Open access peer-reviewed chapter

Advances in Porous Adsorbents for CO2 Capture and Storage

Written By

Arindam Modak and Subhra Jana

Submitted: 09 June 2017 Reviewed: 04 September 2017 Published: 20 December 2017

DOI: 10.5772/intechopen.70796

From the Edited Volume

Carbon Dioxide Chemistry, Capture and Oil Recovery

Edited by Iyad Karamé, Janah Shaya and Hassan Srour

Chapter metrics overview

2,032 Chapter Downloads

View Full Metrics

Abstract

The steady increase of anthropogenic CO2 in ambient air, owing to the fossil fuel, power plants, chemical processing and deforestation caused by the usage of land, is a key challenge in the on-going effort to diminish the effect of greenhouse gases on global climate change by developing efficient techniques for CO2 capture. Global warming as a consequence of high CO2 level in the atmosphere is considered as one of the major long lasting problems in the twenty-first century. Concern over these major issues with regard to severe climate change and ocean acidification motivated us to develop the technologies that capture the evolved CO2 from entering into the carbon cycle. Therefore, CO2 capture and storage technology is attracting increasing interest in order to reduce carbon level in the atmosphere which in turn mitigates global climate issues. In this regard, highly efficient adsorbents e.g.; zeolites, alkali metal oxides, activated carbon, porous silica show considerable progress in post combustion CO2 capture. Recently, metal-organic frameworks (MOFs), porous organic polymers (POPs), porous clays, N-doped carbon etc. are explored as versatile and quite elegant way for next-generation CO2 capture. In this chapter, we will discuss the broad prospect of MOFs, POPs, nanoporous clays and porous carbon for CO2 storage and sequestration through utilization of their nanospace chemistry.

Keywords

  • CO2 capture and sequestration
  • metal-organic frameworks (MOFs)
  • porous organic polymers (POPs)
  • nanoporous clays
  • porous carbon

1. Introduction

In the late nineteenth century, Goddard Institute for Space Studies (GISS) found that global temperature is increasing on account of global warming at least 0.8°C in every year since 1951, which is continuing to increase still twenty-first century because of the release of greenhouse gases [1]. Based on such criteria “The Intergovernmental Panel on Climate Change” (IPCC) declares that this excessive CO2 emission is markedly related to global climate change and hence could adversely affect global temperature [2]. Furthermore, studies reveal that compared to the CO2 level of 280 ppm during pre-industrial era, CO2 concentration increases to 401 ppm in 2015, mainly because of human activity and anthropogenic industrial revolution that needs an ultimatum to check the atmospheric CO2 level [3]. Thus, in order to diminish carbon level, CO2 capture and storage/sequestration (CCS) as well as CO2 capture and utilization (CCU) should have been implemented for considering a ultra-low carbon content environment. Regarding CCS processes, physical adsorption, absorption and membrane separation technologies are being involved to store CO2 in adsorbents which later buried in deep underground. As for instances, liquid amines like monoethanolamine and diethanolamine are traditionally being utilized to store CO2 through CCS processes [4], however, despite the wide use of such methods, it suffers from major drawbacks like corrosive nature and volatility of amines, decomposition of the generated salts as well as expensive regeneration cost [5]. Therefore, the requirement of new supports e.g.; porous organic polymers (POPs) [6], metal organic frameworks (MOFs) [7], zeolites [8], zeolitic imidazolate frameworks (ZIF) [9], microporous carbon [10], perovskites, hydrotalcites, clathrate hydrates [11], etc. are mandatory to overcome such drawbacks. It is worth mentioning, porous materials with suitably decorated narrow micropores/ultra narrow micropores are significant in adsorbing large quantity of CO2 through utilization of hollow nanospace, hence, these materials could be promising in frontier research in order to check the adverse effect of CO2 that is continuously releasing as tens of billions of tons in each year. Again, MOFs, POPs etc. having wide diversity in structure engineering, can stabilize various organometallic complexes through utilization of their surface functional sites, which in turn enhances their CO2 adsorption efficacy. Although CCS has wide applications, but only storage in nanopores could partially solve the carbon emission problems because underground leaking cannot be completely ruled out. This circumstance justifies the implementation of suitable process that could simultaneously convert the captured CO2 to value-added products [12]. Nevertheless, being cheap and easy availability, CO2 is enriched as C1 feedstock and has broad scope of serving as C1 chemistry to produce important products e.g. fuels, commodity chemicals, agrochemicals, valuable materials and so on [13].

In this chapter, we describe the fabrication of several important solid materials for controlling environmental remediation of CO2 through CCS techniques and also highlight their potential utility for CO2 adsorption purposes. However, our major aim is to focus the profound application and usage of several solid-phase adsorbents like, MOFs, POPs, nanoporous carbon, porous clays for CO2 capture and sequestration study.

Advertisement

2. Promising materials for CO2 capture

2.1. Metal-organic frameworks (MOF)

MOFs constitute of a new class of materials which could serve as an ideal platform for the development of next-generation CO2 capture materials because of their large capacity for adsorption of gases and easy tailorability of their structures [14]. MOFs are also called coordination polymers because of their propensity to build up by metal atoms (ions) linked together by multifunctional organic ligands. Therefore, MOFs synthesis is a consequence of the linkage between the inorganic and organic chemistry, and it has been considered that MOFs are pioneered by Hagrman et al. [15], Batten and Robson [16], Zhou and Kitagawa [17] and Yaghi et al. [18].

It is customary to say that the framework components of MOFs are required to precisely control for specific type of CO2 capture e.g. post-combustion capture, pre-combustion capture, oxy fuel combustion and potentially even for the specific power plant in which the capture system is to be installed. In this regard, significant efforts have been made recently in improving the gas separation performances of MOFs and some studies are therefore being under consideration for evaluating the prospects of deploying these materials in real-world CO2 capture [19]. This has been predominately invented by Noro et al. [20] and Yaghi et al. [18]. The wide diversity in its structure makes it possible to construct material with high surface area [21], tailor the material properties and thus its affinity towards specific gas molecules [22, 23]. Accordingly, it would be possible to develop MOF-based sorbents with large capacity and high selectivity for CO2 storage purposes.

It can be stated that most of the MOFs are synthesized using non-renewable organic solvents like N,N-dimethylformamide, dioxane, etc., although few MOFs (CD-MOF) can also be made from green solvents like ethanol, water etc. [27]. During synthesis, high surface area MOFs are filled with solvent molecules. Since MOFs are being used for CO2 storage, therefore, the solvent inside the pores are required to remove without any disturbance in its porous structure. However, special care must be taken to assure that the structure does not collapse. It is worthy to mention that the thermal stability of the MOF is expected to be lower than that of zeolites/zeolite analogues because MOFs have lower metal-linker binding energy than that of zeolites having stronger Si-O, Al-O linkages. In Figure 1, some representative MOF structures are shown.

Figure 1.

Crystal structure of several MOFs, e.g.; MOF-5 [24], DA-MOF [25], CO2(adenine)2-(CH3CO2)2 (bio-MOF-11) [26], with permission from respective references.

The CO2 storage capacity of MOFs is being measured at different temperature and pressure. There has also been established a positive connection between storage capacity and surface area at high pressure. Yaghi et al. [18] were first to explain the relationship between surface area and CO2 uptake capacity. Among several MOFs, MOFs with square channels (MOF-2) [28], pores with open metal sites (MOF-505, Cu3(BTC)2) [29], interpenetrated (IRMOF-11) [29], CO2 (adenine)2-(CH3CO2)2 functionalized bio-MOF-11 are interesting [30]. It is important to mention that MOF-177 has the high surface area (4500 m2 g−1) and possesses high uptake of CO2 (60 wt% at 35 bar) at high pressure. Very recently, ultrahigh porous MOFs are prepared by Furukawa et al. from Zn4O(CO2)6 unit, containing one or two organic linkers and among them MOF-210 showed highest BET surface area (6240 m2 g−1) and high pore volume (3.6 cm3 g−1) reported till date [31]. The ultrahigh porosity of MOF-210 has been achieved mainly because of the expanding organic linkers. As for instance, by extending the size of the 1,3,5-benzenetribenzoate (BTC) in MOF-177 by replacing with larger 4,4′,4″-(benzene-1,3,5- triyl-tris(benzene-4,1-iyl)) tribenzoate (BBC), MOF-200 was obtained with even higher BET than that of MOF-177 [32]. Strikingly, MOF-210 and MOF-200 show ~71 wt% CO2 uptake capacity at 298 K and 50 bar pressure, which are highest and considered as new record among all solid porous materials. For CO2 capture at ambient condition as well as flue gas condition, CO2 capture is not only influenced by surface area but is also dependent on adsorbent-CO2 interaction. In this context, Mg-MOF-74 [Mg2(DOT); DOT: 2,5-dioxidoterephthalate] consisting of an open framework with Mg2+ sites shows high CO2 storage capacity of 35.2 wt% at 298 K and 1 bar [32]. It is pertinent to mention that open metal sites are mandatory in achieving high CO2 storage capacity. In Table 1, we present some MOFs having significance in CO2 capture at low and high pressure.

Chemical formulaeBET (m2 g−1)CO2 storage (wt%)Temp. (K)Pressure (bar)References
MOF-17745006029850[31]
MOF-21062407129850[31]
Mg-MOF-74164035.22981[32]
Cu2(abtc)338.52731[36]
Dy(BTC)65527.22731[37]
Al4(OH)2(OCH3)4(BDC-NH2)3126824.12731[38]
Zn2(BTetB)137019.72731[39]
[In3O(diazDBC)1.5(H2O)3](NO3)89217.92731[40]
Co4(OH)2(p-CDC)3108016.42731[41]
Cu(bpy)2(BF4)213.82731[42]
Cu2(bptb)121712.62731[43]
Zn2(BTetB)(DMF)280012.12731[30]
Zn2(BDoborDC)480012.12731[44]
Ni(bpy)2(BF4)211.92731[34]
Ni2(bpy)3(NO3)410.62731[45]
Cd(mim)224205.62731[46]

Table 1.

CO2 adsorption capacity of metal-organic frameworks at low and high pressure.

On the other hand, thermal degradation of MOFs generally happens on account of metal-ligand bond breaking. As a consequence, thermal stability of MOFs is predominately dependent on both the metal-linker bond strength and the number of linkers connected to metal sites. Indeed, MOFs can be stable up to 300°C without degradation of the framework structure [17, 18], which is crucial for post-combustion CO2 capture from flue gas under moist conditions.

Since open metal sites in MOFs show considerably good CO2 uptake property, one common question that may rise among readers is the effect of water on CO2 uptake because water can easily coordinate to open metal sites. It has been found that in some cases small amount of water can accelerate CO2 adsorption in some MOFs. Yazaydin et al. have observed that water molecules coordinated to open metal sites of Cu-BTC (HKUST-1) and thus significantly increase CO2 adsorption of this framework [33]. However, in other cases; water could only destroy the MOF structure and shows a detrimental effect on CO2 adsorption. Kizzie et al. showed the effect of humidity on the CO2 capture performance of M/DOBDC series of MOFs (M = Zn, Ni, Co and Mg; DOBC = 2,5-dioxidobenzene-1,4-dicarboxylate), demonstrating a significant decrease in the CO2 uptake capacities for Mg/DOBC and Zn/DOBC [34]. Recently, Liu et al. also described water stability of various MOFs after their pretreatment, further suggesting their stability in aqueous condition [35].

2.2. Porous organic polymers (POP)

POPs are recently developed as an important class of porous materials, constructed from lightweight elements (C, H, O, N) and linked by strong covalent bonds, which show huge potential in a variety of applications, e.g.; gas storage and separation, catalysis, sensing, energy storage, optoelectronics and to mention a few [8, 47, 48, 49]. A series of vibrant characteristics of POPs, such as high-specific surface area, good physicochemical stability, tunable pore dimensions, topologies and chemical functionalities, make them suitable adsorbents for CO2 capture.

POPs are widely used and have been classified according to their synthesis conditions, which includes polymers of intrinsic microporosity (PIMs), conjugated microporous polymers (CMPs), covalent organic polymers, hypercrosslinked polymers (HCPs), crystalline triazine-based frameworks (CTFs), porous organic frameworks (POFs), porous aromatic frameworks (PAFs), polymeric organic networks (PONs) and so on. In each networks, the state-of-the art development in the design, synthesis, characterization and the CO2 adsorption performances has been reviewed [50, 51]. Additionally, CO2 uptake capacity and adsorption enthalpy can also be controlled through manipulation of POP surface area, pore size and/or its functionality at the surfaces. Understanding all these aspects could lead to the development of new possibilities of novel POP frameworks to target CO2 capture, where the amorphous characteristics, high cross-linked ultra-microporous structures could be crucial in determining CO2 adsorption performances. Herein, we describe the critical factors that directly influence the CO2 uptake property for POPs.

There is a direct correlation between POP surface area and its porous structures. It is said that CO2 uptake capacity is dependent on the high specific surface area. It has also been suggested that CO2 uptake of an adsorbent increases with increasing CO2 pressure, which demonstrates that controlling of surface property might be advantageous in tailoring CO2 uptake phenomenon. Significant development in the synthesis of high surface area POP materials has been done by several researchers, among which the synthesis of porous aromatic framework (PAF-1) developed by Qiu et al. via Yamamoto homocoupling of tetrahedral monomers (tetrakis(4-bromophenyl)methane) is significant [52]. Owing to the in-build generation of diamondoid framework topology by the cross-linking of tetrahedral monomers, it creates open and interconnected pores which give rise to high surface area (BET: 5600 m2 g−1) and good CO2 storage capacity (29.5 mmol g−1, 298 K, 40 bar). Later significant contribution has been developed by modifying with other quadricovalent building centers [53], which forms PPN-3 (with adamantine-system), PPN-4 (replaced tetrahedral carbon by silicon) and PPN-5 (replaced tetrahedral carbon by Germanium), as shown in Figure 2. Importantly, all these POPs possess exceptionally high surface area. It should be noted that PPN-4 having BET surface area of 6461 m2 g−1 considers as highest surface area among all reported POPs till date. The exceptionally high surface area enables PPN-4 to have excellent CO2 storage capacity (48.2 mmol g−1) at 50 bar and 295 K.

Figure 2.

Synthetic routes for high surface area PPN-4 (X: Si), PPN-5 (X: Ge), PAF-1 (X: C) from tetrahedral building blocks.

Therefore, it can be arguably said that the design followed by the synthesis of high surface area POPs from tetrahedral building block is a promising approach for CO2 storage at high pressure. Apart from BET surface area, porous structure also triggers CO2 uptake phenomenon, as controlling the pore size similar to the kinetic diameter of CO2 molecule (Size: 3.3 Å) can significantly enhance the storage capacity. So far many researches have been made significant effort to make POPs from tetrahedral building units as CO2 adsorbents and among them hypercrosslinked (HCPs) amorphous polymers are important. HCPs exhibit unique property in gas storage capacity due to its highly cross-linked network which prevents the interconnected porous frameworks from being collapsed [54]. In Figure 3, the formation of cross-linked HCP is presented.

Figure 3.

Schematic representation of the formation of hypercrosslinked porous polymers.

The unique advantage for the synthesis of HCPs, possesses huge prospect in materials chemistry research because it requires cheap, readily available precursors and FeCl3 as non-toxic catalyst. Therefore, utilizing this procedure, several aromatic polymers can be easily prepared as described by Cooper et al. [55] In presence of 1,1′-binapthol the resulting HCP shows 1015 m2 g−1 BET surface area with high CO2 uptake capacity (3.96 mmol g−1, 1 bar). The high CO2 capture capacity is hardly affected by aqueous environment owing to strong hydrophobicity of HCPs. Another interesting research has been done by Puthiaraj et al., where the knitting polymerization has been utilized for binding tetraphenyl aromatic moieties (Figure 4) [56]. Unlike the conventional Suzuki/Sonogashira coupling reactions, HCP polymerization promotes the cross-linking between tetrahedral aromatic moieties in a facile way. In continuation with this research, silicon (Si) and germanium (Ge) containing nanoscale hyper-cross-linked aromatic polymers (PHAPs) were reported by FeCl3 mediated Friedel − Crafts alkylation between tetraphenylsilane or tetraphenylgermanium as a building block and formaldehyde dimethyl acetal as a cross-linker, which yields high surface area (1137–1059 m2 g−1) stable polymers having high CO2 adsorption capacity (104.3–114.4 mg g−1) with an isosteric heat of adsorption in the range of 26.5–27.3 kJ mol−1 (Figure 4). It is worthy to mention that controlled synthesis of silicon and germanium-based nanoscale POPs is very difficult and shows intense research activities because Si and Ge are crucial in influencing physical properties of these POPs.

Figure 4.

Synthesis of porous hypercrosslinked aromatic polymers (PHAP) containing central hetero atoms Si, Ge other than carbon. CO2 uptake capacity is given at the right; reproduced with permission from Ref. [56].

Although high BET surface area is necessary for influencing CO2 adsorption behavior, however, it is not the prime reason, indeed adsorbate-adsorbent interaction might be quite effective for adsorption at low CO2 pressure and for selective adsorption of CO2 in presence of other gases. In this regard, N, P, O, S etc. atoms when incorporated into POPs, show dramatic change in adsorption capacity as well as in the selectivity; because heteroatoms can improve isosteric heat of adsorption between sorbent and sorbate. To address this issues POPs with different functional sites e.g. pyridine (PON-2) or thiophene (PON-3) have been prepared, which exhibit favorable interaction with CO2 despite of having low surface areas [PON-2 (189 m2 g−1) and PON-3 (44 m2 g−1)] than only phenyl-based PON (BET 1447 m2 g−1), clearly explaining the effect of hetero-atom in POP structure [57]. Based on this proposal, a variety of N-containing POPs with several functional sites e.g. carbazole, triazine, porphyrin, benzimidazole, azo, etc. have been reported by several researchers as discussed in the subsequent section.

Microporous polycarbazole (CPOP-1) through polymerization of carbazole has been prepared by Chen et al. by oxidative polymerization of carbazole with FeCl3 as cheap and non-toxic catalyst [58]. CPOP-1 shows high BET surface area (2220 m2 g−1), narrow pore size (0.62 nm) together with the presence of electron-rich carbazole units and high charge density at N-sites, which facilitate uptake of polarizable CO2 molecule through local dipole-quadrupole interactions. The CO2 uptake of CPOP-1 is as high as 4.8 mmol g−1 at 273 K and 1 bar, along with its high selectivity for CO2/CH4 (33) and CO2/N2 (25) separation is also noticeable on account of the framework composition. Consequently, they have also prepared CPOP-(2-7) through this reaction and among them CPOP-7 exhibits the best uptake capacity for CO2 (3 mmol g−1 at 273 K and 1 bar) [59]. Among other N-containing polymeric adsorbents, covalent triazine frameworks (CTFs) are interesting. CTFs are efficiently applied to CO2 capture purposes and are generally synthesized at high temperature (>400°C) by molten ZnCl2 which usually acts as Lewis acid catalyst as well as solvents. CTFs possess moderate BET surface area (1235 m2 g−1), while showing excellent CO2 uptake capacity (4.2 mmol g−1, 1 atm, 273 K) [60]. Again, several PCTFs were synthesized through the modifications of CTFs using branched arms precursors. Importantly, PCTF-1 with biphenyl, PCTF-2 with terphenyl and PCTF-3 with quaterphenyl rings have been developed, which show BET surface area as 853, 811 and 395 m2 g−1 respectively [61]. When the middle benzene ring of CTF has been replaced by the strong polar group, such as benzothiadiazole, forming PCTF-4 which shows highest CO2 uptake of 4.7 mmol g−1 at 273 K and 1 bar. Later, Zou and others prepared a nitrogen-rich polysulfone/polymer (PSF/SNW-1) membrane by introducing nano-sized SNW-1 particles through intrinsically small micropores into PSF matrix via spin-coating [62]. Gas sorption measurements demonstrated that SNW-1 exhibits high adsorption capacity and good affinity towards CO2 because of its high microporosity and the presence of functional amino groups. For CO2 separation, it demonstrates an excellent separation performance including high separation factor (34 and 40 for CO2/CH4 and CO2/N2) and high CO2 permeability (22.4 barrer). Furthermore, long thermal stability is also noticeable and further satisfies that SNW-1/PSF membrane is stable and robust; exhibiting its potentiality in the practical application of CO2 capture.

Based on the above discussion, it is evident that the effect of N-substituted POPs markedly influence CO2 adsorption and separation processes. Regarding this, porphyrin containing microporous POP shows potential application in this area. Porphyrin-based POPs are generally synthesized from expensive Pd and Ni catalysts by a tedious way [63]. However, Modak and Bhaumik introduced a new method with FeCl3 for the formation of porphyrin polymers through a one pot bottom up chemistry involving the condensation between pyrrole and several aromatic dialdehydes [64, 65]. In this report, authors reported the formation of Fe-POP-1/2/3 having 750–875 m2 g−1 BET surface area, narrow microporosity (0.75–1.1 nm) and excellent adsorbent for CO2 (19 wt%, 273 K, 1 atm; by Fe-POP-1), shown in Figure 5. This research has further been extended by changing aromatic dialdehydes to aromatic trialdehyde using 4,4′,4″-(1,3,5-triazine-2,4,6-triyl)tris(oxy)tribenzaldehyde as tripodal precursor. The advantage of using this organic precursor is enormous, because it introduces both triazine and porphyrin networks in a single TPOP-1 material, which stores high amount of CO2 (6.2 mmol g−1 or 27.3 wt% at 3 bar, 273 K).

Figure 5.

Schematic representation for the preparation of porphyrin-based porous polymers by electrophilic aromatic substitution of pyrrole with several aromatic dialdehydes and trialdehydes.

2.3. Porous clay materials

The key advantages of any porous materials involve flexible surface modification due to the availability of high surface area together with lower regeneration energy compared to the any conventional solvent methods. Among the several porous materials, clay minerals are one of the significant materials since they have unique physical properties because of their layered structure, small grain size and large surface to volume ratio. Apart from the mentioned advantages, clay possesses low cost, high mechanical and chemical stability that has been fascinated to the researchers and withdrawn specific research attention. A variety of clay minerals have been used for CO2 capture, like halloysite, kaolinite, montmorillonite, nanosepiolite, bentonite, etc.

In a recent report, pristine halloysite nanotubes (HNTs) were pretreated with acid for selective removal of alumina and thus, produce mesoporous silica nanotubes (MSiNTs), which were then impregnated with polyethylenimine (PEI) to prepare an emerging nanocomposite MSiNTs/PEI (MP) for CO2 adsorption [66]. Recently, Jana and coworkers demonstrate the development of several HNTs based solid adsorbents being grafted with diverse aminosilanes containing both primary and/or secondary amine sites over the surfaces of inexpensive and easily available HNTs to capture CO2 from the ambient air under room temperature and ambient pressure [67]. They have also explored the effect of relative humidity in terms of moisture on the adsorption kinetics of the adsorbents by carrying out the experiment on the seasonal ambient air [11], exhibiting that relative humidity basically regulates the atmospheric CO2 adsorption and corresponding adsorption kinetics (Figure 6). Jana et al. also studied the kinetics of several stable isotopes of CO2 present in the ambient air [68]. Schaef et al. reported the adsorption of CO2 onto kaolinite surfaces under geologic sequestration conditions and also verified through the density functional theory (DFT) [69]. Wang et al. developed montmorillonite supported PEI composite for CO2 capture [70]; the CO2 sorption efficacy was found to be 2.54 mmol g−1 under dry environmental condition and 3.23 mmol g−1 under moisture added condition. They also verified that the density of CO2 in the clay pores is comparatively stable over a wide range of CO2 pressures at a given temperature and at the excess sorption maximum. CO2 sorption increases with decreasing temperature while the high pressure sorption properties demonstrate weak temperature dependence. Elliot A. Roth and coworkers synthesized amine-containing solid sorbent for CO2 capture through the modification of the surface of montmorillonite nanoclay using aminopropyltrimethoxysilane and polyethylenimine [71]. They found that in pure CO2 and 10% CO2 in nitrogen gas streams the nanoclay loaded with only one of the amines exhibits ~6 wt% capture efficacy at 85°C and atmospheric pressure, whereas it is ~7.5 wt% CO2 capture efficiency when the nanoclay immobilized with both the amines. Irani et al. utilized nanosepiolite as an inorganic-organic CO2 sorbent by immobilizing tetraethylenepentamine (TEPA) onto acid-modified nanosepiolite, having capacity of 3 mmol g−1 for 1 vol% CO2 in N2, along with ∼1 vol% H2O at 60°C [72]. Xiao and coworkers exploited an inexpensive and commercially available bentonite for CO2 capture from flue gas after modified by sulfuric acid followed by immobilization of TEPA [73]. Hence, these low-cost clay based adsorbents introduce a new pathway in the frontier area of CO2 capture and sequestration study and should make the procedure environmental friendly, robust, sustainable and thus a more attractive strategy.

Figure 6.

Schematic presentation of the trapping of CO2 from the seasonal ambient air in clay based solid nanocomposites, adapted from Ref. [11] with permission from The Royal Society of Chemistry.

2.4. Nanoporous carbon composite

Nanoporous carbon derived from various carbon precursors, such as polymer, biomass, coal, petroleum etc. shows enormous significance in small molecule adsorption due to high specific surface area, narrow pore size distribution, low preparation cost, easy-to-design, low energy requirements for regeneration, high stability in air and water and high flexibility in heteroatom doping/surface functionalization [74]. Porous carbons are prepared from physical/chemical activation of carbonaceous materials with CO2, steam, ZnCl2, KOH etc., which are generally employed as activating agents for the formation of microporous carbon [75]. In addition, templating method using zeolites, ZSM-5, several MOFs (MOF-5, ZIF-8, Al-PCP) acts as an effective carbon precursors in making amorphous carbon (Figure 7) [76].

Figure 7.

Schematic presentation of the formation of porous carbon from MOFs, POPs, foam and waste materials.

Additionally, from the perspective of environmental friendliness and preparation cost, waste materials e.g. fly ash, coconut, carpet are quite popular to prepare high performance carbonaceous materials for CO2 adsorption [77, 78, 79]. As for instance, porous carbons are prepared from hazelnut shells via carbonization and KOH activation. The obtained porous carbon exhibits a high specific surface area of ~1900 m2 g−1 and high pore volume (1–0.7) cm3 g−1, which are approximately five times greater than those of carbon without activation [80]. Again, coffee residue has also been used as a sustainable source for the preparation of activated carbon with ZnCl2 [81]. The textural properties of the samples thus prepared significantly depend on the ZnCl2/C ratio. On the other hand, porous carbons are also prepared through a facile one step carbonization from porous polymers. Recently, Modak et al. have prepared microporous carbon (K-COP-M) having extremely high surface area of 2186 m2 g−1 and contains micropore volume of 0.85 cm3 g−1 [82]. K-COP-M has been produced by KOH activation of COP-M polymer in an inert condition and at very high temperature (600–700°C). CO2 adsorption potential of K-COP-M-600/700 have been evaluated, indicating that these frameworks adsorb 160–170 cm3 g−1 (7.6–7.1 mmol g−1) CO2 at 1 atm and 273 K, which is higher than that of non-activated carbon derived from COP-M (i.e. COP-M-600/700 stores 77–83 cm3 g−1 CO2). This signifies the utility of KOH in making effective adsorbents by creating ultra-small micropores in carbons. While heteroatom doping is significantly used in controlling the textural properties of porous carbon, carbonization of amine containing organic precursors substantially shows high N-doping and considers as an effective adsorbent of CO2 with very good selectivity. There are several reports available in this respect [49, 83], among which N-doped composite developed by Kim et al. is interesting, where polyindole-reduced graphene oxide (PIG) hybrid was synthesized and later carbonized at 400–800°C temperature and thus produces N-doped graphene sheets [84]. The N-doped graphene sheets are microporous having 0.6 nm pores with BET surface area of about 936 m2 g−1 and show a maximum CO2 uptake of 3.0 mmol g−1 at 25°C and 1 atm pressure. Nevertheless, high recycling stability of CO2 uptake is noticed even after 10 recycling cycles; additionally this N-doped carbon shows CO2/N2, CO2/CH4 separation ability of 23 and 4 respectively.

Alternatively, polyurethane foams (PUFs) are important thermosetting polymers and owing to its high nitrogen contents, it can be used as good precursor for N-doped carbon [85]. The global demand for polyurethanes was estimated to be 13.6 million tons in 2010, which leads to the generation of huge wastes. However, regeneration of spent polyurethane is not only high energy-consuming process, but also it produces toxic nitrogen oxides, carbon oxides etc. and causing severe environmental pollution. These wastes when carbonized at high temperature can produce nitrogen-doped carbon that can further reduce environmental pollution since these N-doped carbon shows CO2 uptake capacities of 6.67 and 4.33 mmol g−1 at 0°C and 25°C under 1 bar, respectively. Finally, it can be said that like MOFs, POPs and porous clays, microporous carbon is also an alternative for CO2 storage purposes.

Advertisement

3. Conclusion

In this chapter, we highlight important aspect of some promising materials, like MOFs, POPs, nanoporous carbon and porous clays as CO2 adsorbents, which possess high BET surface area, tunable microporosity and facile surface engineering for enhancing interaction with CO2. All such features are however exempted from conventional zeolites, alkali metal oxides, activated carbon, porous silica, therefore, demonstrating the significance of new porous materials in developing carbon capture techniques. Considering the growth of rapid industrialization and abrupt emission of CO2 in the atmosphere, an increasing concern to the social as well as marine lives, would be diminished through utilization of aforesaid functional materials to adsorb CO2. In the abovementioned sections, we have thus tried to summarize the recent advancement made in the synthesis and broad prospect of MOFs, POPs, nanoporous clays and porous carbon as potential adsorbents for CO2 capture and storage. Nevertheless, in the near future, such promising materials would motivate to the researchers to extend this work towards low pressure CO2 adsorption, which may open a new route in the frontier area of carbon capture and sequestering study through the real world CO2 capture under ultra-dilute condition.

Advertisement

Acknowledgments

Authors would like to acknowledge Department of Science and Technology (DST), Government of India for funding (SR/NM/NS-18/2014 and SB/WEA-008/2016) and S. N. Bose National Centre for Basic Sciences, Kolkata, India.

References

  1. 1. Hansen J, Ruedy R, Sato M, Lo K. Global surface temperature change. Reviews of Geophysics. 2010;48:RG4004
  2. 2. GISS Surface Temperature Analysis. Available from: http://data.giss.nasa.gov/gistemp/graphs_v3/ [Accessed: January 2016]
  3. 3. Pachauri RK, Meyer, L. A IPCC. Climate change 2014: Synthesis report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Core Writing Team, editors. IPCC: Geneva, Switzerland, 2014
  4. 4. Reichle D, Houghton J, Kane B, Ekmann J. Carbon Sequestration. Washington, DC: State of the Science; U.S. Department of Energy, Office of Science, Office of Fossil Energy; 1999
  5. 5. Oyenekan BA, Rochelle GT. Alternative stripper configurations for CO2 capture by aqueous amines. AICHE Journal. 2007;53:3144-3154
  6. 6. Wang W, Zhou M, Yuan D. Carbon dioxide capture in amorphous porous organic polymers. Journal of Materials Chemistry A. 2017;5:1334-1347
  7. 7. Sumida K, Rogow DL, Mason JA, McDonald TM, Bloch ED, Herm ZR, Tae-Hyun Bae TH, Long JR. Carbon dioxide capture in metal-organic frameworks. Chemical Reviews. 2012;112:724-781
  8. 8. Stuckert NR, Yang RT. CO2 capture from the atmosphere and simultaneous concentration using zeolites and amine-grafted SBA-15. Environmental Science & Technology. 2011;45:10257-10264
  9. 9. Sajani A, Basnayake JS, Xiadong Z, Kenneth JB. Carbonate-based zeolitic imidazolate framework for highly selective CO2 capture. Inorganic Chemistry. 2015;54:1816-1821
  10. 10. Seema H, Kemp KC, Le NH, Park SW, Chandra V, Lee JW, Kim KS. Highly selective CO2 capture by S-doped microporous carbon materials. Carbon. 2014;66:320-326
  11. 11. Das S, Ghosh C, Jana S. Moisture induced isotopic carbon dioxide trapping from ambient air. Journal of Materials Chemistry A. 2016;4:7632-7640
  12. 12. Duyara MS, ArellanoTreviñob MA, Farrautoa RJ. Dual function materials for CO2 capture and conversion using renewable H2. Applied Catalysis B: Environmental. 2015;168-169:370-376
  13. 13. Gao X, Yu B, Yang Z, Zhao Y, Zhang H, Hao L, Han B, Liu Z. Ionic liquid-catalyzed C-S bond construction using CO2 as a C1 building block under mild conditions: A metal-free route to synthesis of benzothiazoles. ACS Catalysis. 2015;5:6648-6652
  14. 14. Kang Z, Fan L, Sun D. Recent advances and challenges of metal-organic framework membranes for gas separation. Journal of Materials Chemistry A. 2017;5:10073-10091
  15. 15. Hagrman PJ, Hagrman D, Zubieta J. Organic-inorganic hybrid materials: From “simple” coordination polymers to organodiamine-templated molybdenum oxides. Angewandte Chemie, International Edition. 1999;38:2638-2684
  16. 16. Batten SR, Robson R. Interpenetrating nets: Ordered, periodic entanglement. Angewandte Chemie, International Edition. 1998;37:1460-1494
  17. 17. Zhou HC, Kitagawa S. Metal-organic frameworks (MOFs). Chemical Society Reviews. 2014;43:5415-5418
  18. 18. Trickett CA, Helal A, Al-Maythalony BA, Yamani ZH, Cordova KE, Yaghi OM. The chemistry of metal-organic frameworks for CO2 capture, regeneration and conversion. Nature Reviews Materials. 2017;2:17045
  19. 19. Li JR, Ma Y, McCarthy MC, Sculley J, Yu J, Jeong HK, Balbuena PB, Zhou HC. Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks. Coordination Chemistry Reviews. 2011;255:1791-1823
  20. 20. Noro SI, Kitaura R, Kondo M, Kitagawa S, Ishii T, Matsuzaka H, Yamashita M. Framework engineering by anions and porous functionalities of cu(II)/4,4′-bpy coordination polymers. Journal of the American Chemical Society. 2002;124:2568-2583
  21. 21. Jung OS, Kim YJ, Lee YA, Park JK, Chae HK. Smart molecular helical springs as tunable receptors. Journal of the American Chemical Society. 2000;122:9921-9925
  22. 22. Uemura K, Kitagawa S, Fukui K, Saito K. A contrivance for a dynamic porous framework: Cooperative guest adsorption based on square grids connected by amide-amide hydrogen bonds. Journal of the American Chemical Society. 2004;126:3817-3828
  23. 23. Ramsahye NA, Maurin G, Bourrelly S, Llewellyn PL, Devic T, Serre C, Loiseau T, Ferey G. Adsorption of CO2 in metal organic frameworks of different metal centres: Grand canonical Monte Carlo simulations compared to experiments. Adsorption. 2007;13:461-467
  24. 24. Sillar K, Hofmann A, Sauer J. Ab initio study of hydrogen adsorption in MOF-5. J. Am. Chem. Soc. 2009;131:4143-4150
  25. 25. Son HJ, Jin S, Patwardhan S, Wezenberg SJ, Jeong NC, So M, Wilmer CE, Sarjeant AA, Schatz GC, Snurr RQ, Farha OK, Wiederrecht GP, Hupp JT. Light-harvesting and ultrafast energy migration in porphyrin-based metal−organic frameworks. J. Am. Chem. Soc. 2013;135:862−869
  26. 26. An J, Geib SJ, Rosi NL. High and selective CO2 uptake in a cobalt adeninate metal-organic framework exhibiting pyrimidine-and amino-decorated pores. J. Am. Chem. Soc. 2010;132:38-39
  27. 27. Gassensmith JJ, Furukawa H, Smaldone RA, Forgan RS, Botros YY, Yaghi OM, Stoddart JF. Strong and reversible binding of carbon dioxide in a green metal–organic framework. Journal of the American Chemical Society. 2011;133:15312-15315
  28. 28. Millward AR, Yaghi OM. Metal-organic frameworks with exceptionally high capacity for storage of carbon dioxide at room temperature. Journal of the American Chemical Society. 2005;127:17998-17999
  29. 29. Li H, Eddaoudi M, O’Keeffe M, Yaghi OM. Design and synthesis of an exceptionally stable and highly porous metal-organic framework. Nature. 1999;402:276-279
  30. 30. Eddaoudi M, Kim J, Rosi N, Vodak D, Wachter J, O’Keeffe M, Yaghi OM. Systematic design of pore size and functionality in isoreticular MOFs and their application in methane storage. Science. 2002;295:469-472
  31. 31. Furukawa H, Ko N, Go YB, Aratani N, Choi SB, Choi E. Ultrahigh porosity in metal-organic frameworks. Science. 2010;329:424-428
  32. 32. Britt D, Furukawa H, Wang B, Glover TG, Yaghi OM. Highly efficient separation of carbon dioxide by a metal-organic framework replete with open metal sites. Proceedings of the National Academy of Sciences of the United States of America. 2009;106:20637-20640
  33. 33. Yazaydın AO, Benin AI, Faheem SA, Jakubczak P, Low JJ, Willis RR, Snurr RQ. Enhanced CO2 adsorption in metal-organic frameworks via occupation of open-metal sites by coordinated water molecules. Chemistry of Materials. 2009;21:1425-1430
  34. 34. Kizzie AC, Wong-Foy AG, Matzger AJ. Effect of humidity on the performance of microporous coordination polymers as adsorbents for CO2 capture. Langmuir. 2011;27:6368-6373
  35. 35. Liu J, Thallapally PK, McGrail BP, Brown DR. Progress in adsorption-based CO2 capture by metal-organic frameworks. Chemical Society Reviews. 2012;41:2308-2322
  36. 36. Lee YG, Moon HR, Cheon YE, Suh MP. A comparison of the H2 sorption capacities of isostructural metal-organic frameworks with and without accessible metal sites:[{Zn2(abtc)(dmf)2}3] and [{Cu2(abtc)(dmf)2}3] versus [{Cu2(abtc)}3]. Angewandte Chemie, International Edition. 2008;47:7741-7745
  37. 37. Guo X, Zhu G, Li Z, Sun F, Yang Z, Qiu SA. Lanthanide metal-organic framework with high thermal stability and available Lewis-acid metal sites. Chemical Communications. 2006;3172-3174
  38. 38. Si X, Jiao C, Li F, Zhang J, Wang S, Liu S, Li Z, Sun L, Xu F, Gabelica Z, Schick C. High and selective CO2 uptake, H2 storage and methanol sensing on the amine-decorated 12-connected MOF CAU-1. Energy & Environmental Science. 2011;4:4522-4527
  39. 39. Bae Y-S, Farha OK, Hupp JT, Snurr RQ. Enhancement of CO2/N2 selectivity in a metal-organic framework by cavity modification. Journal of Materials Chemistry. 2009;19:2131-2134
  40. 40. Moellmer J, Celer EB, Luebke R, Cairns AJ, Staudt R, Eddaoudi M, Thommes M. Insights on adsorption characterization of metal-organic frameworks: A benchmark study on the novel soc-MOF. Microporous and Mesoporous Materials. 2010;129:345-353
  41. 41. Farha OK, Spokoyny AM, Mulfort KL, Galli S, Hupp JT, Mirkin CA. Gas-sorption properties of cobalt(II)–carborane-based coordinat ion polymers as a function of morphology. Small. 2009;(15):1727-1731
  42. 42. Kondo A, Chinen A, Kajiro H, Nakagawa T, Kato K, Takata M, Hattori Y, Okino F, Ohba T, Kaneko K, Kanoh H. Metal-ion-dependent gas sorptivity of elastic layer-structured MOFs. Chemistry – A European Journal. 2009;15:7549
  43. 43. Wu S, Ma L, Long LS, Zheng LS, Lin W. Three-dimensional metal-organic frameworks based on functionalized tetracarboxylate linkers: Synthesis, strucures, and gas sorption studies. Inorganic Chemistry. 2009;48:2436-2442
  44. 44. Spokoyny AM, Farha OK, Mulfort KL, Hupp JT, Mirkin CA. Porosity tuning of carborane-based metal-organic frameworks (MOFs) via coordination chemistry and ligand design. Inorganica Chimica Acta. 2010;364:266-271
  45. 45. Fletcher AJ, Cussen EJ, Bradshaw D, Rosseinsky MJ, Thomas KM. Adsorption of gases and vapors on nanoporous Ni2(4,4′-bipyridine)3(NO3)4 metal-organic framework materials templated with methanol and ethanol: Structural effects in adsorption kinetics. Journal of the American Chemical Society. 2004;126:9750-9759
  46. 46. Tian YQ, Yao SY, Gu D, Cui KH, Guo DW, Zhang G, Chen ZX, Zhao DY. Cadmium imidazolate frameworks with polymorphism, high thermal stability, and a large surface area. Chemistry – A European Journal. 2010;16:1137-1141
  47. 47. Modak A, Mondal J, Sasidharan M, Bhaumik A. Triazine functionalized ordered mesoporous polymer: A novel solid support for Pd-mediated C–C cross-coupling reactions in water. Green Chemistry. 2011;13:1317-1331
  48. 48. Modak A, Mondal J, Bhaumik A. Highly porous organic polymer containing free –CO2H groups: A convenient carbocatalyst for indole C–H activation at room temperature. ChemCatChem. 2013;5:1749-1753
  49. 49. Modak A, Bhaumik A. High-throughput acid-base tandem organocatalysis over hollow tube-shaped porous polymers and carbons. Chemistry Select. 2016;(6):1192-1200
  50. 50. Xu Y, Jin S, Xu H, Nagai A, Jiang D. Conjugated microporous polymers: Design, synthesis and application. Chemical Society Reviews. 2013;42:8012-8031
  51. 51. Das S, Heasman P, Ben T, Qiu S. Porous organic materials: Strategic design and structure–function correlation. Chemical Reviews. 2017;117:1515-1563
  52. 52. Ben T, Ren TH, Ma S, Cao D, Lan J, Jing X, Wang W, Xu J, Deng F, Simmons JM, Qiu S, Zhu G. Targeted synthesis of a porous aromatic framework with high stability and exceptionally high surface area. Angewandte Chemie, International Edition. 2009;48:9457-9460
  53. 53. Yuan D, Lu W, Zhao D, Zhou HC. Highly stable porous polymer networks with exceptionally high gas-uptake capacities. Advanced Materials. 2011;23:3723-3725
  54. 54. Tan L, Tan B. Hypercrosslinked porous polymer materials: Design, synthesis, and applications. Chemical Society Reviews. 2017;46:3322-3356
  55. 55. Dawson R, Stevens LA, Drage TC, Snape CE, Smith MW, Adams DJ, Cooper AI. Impact of water coadsorption for carbon dioxide capture in microporous polymer sorbents. Journal of the American Chemical Society. 2012;134:10741-10744
  56. 56. Puthiaraj P, Ahn WS. CO2 capture by porous hyper-cross-linked aromatic polymers synthesized using tetrahedral precursors. Industrial and Engineering Chemistry Research. 2016;55:7917-7923
  57. 57. Jeon HJ, Choi JH, Lee Y, Choi KM, Park JH, Kang JK. Highly selective CO2-capturing polymeric organic network structures. Advanced Energy Materials. 2012;2:225-228
  58. 58. Chen Q, Luo M, Hammershoj P, Zhou D, Han Y, Laursen BW, Yan CG, Han BH. Microporous polycarbazole with high specific surface area for gas storage and separation. Journal of the American Chemical Society. 2012;134:6084-6087
  59. 59. Chen Q, Liu DP, Luo M, Feng LJ, Zhao YC, Han BH. Nitrogen-containing microporous conjugated polymers via carbazole-based oxidative coupling polymerization: Preparation, porosity, and gas uptake. Small. 2014;10:308-315
  60. 60. Katekomol P, Roeser J, Bojdys M, Weber J, Thomas A. Covalent triazine frameworks prepared from 1,3,5-tricyanobenzene. Chemistry of Materials. 2013;25:1542-1548
  61. 61. Gu C, Liu D, Huang W, Liu J, Yang R. Synthesis of covalent triazine-based frameworks with high CO2 adsorption and selectivity. Polymer Chemistry. 2015;6:7410-7417
  62. 62. Gao X, Zou X, Ma H, Meng S, Zhu G. Highly selective and permeable porous organic framework membrane for CO2 capture. Advanced Materials. 2014;26:3644-3648
  63. 63. Wang Z, Yuan S, Mason A, Reprogle B, Liu DJ, Yu L. Nanoporous porphyrin polymers for gas storage and separation. Macromolecules. 2012;45:7413-7419
  64. 64. Modak A, Nandi M, Mondal J, Bhaumik A. Porphyrin based porous organic polymers: Novel synthetic strategy and exceptionally high CO2 adsorption capacity. Chemical Communications. 2012;48:248-250
  65. 65. Modak A, Pramanik M, Inagaki S, Bhaumik A. A triazine functionalized porous organic polymer: Excellent CO2 storage material and support for designing Pd nanocatalyst for C–C cross-coupling reactions. Journal of Materials Chemistry A. 2014;2:11642-11650
  66. 66. Mengya N, Huaming Y, Xiangchao Z, Yutang W, Aidong T. Amine-impregnated mesoporous silica nanotube as an emerging nanocomposite for CO2 capture. ACS Applied Materials & Interfaces. 2016;8:17312-17320
  67. 67. Das S, Maity A, Pradhan M, Jana S. Assessing atmospheric CO2 entrapped in clay nanotubes using residual gas analyzer. Analytical Chemistry. 2016;88:2205-2211
  68. 68. Jana S, Das S, Ghosh C, Maity A, Pradhan M. Halloysite nanotubes capturing isotope selective atmospheric CO2. Scientific Reports. 2015;5:8711
  69. 69. Schaef HT, Glezakou VA, Owen AT, Ramprasad S, Martin PF, McGrail BP. Surface condensation of CO2 onto kaolinite. Environmental Science & Technology Letters. 2014;1:142-145
  70. 70. Wang W, Xiao J, Wei X, Ding J, Wang X, Song C. Development of a new clay supported polyethylenimine composite for CO2 capture. Applied Energy. 2014;113:334-341
  71. 71. Roth EA, Agarwal S, Gupta RK. Nanoclay-based solid sorbents for CO2 capture. Energy & Fuels. 2013;27:4129-4136
  72. 72. Irani M, Fan M, Ismail H, Tuwati A, Dutcher B, Russell AG. Modified nanosepiolite as an inexpensive support of tetraethylenepentamine for CO2 sorption. Nano Energy. 2015;11:235-246
  73. 73. Wang W, Wang X, Song C, Wei X, Ding J, Xiao J. Sulfuric acid modified bentonite as the support of tetraethylenepentamine for CO2 capture. Energy & Fuels. 2013;27:1538-1546
  74. 74. Yang Z, Ren J, Zhang Z, Chen X, Guan G, Qiu L, Zhang Y, Peng H. Recent advancement of nanostructured carbon for energy applications. Chemical Reviews. 2015;115:5159-5223
  75. 75. Lozano-Castello D, Lillo-Rodenas MA, Cazorla-Amoros D, Linares-Solano A. Preparation of activated carbons from Spanish anthracite: I. Activation by KOH. Carbon. 2001;39:741-749
  76. 76. Radhakrishnan L, Reboul J, Furukuwa S, Srinivasu P, Kitagawa S, Yamauchi Y. Preparation of microporous carbon fibers through carbonization of al-based porous coordination polymer (Al-PCP) with furfuryl alcohol. Chemistry of Materials. 2011;23:1225-1231
  77. 77. Olivares-Marín M, Sanz-Pérez ES, Wong MS, Maroto-Valer MM. Development of regenerable sorbents from abundant wastes for capture of CO2. Energy Procedia. 2011;4:1118-1124
  78. 78. Liu L, Singh R, Xiao P, Webley P, Zhai Y. Zeolite synthesis from waste fly ash and its application in CO2 capture from flue gas streams. Adsorption. 2011;17:795-800
  79. 79. Ello AS, de Souza LKC, Trokourey A, Jaroniec M. Coconut shell-based microporous carbons for CO2 capture. Microporous and Mesoporous Materials. 2013;180:280-283
  80. 80. Lewicka K. Activated carbons prepared from hazelnut shells, walnut shells and peanut shells for high CO2 adsorption. Polish Journal of Chemical Technology. 2017;19:38-43
  81. 81. Khenniche L, Aissani F. Preparation and characterization of carbons from coffee residue: Adsorption of salicylic acid on the prepared carbons. Journal of Chemical & Engineering Data. 2010;55:728-734
  82. 82. Modak A, Bhaumik A. Porous carbon derived via KOH activation of a hypercrosslinked porous organic polymer for efficient CO2, CH4, H2 adsorptions and high CO2/N2 selectivity. Journal of Solid State Chemistry. 2015;232:157-162
  83. 83. Modak A, Bhaumik A. Surface-exposed Pd nanoparticles supported over nanoporous carbon hollow tubes as an efficient heterogeneous catalyst for the C–C bond formation and hydrogenation reactions. Journal of Molecular Catalysis A: Chemical. 2016;425:147-156
  84. 84. Saleh M, Chandra V, Kemp KC, Kim KS. Synthesis of N-doped microporous carbon via chemical activation of polyindole-modified graphene oxide sheets for selective carbon dioxide adsorption. Nanotechnology. 2013;24:255702
  85. 85. Wang J, Krishna R, Wu X, Sun Y, Deng S. Polyfuran-derived microporous carbons for enhanced adsorption of CO2 and CH4. Langmuir. 2015;31:9845-9852

Written By

Arindam Modak and Subhra Jana

Submitted: 09 June 2017 Reviewed: 04 September 2017 Published: 20 December 2017