsorption in micropores is that of pore filling rather than layer-by-layer surface coverage, generally applies well to adsorption systems involving only van der Waals forces and is especially useful to describe adsorption on activated carbon.109,234 This equation has been used to calculate micropore volumes and to estimate the mean pore size widths (L) for both N2 and CO2 adsorption on different samples.88,89,235,236 The mean pore width could been estimated using the equation of Stoeckli et al. (eqn (1.15))237 when the characteristic energy (E0) is between 42 and 20 kJ mol−1:
Its validity has been tested by direct determination of the distribution in the range 0.35–1.3 nm. A well-defined linear D–R plot over the entire log2(P0/P) range indicates a uniform micro-PSD, while the non-linear one usually indicates a wide microporosity with various micropore sizes (Figure 1.21).238
Evaluating the mean micropore size from the slope of D–R plots of: (a) KAC (b) LiAC. Adapted from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Evaluating the mean micropore size from the slope of D–R plots of: (a) KAC (b) LiAC. Adapted from ref. 89, with permission from John Wiley and Sons, Copyright 2016 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
1.6.1.2 Micropore Filling Mechanism of CO2 Adsorption
In order to reveal the adsorption mechanism, isotherms are simulated using two typical adsorption models including Langmuir and Dubinin–Astakhov (D–A) models (Figure 1.22).
(a) CO2 sorption isotherm of a PANI-based porous carbon at 25 °C, the correlation of CO2 adsorption data with (b) the D–A model and (c) the Langmuir equation.
(a) CO2 sorption isotherm of a PANI-based porous carbon at 25 °C, the correlation of CO2 adsorption data with (b) the D–A model and (c) the Langmuir equation.
The well-known Langmuir isotherm model can be expressed by eqn (1.16).
where Q (cm3 g−1) denotes the amount adsorbed, Q0 (cm3 g−1) denotes the saturated amount adsorbed, P (kPa) denotes the equilibrium pressure, and b/kP−1 denotes the adsorption affinity.
A linear expression for the Langmuir equation is
Figure 1.22 shows the correlation of adsorption data with the Langmuir equation for a PANI-based activated carbon.105 It was found that the CO2 adsorption isotherms could not be fitted well by the Langmuir model.
The Dubinin–Astakhov (D–A) model is applied to the adsorption equilibrium of CO2 onto the activated carbons. The D–A equation is proposed to describe a wide microporosity as a general form of the D–R equation, which can be written as
or
where W denotes the volume adsorbed, W0 denotes the limiting micropore volume, E denotes the characteristic energy of the system, n denotes the heterogeneity parameter, R is the gas constant, T is the equilibrium temperature, and P0 is the saturated pressure. Apparently, the experimental data can be fitted well by the D–A equation (Figure 1.22).
This simulation shows evidence that CO2 adsorption on porous carbons can be depicted well by a micropore filling mechanism, not by monolayer adsorption. In other words, the CO2 sorption capacity should depend on the pore volume of a certain pore size range but not the specific surface area.
1.6.1.3 Pore Size Effect at Different Sorption Pressures
The main contribution of micropores, especially ultramicropores, to CO2 uptake, has been widely recognized. Therefore, it is necessary to finely explore the relationship between microporosity and CO2 capture capacity, which is significant for designing high-performance carbon adsorbents for CO2 capture. A pioneering study of the relationship between porosity and CO2 capacity was conducted by Gogotsi and his co-workers.106 They prepared a large number of different carbide-derived carbons (CDCs) derived from nano- and micrometer-sized precursors, with and without activation, and investigated their CO2 sorption performances. Various porosity parameters, including specific surface area, pore size, and pore volume, were correlated with the resulting sorption properties. They found that there was no clear dependency between the total pore volume or the volume of pores larger than 1 nm and CO2 capacities at 1 bar (Figure 1.23). Moreover, there is no clear trend correlating CO2 uptake and average pore size. Hence, they concluded that the CO2 sorption behavior at ambient conditions may be largely governed by the volume of micropores, similar to the result with H2 sorption.
The relationship of CO2 uptake at 1 bar (0 °C) versus total pore volume, specific surface area, volume of pores smaller than 0.8 nm, and the relationship of CO2 uptake at 0.1 bar (0 °C) versus volume of pores smaller than 0.5 nm. Adapted from ref. 106 with permission from the Royal Society of Chemistry.
The relationship of CO2 uptake at 1 bar (0 °C) versus total pore volume, specific surface area, volume of pores smaller than 0.8 nm, and the relationship of CO2 uptake at 0.1 bar (0 °C) versus volume of pores smaller than 0.5 nm. Adapted from ref. 106 with permission from the Royal Society of Chemistry.
They further correlated the micropore volumes of various pore sizes with CO2 uptake at different sorption pressures. At 1 bar, pores smaller than 0.8 nm were found to contribute the most to CO2 uptake and at 0.1 bar, pores smaller or equal to 0.5 nm were preferred (Figure 1.23). This result is regardless of carbon particle size and activation/non-activation. Coefficients of determination between CO2 uptake at a certain pressure (0.1–1.0 bar) and the cumulative pore volume of pores smaller or equal to a certain pore size (0.4–1.5 nm) were also obtained (Figure 1.24). With a lower total pressure, smaller pores contributed more to the measured CO2 uptakes.
Coefficient of determination between CO2 uptake from 0.1 to 1.0 bar and the volume of pores smaller than critical size. Reproduced from ref. 106 with permission from the Royal Society of Chemistry.
Coefficient of determination between CO2 uptake from 0.1 to 1.0 bar and the volume of pores smaller than critical size. Reproduced from ref. 106 with permission from the Royal Society of Chemistry.
1.6.1.4 Pore Size Effect at Different Adsorption Temperatures
Almost all of the studies on CO2 capture were carried out at low temperatures such as 0 °C and room temperature (Tables 1.1–1.3). However, industrial flue gas, the biggest contributor of CO2, is at a relatively high temperature (>70 °C). The sorption temperature significantly influences gas uptake on porous sorbents, and a high temperature always leads to a low gas uptake. Apparently, a well-designed porosity for carbon adsorbents is required for CO2 capture at a high temperature. An investigation into the relationship between CO2 uptake and the porosity of carbons at different sorption temperatures, especially a high temperature (e.g., 75 °C), is also necessary.
In 2012, our group prepared a series of N-doped porous carbon materials by using PANI as a carbon precursor via a pre-carbonization and post-KOH activation method.105 The micropore size distribution of the prepared carbons varied with the activation conditions (KOH–carbon ratio and activation temperature). These carbons showed very high CO2 uptakes of up to 1.86 and 1.39 mmol g−1 under 1 bar, 75 °C and 0.15 bar, 25 °C, respectively, which are amongst the highest of the known carbon materials for CO2 capture.
It is well known that the smaller the micropore, the stronger the adsorption potential. The smaller micropores will be the preferred spaces for adsorption of CO2 molecules, and could be filled to a higher degree due to their stronger adsorption potentials. Based on this view, the CO2 sorption capacity per pore volume is used as a factor reflecting the adsorption capability of porous carbons. The correlations of CO2 sorption capacities per pore volume with the volume fractions of micropores smaller than a critical size under different sorption conditions were investigated (Figure 1.25). At low temperatures (0 °C), the CO2 sorption capacity per pore volume is strictly linear with the volume fraction of the pores smaller than 0.80 nm in total pore volume (high correlation coefficient of 0.97, Figure 1.25), agreeing well with the results of Gogosti.106 Similar results are obtained at 25 °C and 75 °C, while this critical pore size (0.70 nm and 0.54 nm for 25 °C and 75 °C, respectively) decreases as the sorption temperature increases. Furthermore, the simulated lines plotted in Figure 1.25 are found to have an intercept of near zero, indicating that the small micropore is the most important (even exclusive) factor for CO2 uptake. In a word, the CO2 adsorption capacity of porous carbons is associated with the pores below a temperature-dependent size.
Linear fitting of the experimental data. A plot of CO2 sorption capacity per pore volume vs. the volume percentage of small micropores in total pores at adsorption temperatures of 0, 25 and 75 °C. Adapted from ref. 105 with permission from the PCCP Owner Societies.
Linear fitting of the experimental data. A plot of CO2 sorption capacity per pore volume vs. the volume percentage of small micropores in total pores at adsorption temperatures of 0, 25 and 75 °C. Adapted from ref. 105 with permission from the PCCP Owner Societies.
The correlation of CO2 sorption capacity per pore volume at 0.15 bar, 25 °C, with the N contents of the carbons or the volume fractions of small micropores (<0.54 nm) in the total pores are also presented (Figure 1.26). Interestingly, the CO2 sorption capacity per pore volume shows a very good linear correlation with the volume percentage of extremely small micropores (<0.54 nm). By contrast, the correlation of CO2 sorption capacity per pore volume at 0.15 bar, 25 °C, with the N contents of the carbons obviously deviates from the linear. This fact further confirmed that the CO2 capture capabilities of the investigated carbons are mainly contributed by the small micropores (<0.54 nm for 0.15 bar at 25 °C) but not the N-doping.
The correlation of CO2 sorption capacity per pore volume at 0.15 bar, 25 °C, with (a) N contents of the carbons and (b) the volume fractions of small micropores (<0.54 nm) in the total pores. Adapted from ref. 105 with permission from the PCCP Owner Societies.
The correlation of CO2 sorption capacity per pore volume at 0.15 bar, 25 °C, with (a) N contents of the carbons and (b) the volume fractions of small micropores (<0.54 nm) in the total pores. Adapted from ref. 105 with permission from the PCCP Owner Societies.
1.6.2 Surface Chemistry Effect on CO2 Adsorption
To enhance the interactions of porous carbons with CO2 molecules, many authors have tried to introduce polarity (e.g., N-doping, S-doping, or cation-introducing, etc.) into the carbon frameworks because the CO2 is highly quadrupolar and weakly acidic. However, the effect of heteroatom-doping, especially N-doping, on CO2 capture is still unclear and is often controversial. Recently, there have been some experimental and simulation studies aimed at clarifying the effects of surface chemistry on the CO2 uptake of carbons. Herein, we briefly summarize the recent relevant progress.
1.6.2.1 Effect of Nitrogen Doping
N-doping is widely used. Many authors have suggested that N-doping enhances the CO2 capture performance of porous carbon adsorbents. In these works, the enhancement effect of N-doping is generally assumed to be chemical sorption based on a one-on-one base–acid interaction. Based on this view, our group prepared N-doped microporous carbons by a two-step casting process using zeolite NaY as a hard template, furfuryl alcohol as a carbon precursor and acetonitrile as the CVD (chemical vapor decomposition) substrate, in which the use of acetonitrile is to introduce nitrogen functionalities to the microporous carbons.239 Based on the analysis results of X-ray photoelectron spectra, it was found that the type of nitrogen group strongly depends on the treatment temperature of the carbon samples (Figure 1.27). The nitrogen groups on the carbons prepared at 500 and 600 °C are mainly Lewis-basic species of amides/pyrrolic (400 eV) and pyridinic (398.9 eV) groups, while most of these basic groups decomposed at a higher temperature (>600 °C) and were converted into aromatic nitrogen species built into the carbon matrix (mainly quaternary N with typical Lewis acidity, ∼401 eV). Although the carbon prepared at 500 °C possessed a very low CO2 uptake (57 mg g−1), this carbon showed the highest heat of adsorption (35–40 kJ mol−1), much higher than the typical heat of physical adsorption (∼20 kJ mol−1), indicating a strong affinity between the surface of this carbon with CO2 molecules. As the carbon temperature increased, the heat of adsorption significantly decreased due to a decrease in the interactions between CO2 and carbon materials resulting from the loss of unstable basic nitrogen groups at high carbonization temperatures. These results demonstrated that the basic N groups (amides/pyrrolic and pyridinic) had a positive effect on CO2 capture via a mechanism of base–acid interactions, especially for the N-doped carbons prepared at a low carbon temperature.239
N1s spectra of the prepared carbons, the type of nitrogen group, reflected by the peaks of the N1s spectra, depends on the treatment temperature of acetonitrile. Reproduced from ref. 239 with permission from the Royal Society of Chemistry.
N1s spectra of the prepared carbons, the type of nitrogen group, reflected by the peaks of the N1s spectra, depends on the treatment temperature of acetonitrile. Reproduced from ref. 239 with permission from the Royal Society of Chemistry.
However, the preparation temperature of most carbon adsorbents is usually higher than 600 °C with the purpose of developing high microporosity. The nitrogen species in these N-doped carbons frequently appear in four weakly alkaline or Lewis-acidic types: pyridinic N, pyrrolic N, quaternary N, and pyridine-N-oxide (Figure 1.28). The one-on-one base–acid interactions could only be related to the lone pair of electrons in pyridinic-N and pyrrolic-N. Further, considering the relatively low N-doping level of the carbon adsorbents reported (<5 at%), the N-doping effect of acid–base interaction may be marginal.
Nitrogen-containing groups appearing frequently in the carbon framework.
In 2012, our group suggested a hydrogen-bonding interaction mechanism on the N-doped activated carbons through quantum chemical calculations.240 A theoretical N-doped carbon surface model (NCSM) containing five different typical N-containing functional groups was developed. For comparison, a pure carbon surface model (CSM) was also devised, as is shown in Figure 1.29. Density functional theory B3LYP was employed to study the interactions between these models and CO2 at 12 different positions, corresponding to 12 different H atoms in the NCSM model, and all the configurations were optimized with the 6-31+G* basis set for all atoms using the Gaussian-03 suite package. The optimized results at the 6-31+G* level showed that there were eight NCSM–CO2 complexes and seven CSM–CO2 complexes. In these complexes, hydrogen bonds between CO2 and NCSM/CSM were formed due to the high electronegativity of the oxygen atom in the CO2 molecule. This type of weak hydrogen bond has been widely studied in recent years, although the interaction of C–H/O was weaker than that of typical hydrogen bonds such as O–H/O and N–H/O. Computational results indicated that binding energies for such hydrogen bonds were different at various positions. The larger the binding energy ΔE (kJ mol−1), the stronger the adsorption affinity. The average binding energy of eight NCSM–CO2 complexes was 7.84 kJ mol−1, and that of CSM–CO2 complexes was only 1.26 kJ mol−1, suggesting that the hydrogen bonds in the NCSM–CO2 complexes were much stronger than those in CSM–CO2 complexes. This demonstrated that the introduction of N into a carbon surface facilitated the hydrogen-bonding interactions between the carbon surface and CO2 molecules. The existence of this weak hydrogen-bonding interaction was further confirmed by the FTIR analyses of SK-0.3-700 carbon (Figure 1.30), which showed that the peak due to a C–H anti-symmetric vibration was broadened and red-shifted to a low wavenumber under a CO2 atmosphere compared with that measured under a N2 atmosphere.
Theoretical models for (a) an N-doped carbon surface and (b) a pure carbon surface (red ball: oxygen atom; blue ball: nitrogen; grey ball: carbon; small grey ball: hydrogen). (c) Hydrogen bond energies at different adsorption sites. Reproduced from ref. 240 with permission from the Royal Society of Chemistry.
Theoretical models for (a) an N-doped carbon surface and (b) a pure carbon surface (red ball: oxygen atom; blue ball: nitrogen; grey ball: carbon; small grey ball: hydrogen). (c) Hydrogen bond energies at different adsorption sites. Reproduced from ref. 240 with permission from the Royal Society of Chemistry.
FT-IR spectra of the activated carbon measured under different atmospheres. Reproduced from ref. 240 with permission from the Royal Society of Chemistry.
FT-IR spectra of the activated carbon measured under different atmospheres. Reproduced from ref. 240 with permission from the Royal Society of Chemistry.
In the previous section, we discussed a temperature-dependent pore size effect on CO2 uptake. In that work, the nitrogen content of the PANI-based carbons was also correlated with CO2 sorption capacities at different sorption temperatures (0–75 °C). What was surprising was that there was no positive correlation between the CO2 sorption capacity and the N content, concluded from the considerably low coefficient of determination (Figure 1.31, R2 ≈ 0.20). According to the D–A theory, the CO2 molecules fill the micropores in a liquid-like state. We evaluated the CO2 sorption capacities at 25 °C by multiplying the density of CO2 at 25 °C by micropore volume for pore sizes smaller than 0.7 nm. These calculated values were very close to the measured CO2 sorption capacities although the investigated carbons possess very different N contents (from 2.8 wt% to 6.7 wt%). Based on the these facts, we suggested an innovative view that the small micropore plays a critical role in high CO2 uptake, and the N-doping doesn't enhance CO2 capture capacities for porous carbons.
Linear fitting of the experimental data. A plot of CO2 sorption capacity per pore volume vs. N-doping level at adsorption temperatures of 0, 25 and 75 °C. Adapted from ref. 105 with permission from the PCCP Owner Societies.
Linear fitting of the experimental data. A plot of CO2 sorption capacity per pore volume vs. N-doping level at adsorption temperatures of 0, 25 and 75 °C. Adapted from ref. 105 with permission from the PCCP Owner Societies.
Sevilla et al. also investigated the role of micropore size and N-doping in CO2 capture, and confirmed our view.235 They prepared two types of activated carbons with analogous pore textures and different surface chemistry (N-free and N-doped), and comprehensively compared the CO2 capture performances of these carbons. The CO2 capture capacities of N-free and N-doped carbons were analogous, and no significant differences could be observed between the isotherms of the N-free and N-doped samples over the whole pressure range at three sorption temperatures: −15, 0, and 25 °C. This observation further showed that the nitrogen functionalities present in these N-doped samples do not influence CO2 adsorption capacities.
Kumar used molecular simulations to study the effects of quaternary N-doping on CO2 uptake and CO2–N2 selectivity in representative carbon pore architectures (slit and disordered carbon structures) at 298 K.241 The simulation results showed that N-doping can only deliver a marginal enhancement on CO2 uptake, but it can improve CO2–N2 selectivity in smaller micropores. Therefore, they suggested that CO2 uptake and CO2–N2 selectivity were predominantly controlled by the pore architecture as well as ultramicropores. They further found that the tendency of linear CO2 molecules to lie flat on the carbon surface favored CO2 uptake in slit pore architectures rather than disordered carbon pore structures. It should be noted that the N-doping effect demonstrated here may be difficult to exemplify experimentally if the material has a disordered pore architecture and complex surface chemistry (such as the presence of other functional groups).241
There are some other studies on the effect of N-doping. For instance, Sánchez-Sánchez et al. suggested that nitrogen functionalities acted as basic sites and oxygen and phosphorus groups as acidic ones toward the adsorption of CO2 molecules, and among the N-containing groups, pyrrolic groups exhibited the highest influence, while the positive effect of pyridinic and quaternary functionalities was much smaller.134 Their results agreed well with our experimental results.239 On the contrary, a DFT simulation by Lim et al. indicated that pyridine groups were most effective at enhancing affinity with CO2 via Lewis base–acid interactions, and pyrrolic groups generated weaker interactions with CO2.242 Babu et al. carried out experimental research on the CO2 capture of mesoporous carbon nanotubes, and suggested that the presence of nitrogen functionalities has a beneficial influence on the CO2 adsorption characteristics over a wide range of pressure (0–36 bar).243 In a word, the influence of nitrogen incorporation on CO2 capture is still a controversial research topic with conflicting conclusions, therefore, we can't draw firm conclusions here.
1.6.2.2 Effect of Other Heteroatom-doping
Besides N-doping, other types of heteroatom-doping, such as S-doping,37,121,244–246 O-doping,134,247–250 and P-doping,251 etc. have also been researched.
There are several reports on the effects of S-doping on CO2 capture. For instance, Xia et al. found that S-doped microporous carbon materials templated by EMC-2 possessed a homogeneous thiophenic S-doping, and exhibited a large isosteric heat of CO2 adsorption. They suggested that the thiophenic groups strongly interact with CO2 molecules via strong base–acid interactions and strong pole–pole interactions between the large quadrupole moment of CO2 and the polar carbon surface associated with S-doping.121 Seema and his co-workers suggested that the oxidized S content played a more significant role in increased CO2 adsorption, while the plot of total S content (the sum of oxidized S and thiophenic S) vs. CO2 uptake showed a negative correlation.37 DFT calculations indicated that the oxidized S groups attracted CO2 molecules more strongly than pyrrolic N groups (Figure 1.32). The attraction energy of oxidized S groups with CO2 was mainly due to the attraction energy between the negative O atom of thiophene (−0.94) and the positive C atom of the CO2 molecule (+1.07). The enhanced effect of oxidized S (sulfone group) on CO2 uptake was also reported for MOF-based CO2 adsorbents.252 Li et al. reported that sulfur, nitrogen, and phosphorus doping were all effective to enhance the gas adsorption capacity of carbon materials; among them, sulfur doping was the most positive.251 However, Bandosz et al. recently reported that the S-containing groups were not stable in a CO2 atmosphere; the thiophenes were oxidized by CO2 molecules to sulfones and sulfonic acids, which were thermodynamically unstable and further decomposed forming SO/SO2 and water, providing additional electrons for CO2 reduction.245,246,253
Molecular interactions of CO2 with (a) the di-oxidized S of thiophene, (b) mono-oxidized S of thiophene, (c) pyrrolic N and (d) pyrrolic hydrogen. The binding energies are 4.3, 6.0, 3.4 and 3.1 kcal mol−1, respectively. The dotted distances are denoted in Angstroms. All calculations were performed using the DFT (M062X/6-31+G*) method. Reprinted from H. Seema, K. C. Kemp, N. H. Le, S.-W. Park, V. Chandra, J. W. Lee and K. S. Kim, Highly selective CO2 capture by S-doped microporous carbon materials, Carbon, 66, 320–326, Copyright 2013, with permission from Elsevier.
Molecular interactions of CO2 with (a) the di-oxidized S of thiophene, (b) mono-oxidized S of thiophene, (c) pyrrolic N and (d) pyrrolic hydrogen. The binding energies are 4.3, 6.0, 3.4 and 3.1 kcal mol−1, respectively. The dotted distances are denoted in Angstroms. All calculations were performed using the DFT (M062X/6-31+G*) method. Reprinted from H. Seema, K. C. Kemp, N. H. Le, S.-W. Park, V. Chandra, J. W. Lee and K. S. Kim, Highly selective CO2 capture by S-doped microporous carbon materials, Carbon, 66, 320–326, Copyright 2013, with permission from Elsevier.
Oxygen-containing functional groups inherently exist in the carbon frameworks. To some extent, O-containing groups, such as carbonyls, alcohols and ethers, can also induce Lewis base–acid interactions with CO2 due to the weak electron-donating effect of the oxygen atom. CO2 adsorption measurements showed that the oxidation process led to an increase in CO2 adsorption capacity for the porous carbons.247,248 Liu and Wilcox carried out a simulation investigation on the effect of oxygen-containing surface functionalities on the CO2 adsorption of microporous carbons.250 Bader charge analysis indicated that the oxygen atom was highly electronegative and thus can serve as a basic adsorption site to attract more CO2 molecules. GCMC simulations predicted that oxygen-containing functional groups enhanced CO2 adsorption in microporous carbon materials in the absence of water vapor. The surface heterogeneity leads to a more condensed packing pattern of CO2 molecules in the micropores, allowing for maximum utilization of the limited pore space (Figure 1.33). They further observed that the induced polarity of the surface functionalization could enhance the selectivity of CO2 over CH4 and N2, especially in a low-pressure regime.249
Comparisons of CO2 adsorbed in functionalized micropores with that in the perfect graphite slit pore: left, side views of adsorbed CO2 in various functionalized graphite slit pores with a pore width of 9.2 Å; right, side views of adsorbed CO2 in various functionalized graphite slit pores with a pore width of 20 Å. Adapted with permission from Y. Liu and J. Wilcox, Environ. Sci. Technol., 2012, 46, 1940, Copyright 2012 American Chemical Society.
Comparisons of CO2 adsorbed in functionalized micropores with that in the perfect graphite slit pore: left, side views of adsorbed CO2 in various functionalized graphite slit pores with a pore width of 9.2 Å; right, side views of adsorbed CO2 in various functionalized graphite slit pores with a pore width of 20 Å. Adapted with permission from Y. Liu and J. Wilcox, Environ. Sci. Technol., 2012, 46, 1940, Copyright 2012 American Chemical Society.
Metal ion-doping was also proved to enhance the CO2 capture performance. For example, the K-doped porous carbons reported by Han et al. exhibited a high CO2 adsorption capacity of 1.62 mmol g−1 at 0.1 bar and 25 °C and a high kinetic CO2–N2 selectivity of 44.254 The calculation suggested that the highly ionic K–O bonds led to a strong polarization and charge separation within the carbon cluster and thus significant improvement of CO2 adsorption. Zhu et al. reported a carbon adsorbent derived from pine-cone biomass, and suggested that the well-dispersed nitrogen and calcium dopants (Ca2+) were positive to CO2 capture.255
Undoubtedly, surface chemistry influences the CO2 capture of carbon-based adsorbents, including CO2 capacity, heat of adsorption, and adsorption selectivity of CO2 over other gases. Many experimental and simulation investigations have proved this conclusion. Based on the results obtained so far, the interactions between the surface doping and CO2 molecules could be generally summarized as Lewis acid–base interactions, hydrogen bonding, and electrostatic interactions. However, the exact role of heteroatom-doping, particularly N-doping, remains controversial and more studies are needed. Furthermore, the effect of heteroatom-doping should be evaluated in a practical (or simulated) condition with flue gas because the polar surface sites will enhance the affinity for the H2O that saturates the flue gas, which will significantly degrade the CO2 capture.
1.7 Summary and Outlook
In this chapter, the recent research progress in the field of carbon-based materials for post-combustion CO2 capture has been reviewed. The carbon-based CO2 adsorbents touched on here are classified into porous carbons, graphene-based carbons, carbon nanotubes, and carbon-based hybrid adsorbents. In the section on porous carbon materials, activated carbons, templated carbons, hierarchical porous carbons, and MOFs (ZIFs, COFs)-based carbons, were respectively discussed. The carbon-based hybrid adsorbents were divided into carbon–organic hybrid materials and carbon–inorganic hybrid materials according to the nature of the incorporated component. Since the CO2 capture of porous carbon adsorbents is primarily determined by the microporosity, and is strongly influenced by the surface chemistry of the carbon, the effects of pore size and surface chemistry on CO2 capture are specifically summarized. Overall, carbon-based materials have been extensively researched, and show attractive prospects for pre-combustion CO2 capture. Nevertheless, this research is in its infancy, and further investigations are needed in many areas.
Improving the CO2 capture performances of carbon adsorbents: post-combustion CO2 capture from flue gases possesses the characteristic of a low partial pressure of CO2 (about 10–15% in flue gases) and selective separation of CO2 and N2. The capacity and selectivity of carbon materials are still low for this application. Tuning the pore size, inducing polarity by heteroatom-doping, and making hybrid composites, are efficient strategies for enhancing the CO2 capture performances of carbons.256 On the subject of selectivity, three calculation methods are commonly used in the literature, including the capacity ratio of CO2–N2, the selectivity based on Henry's law, and the IAST selectivity, resulting in a difficulty of comparison between different carbon adsorbents. Among them, IAST selectivity is recommended because this is the standard in mixture adsorption predictions.257
Relationship between CO2 capture and the carbon's structure: in this chapter, we have reviewed the research development of the effects of pore size and surface chemistry. However, controversies remain in this field. Besides, other structure factors of carbon adsorbents, such as the hierarchical pore structure, the ratio of micro/meso/macro porosity, the length of pores, the size and shape of the carbon particle, etc., also influence CO2 capture (e.g., adsorption diffusion dynamics of CO2), and need to be studied.
Evaluation of CO2 separation under real (or simulated) adsorption conditions: the current works mostly focus on the static adsorption of a single-component gas. However, real CO2 separations are performed as dynamic adsorption processes. Furthermore, flue gas in the field of post-combustion capture is a mixture of mostly N2, CO2, and some moisture. The co-existing gas components can cause a decline in the CO2 adsorption capacity. Therefore, evaluation of CO2 capture should be preferentially conducted on real (or simulated) flue gas under dynamic conditions.
Development of low-cost and sustainable carbon adsorbents: the cost and sustainability of CO2 sorbents is crucial in practical applications due to the huge demand for them. In view of this regard, utilization of renewable precursors (typically various biomass sources), reduction/elimination of the activating agents (e.g., single-ion activation, self-template method), and lowering the energy consumption of carbonization (e.g., low carbonization temperature, microwave heating, etc.), can be considered in the manufacture of low-cost carbon adsorbents.
This work was financially supported by the Natural Science Foundation of China (NSFC21576158, 51302156, 21476264), Natural Science Foundation of Shandong Province (ZR2017JL014) and Distinguished Young Scientist Foundation of Shandong Province (JQ201215).