Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Interactions of CO2 with various functional molecules

Han Myoung Lee *, Il Seung Youn , Muhammad Saleh , Jung Woo Lee and Kwang S. Kim *
Center for Superfunctional Materials, Department of Chemistry, Ulsan National Institute of Science and Technology (UNIST), Ulsan 689-798, Korea. E-mail: hmlee@unist.ac.kr; kimks@unist.ac.kr

Received 3rd February 2015 , Accepted 17th March 2015

First published on 20th March 2015


Abstract

The CO2 capturing and sequestration are of importance in environmental science. Understanding of the CO2-interactions with various functional molecules including multi-N-containing superbases and heteroaromatic ring systems is essential for designing novel materials to effectively capture the CO2 gas. These interactions are investigated using density functional theory (DFT) with dispersion correction and high level wave function theory (resolution-of-identity (RI) spin-component-scaling (scs) Möller–Plesset second-order perturbation theory (MP2) and coupled cluster with single, double and perturbative triple excitations (CCSD(T))). We found intriguing molecular systems of melamine, 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD), 7-azaindole and guanidine, which show much stronger CO2 interactions than the well-known functional systems such as amines. In particular, melamine could be exploited to design novel materials to capture the CO2 gas, since one CO2 molecule can be coordinated by four melamine molecules, which gives a binding energy (BE) of ∼85 kJ mol−1, much larger than in other cases.


1. Introduction

Carbon dioxide (CO2) is an important green-house gas, which is known to cause serious environmental damage to global weather and human life.1–3 Power plant flue gases contain about 75% N2, 14% CO2, and 10% moisture. Natural gas reserves contain about 40% CO2 and N2 gases.4 Since CO2 gas can be converted into diverse valuable organic molecules, it is highly desirable to develop novel materials that capture CO2 selectively.5,6 Recently, new intriguing methods have been introduced for CO2 capture, storage, and utilization.7,8 Various materials are being developed, such as metal organic frameworks (MOFs),9–12 zeolite-like sorbents,13,14 covalent organic frameworks (COFs),15 polymers with light organic functional groups,16–19 boron nitride nanotubes,20 and many kinds of amines including aminoalcohols,21–23 aqueous ammonia,24 and ionic liquids.25,26 However, simple MOF and zeolite materials show low capacity for separating CO2 from combustion–exhaust gas mixtures.6 The amine-based wet processes of CO2 capture show the degradation problem of amines. The CO2-covalent bonding interactions by amines, ammonia, and ionic liquids consume high energy in regeneration cycles. Furthermore, the aqueous ammonia processes show ammonia loss.27–30 Various functionalized MOF/zeolite materials have been designed for selective CO2 capture,31 and their general functional group is an amine.32–34 Recently a N-containing polymer sphere was reported to show high CO2 adsorption capacity.35 In this material the porous carbon spheres contain intrinsic nitrogen-containing groups. The cooperative CO2-interactions enhance the CO2 adsorption enthalpy with the CO2-interaction energy of a functional group.19,33 To reduce the degradation problem of amines, aromatic molecules can be used to enhance the stability. Substituted aromatic or heteroaromatic systems can have enhanced CO2-BEs as compared with benzene.19,36 On the other hand, CO2 shows some solubility by physisorption in non-polar or weak polar solvents such as benzene, chloroform and dichloromethane.37,38 Since CO2 capture by physisorption shows low CO2-release energy, many CO2 capture materials have been developed based on physisorption. It is thus vital to understand the physisorption strengths of CO2 with many functional groups/molecules. However, limited theoretical investigations were performed.39–44 Moreover, systematic investigation employing reliable high level ab initio methods has hardly been reported for the CO2 interactions with diverse functional molecules. Therefore, we have systematically selected various functional molecules and calculated their BEs with CO2, using reliable high-level computational methods. We have found intriguing functional molecules showing large CO2-interaction energies, which can be used to design novel materials to capture CO2. Since ionic forms must have counter parts or special conditions to be used to capture CO2, ionic forms are excluded in this study.

2. Computational details

The CO2 interactions with various functional molecules were calculated at the M06-2X45 level with the aug-cc-pVDZ basis set (abbreviated as aVDZ) and the resolution-of-identity (RI) spin-component-scaling (scs) Möller–Plesset second-order perturbation theory (MP2) (RI-scs-MP2)46,47 level with the aug-cc-pVTZ (aVTZ) basis set. The geometries were fully optimized without symmetry constraints at each calculation level. The M06-2X functional (hybrid-meta GGA with dispersion correction) has shown good performance in the investigation of the dispersion interaction as well as the electrostatic interaction (H-bonding, H–π interaction, π–π interaction, additional electrostatic and induction energies of neutral and charged dimeric systems).48 Single point (SP) calculations using the RI-coupled cluster theory with single, double and perturbative triple excitations (RI-CCSD(T)) were performed by employing the aVTZ and aug-cc-pVQZ (aVQZ) basis sets at the RI-scs-MP2/aVTZ geometries. The CO2-BEs were calculated at the complete basis set (CBS) limit at the RI-CCSD(T) level with the aVTZ and aVQZ basis sets by employing the extrapolation approximation.49,50 The complete basis set (CBS) energies were estimated with the extrapolation scheme utilizing the electron correlation error proportional to N−3 for the aug-cc-pVNZ basis set (N = 3: T, N = 4: Q). It is generally known that the zero-point-energy (ZPE)-uncorrected BE (−ΔEe) is closer to the experimental CO2-adsorption enthalpy (ΔHads) than the ZPE-corrected BE (−ΔE0).19,51 Therefore, the values of −ΔEe are reported as the CO2-BEs.

Polar σ-bonding functional molecules give significant electrostatic interactions with CO2. Aromatic and heteroaromatic functional molecules give significant dispersion force contributions as well as electrostatic interaction contributions. We analyzed the compositions of BEs using symmetry-adapted perturbation theory (SAPT) at the DFT-PBE0 level with the aVDZ basis set, so-called DFT-SAPT.52 The energy components are the electrostatic energy (Ees), the effective induction energy including the induction-induced exchange energy (Eind* = Eind + Eind-exch), the effective dispersion energy including the dispersion-induced exchange energy (Edisp* = Edisp + Edisp-exch), and the effective exchange repulsion energy with the induction-induced and dispersion-induced exchange energies excluded (Eexch* = Eexch − (Eind-exch + Edisp-exch)).53 In this study, the asymptotically corrected PBE0 (PBE0AC) exchange–correlation (xc) functional with the adiabatic local density approximation (ALDA) xc kernel was used. In the PBE0AC-SAPT calculations, a purely local ALDA xc kernel was used for the hybrid xc functional.

The interaction energies were corrected with the basis set superposition error (BSSE) at the M06-2X and RI-CCSD(T) levels of theory. The RI-scs-MP2 method is known to produce slightly underestimated interaction energies,47 thus, the BSSE corrections were not carried out at the RI-scs-MP2 level. Thermal energies were calculated by employing the M06-2X harmonic vibrational frequencies. The calculations were performed by using the Turbomole package54 and the Molpro package.55

3. Results and discussion

3.1. Understanding of CO2 interactions with various normal functional molecules

The M06-2X CO2-binding structures and energies (in kJ mol−1) of all the functional molecules considered here are given in Fig. 1 and Table 1. The first five structures show simple electrostatic interactions of each polar molecule with CO2. Among them, NH3 having the largest BE (−ΔEe = 14.1 kJ mol−1) with CO2 implies that the sp3 nitrogen atom is the best electron-pair donor to the electron deficient central C atom of CO2. The polar molecules containing the second-row elements have larger BEs with CO2 than those containing third-row elements. The dipole moments of the polar molecules significantly affect their CO2-interactions. The increase of the atomic size or the polarizability in the same group elements has no significant effect on the electrostatic interaction component. Since the fluoric acid (HF) is a good proton donor rather than an electron donor, it shows the H-bond interaction with one electronegative O atom of CO2 (Fig. 1).
image file: c5cp00673b-f1.tif
Fig. 1 M06-2X/aVDZ structures of various functional molecules involved in CO2-interaction.
Table 1 M06-2X/aVDZ CO2-interaction energies (kJ mol−1) with various functional moleculesa
  −ΔEe −ΔE0 −ΔHr
a ΔE0 is the zero-point-energy (ZPE) corrected interaction energy and ΔHr is the enthalpy change at room temperature (298 K). Each interaction energy was corrected by the basis set superposition error (BSSE). Subscripts “-s” and “-i” indicate “stacking” and “in-plane” conformations, respectively.
HF 12.1 7.2 9.0
H2O 13.2 8.0 9.1
NH3 14.1 10.6 10.5
SH2 7.6 3.5 3.7
PH3 5.6 2.9 1.7
HCN 8.6 7.2 6.6
FCH3 10.1 7.4 6.5
OMe2 17.4 15.5 14.2
NMe3 20.3 19.6 18.2
OCH2 9.6 6.8 5.8
NHCH2 17.4 14.5 14.1
MeCN 10.5 8.6 6.9
HCO2H 20.3 16.8 16.4
HCONH2 20.7 17.4 17.1
1,2,3-Triazole 20.0 17.5 16.3
1,2,4-Triazole 20.1 17.1 16.0
CHCl3 11.6 9.5 7.7
CH2Cl2 12.9 10.7 9.3
Guanidine 24.4 21.3 20.3
7-Azaindole 24.4 21.4 20.1
7-A_Tautomer 29.1
TBD 26.9 24.5 23.0
Melamine 27.2 22.5 22.1
Benzene-s 10.6 9.5 7.5
Pyrrole-s 15.7 12.6 11.4
Furan-i 11.6 9.0 7.1
Furan-s 10.1 7.8 6.1
Thiophene-i 6.6 5.4 3.3
Thiophene-s 11.3 10.9 8.7
Pyridine-i 19.3 16.8 15.2
Pyridine-s 11.0 8.5 6.7
Imidazole-i 19.7 16.5 15.2
Imidazole-s 15.1 12.3 10.9
Pyrazine-i 17.0 14.4 12.9
Pyrazine-s 9.0 6.6 4.8
Bisazobipyridine-s 8.5 7.0 5.0
Indole-s 16.7 13.1 11.5


Hydrogen cyanide (HCN) has an sp-hybrid N, which is a relatively poor electron donor and has a smaller BE with CO2 than NH3 does. On the other hand, for the trimethylamine (NMe3)–CO2 binding, the methyl group is electron-donating to the electro-negative N and then enhances the electrostatic interaction strength of the sp3 N as compared with the NH3–CO2 binding (−ΔEe = 20.3 kJ mol−1 for NMe3–CO2; 14.1 kJ mol−1 for NH3–CO2). As N changes from sp3 to sp hybridization, the CO2-BE becomes smaller due to the contraction of the lone pair of electrons (20.3 kJ mol−1 for NMe3–CO2, 17.4 kJ mol−1 for NHCH2–CO2, 8.6 kJ mol−1 for HCN–CO2). The OMe2–CO2 interaction is stronger than the H2O–CO2 interaction, and the OCH2–CO2 interaction has a relatively small BE among the O-containing functional systems. In CO2–OMe2/CO2–NMe3 systems, the simultaneous interactions of the electron deficient central C atom of CO2 with the O/N atom of the functional molecules and the electron rich terminal O atoms of CO2 with the methyl H atoms exhibit so-called cooperative intermolecular interactions, which increase the CO2-BEs. Therefore, the sp3-N containing functional groups (or amine-functionalization) have often been used.26,32–34

The fluoromethane (FCH3) has a considerable CO2-BE (10.1 kJ mol−1) but this is smaller than that of fluoric acid (HF). Nevertheless, some newly designed materials with F-containing functional groups have been introduced to enhance the CO2-adsorption enthalpy.18,56,57 The carbonyl group (C[double bond, length as m-dash]O) of formamide is more polar due to the resonance effect by the amino group (–NH2) than that of formic acid. Thus, formamide has a stronger CO2-BE (−ΔEe = 20.7 kJ mol−1) than formic acid (20.3 kJ mol−1). This explains how MOF materials functionalized by carboxylic acid work well for CO2 capture.58 However, no amide-functionalized material has been investigated for CO2 capture. Amide-based materials could show considerable performance. NMe3 shows strong CO2-BE (20.3 kJ mol−1). The CO2-interaction energies of formic acid and formamide are compatible with that of NMe3. Some interesting research on environmentally friendly amino acids was reported by employing the amino acids as linkers in porous solid materials for CO2 capture in the process of CO2 physisorption.59 The amino acids and aminoalcohols have multiple interaction sites. Neutral amino acids and aminoalcohols can have strong intramolecular H-bonding between their hydroxyl proton and their amine N atom, which can somewhat hinder the CO2 physisorption. 1,2,3- and 1,2,4-triazole molecules also have large CO2-BEs (20.0 and 20.1 kJ mol−1). In the CO2–chloroform (CHCl3) and CO2–dichloromethane (CH2Cl2) interactions, two Cl atoms interact with the central C atom of CO2. The CO2 gas is somewhat soluble in both chloroform and dichloromethane solvents.38

3.2. Special molecules showing strong CO2-interactions

Multi-N containing guanidine, 7-azaindole, 1,5,7-triazabicyclo[4.4.0]dec-5-ene (TBD) and melamine are tautomerizable, showing strong amphoteric properties and having strong CO2-bindings (24.4, 24.4, 26.9 and 27.2 kJ mol−1, respectively). Guanidine and TBD are well known as superbases, and thus it is reasonable that they have large CO2-BEs. The tautomer (7-A_Tautomer) of 7-azaindole shows a very strong CO2-BE (29.1 kJ mol−1). However, the tautomer is 13.4 kcal mol−1 less stable than 7-azaindole at the M06-2X/aVDZ level. Thus, this tautomer cannot be used for practical materials, and so no discussion will be made here. Analogues and derivatives (purine BS3, imidazopyridine, adenine and imidazopyridamine) of 7-azaindole have been reported to have large CO2 BEs.44 At the M06-2X/aVDZ level they show large CO2-BEs of 23.2, 24.3, 26.0 and 25.0 kJ mol−1, respectively. The intriguing point is that the molecules mentioned here show larger BEs than amine species. Their stronger binding with CO2 could imply higher selectivity than amine species. Among the molecules studied here, melamine gives the largest CO2 BE.

3.3. π–π stacking CO2 interactions of aromatic functional molecules

As the intermolecular interactions become important in self-assembly,60,61 the dispersion interactions of π-systems62–64 have recently received much attention. The CO2-binding sometimes shows intriguing competition between electrostatic and dispersion interactions.39–44,65–71 Aromatic systems have extra stability due to the resonance effect, which can reduce the degradation problem of amine cases in the regeneration cycles. Some heteroaromatic systems can have two possible interaction structures with CO2. Subscripts “-i” and “-s” for the aromatic systems of Fig. 1 and Table 1 indicate “electrostatic in-plane” and “dispersive π–π stacking” conformations, respectively. In most cases the in-plane conformations are more stable than the stacking conformations except for the case of thiophene. The CO2-binding of poly-thiophene was applied to CO2 capture and CO2 polymerization.72,73 A poly-pyrrole shows good CO2 adsorption capacity.74 Pyrrole and indole do not have the in-plane conformations with CO2. For the N-containing heteroaromatic systems (imidazole, pyridine and pyrazine), the in-plane conformations (−ΔEe = 19.7, 19.3 and 17.0 kJ mol−1) are much more stable than their stacking conformations with CO2. For the furan, the in-plane conformation (−ΔEe = 11.6 kJ mol−1) is slightly more stable than the stacking conformation (10.1 kJ mol−1) with CO2. The stacking conformation shows somewhat weak binding strength comparable to that of benzene (10.6 kJ mol−1). The attractive dispersion interaction between CO2 and a phenyl ring was experimentally reported.75 The stacking conformations of N-containing pyrazine and bisazobipyridine with CO2 give weak CO2-BEs (9.0 and 8.5 kJ mol−1). Although bisazobipyridine has small CO2-BE, it has been applied to the functionalized MOF material to capture CO2.76 Heteroaromatic systems of pyrrole, thiophene, imidazole and indole show strong stacking interactions with CO2. Among them, indole has the largest CO2-BE (16.7 kJ mol−1). Such heteroaromatic systems were reported as functional materials for CO2 capture.19,72–74,77–80 The stacking conformations of aromatic systems with CO2 include the electrostatic interaction between the electronegative aromatic ring and electropositive central carbon atom of CO2 and the bent H-bond interaction between one electropositive aromatic H atom and one electronegative O atom of CO2, as well as the dispersion interaction.

We performed SAPT calculations at the PBE0/aVDZ level for the in-plane and stacking complexes of CO2 with benzene, pyrrole, thiophene, pyridine and indole. Their interaction energy decompositions are analyzed in Table 2. Their electrostatic interaction energy terms (Ees) are over-compensated by the counter repulsive exchange interaction energy terms (Eexch*). In the case of benzene–CO2 interaction, the dispersion interaction component is much larger than the electrostatic interaction component. For thiophene the stacking conformation has a large dispersion energy (−9.8 kJ mol−1) in comparison with the in-plane conformation in the SAPT calculations. However, for pyridine, the electrostatic component is much larger in the in-plane conformation, and so the in-plane conformation is much more stable than the stacking conformation. The stacked thiophene–CO2 structure has enhanced electrostatic and dispersion interaction energy components in comparison with benzene–CO2. Pyrrole and indole have enhanced electrostatic interaction energy components and enhanced repulsive exchange interaction energy components in comparison with benzene, while they have enhanced attractive dispersion interaction energy components. Indole–CO2 interaction has the largest dispersion interaction component among them.

Table 2 DFT–SAPT interaction energy decompositions (kJ mol−1) of the CO2-interactions with functional molecules for the stacking (-s) (in-plane (-i)) conformations (Etot – total interaction energy; Ees –electrostatic interaction energy; Eexch* – exchange energy term; Eind* – induction energy term; Edisp* – dispersion interaction energy)
  Benzene-s Pyrrole-s Thiophene-s(-i) Pyridine-s(-i) Indole-s
E tot −7.98 −11.65 −8.30 (−5.68) −6.31 (−15.74) −12.47
E es −5.23 −10.77 −6.49 (−4.51) −3.52 (−30.14) −9.53
E exch* 7.12 12.40 9.12 (6.39) 5.64 (33.40) 11.53
E ind* −0.61 −1.12 −0.74 (−0.50) −0.46 (−3.65) −0.99
E disp* −8.97 −11.57 −9.80 (−6.82) −7.77 (−13.90) −12.95


All the results of the BEs of ring compounds studied here are compatible with amine species, e.g. ammonia and NMe3. Moreover, unlike the amine-based wet processes, which exhibit covalent bond breaking, followed by amine loss, the aromatic ring compounds are not expected to have such chemical reactions due to their non-covalent stacking interactions, and thus show no significant amine loss. Therefore, they can be applied to develop novel materials to capture the CO2 gas.

3.4. High level calculations of strong CO2-interaction systems

Based on the M06-2X BEs of the functional molecules with a CO2 molecule, we selected important complexes and calculated their optimal structures and interaction energies at the RI-scs-MP2/aVTZ level. The RI-CCSD(T)/CBS BEs were estimated by the RI-CCSD(T)/aVTZ and RI-CCSD(T)/aVQZ single point calculations at the RI-scs-MP2/aVTZ geometries. These CO2-BEs (−ΔEe) are given in Table 3. Their RI-scs-MP2/aVTZ structures are shown in Fig. 2. Based on the RI-CCSD(T)/CBS BEs, NMe3 has a large CO2-BE (16.0 kJ mol−1). The CO2-BEs of formamide and formic acid are 18.1 and 17.3 kJ mol−1, respectively, which are larger than that of NMe3. The tautomerizable multi-N-containing systems (guanidine, 7-azaindole, TBD and melamine) show much larger CO2-BEs (23.0, 24.3, 25.9, and 26.4 kJ mol−1, respectively). Amine, carboxylic acid and amide have considerably large CO2-BEs. Pyridine and imidazole have larger CO2-BEs (16.9 and 17.3 kJ mol−1) than NMe3 due to larger dipole moments (2.24/2.33 Debye for pyridine, 3.86/3.85 Debye for imidazole, and 0.62/0.79 Debye for NMe3 at the M06-2X/RI-CCSD(T) level). This effect also appears in the 7-azaindole–CO2 interaction. The CO2-BE of indole is 15.5 kJ mol−1. A polymer synthesized with indole shows a larger CO2 adsorption enthalpy (49.0 kJ mol−1) at zero coverage, which is much more than three times the CO2–indole BE due to the cooperative interactions.19 This indicates the binding of a CO2 molecule mostly with three indole molecules and sometimes with four indole molecules. The BE of a CO2 molecule with one indole molecule is 15.5 kJ mol−1, and so the CO2-BEs with three and four indole molecules can be roughly estimated to be 41.5 and 62 kJ mol−1, respectively, when the binding is assumed not to be seriously disturbed by the presence other indole molecules. This could be possible because the CO2–indole interaction is based on the stacking interaction, and the three or four fold interactions with one CO2 molecule the same could be feasible when the stackings are made in the shape of three or four propeller blades of indole surrounding the linear CO2 molecular axis (see, for example, Fig. 3). In reality, the side H-bond interaction of the CO2–indole system would be no more than that of the CO2–water system in which the OC[double bond, length as m-dash]O⋯H–OH interaction energy was 5.5 kJ mol−1.40 Indole is an extended derivative of pyrrole, and carbazole is a larger extended derivative of indole. Among them, indole shows the strongest stacking interaction with CO2.19 The bigger aromatic systems do not show larger CO2-BEs.
Table 3 RI-scs-MP2/aVTZ, RI-CCSD(T)/aVTZ and RI-CCSD(T)/CBS CO2-BEs (−ΔEe in kJ mol−1) on the RI-scs-MP2/aVTZ optimized geometriesa
  scs-MP2 CCSD(T)/aVTZ CCSD(T)/CBS
a The values in parentheses are obtained with the CBS estimates obtained from the MP2 aVTZ and aVQZ energies and the CCSD(T) aVTZ energies.
H2O 10.5 11.3 11.3
NH3 10.7 11.9 11.8
OMe2 15.5 15.9 15.5
NMe3 17.6 17.4 16.0
NHCH2 14.4 15.1 14.9
HCO2H 17.0 17.8 17.3
HCONH2 17.5 18.5 18.1
Guanidine 20.6 21.4 23.0
7-Azaindole 23.1 23.0 (24.3)
TBD 24.1 24.3 (25.9)
Melamine 24.0 24.1 (26.4)
Benzene-s 11.8 9.4 (10.3)
Pyrrole-s 14.3 13.2 12.1
Furan-i 11.5 11.7 10.8
Furan-s 10.1 9.2 8.1
Thiophene-i 7.3 6.8 5.2
Thiophene-s 11.5 9.9 9.9
Pyridine-i 17.3 17.6 16.9
Pyridine-s 11.7 10.8 9.8
Imidazole-i 17.5 16.4 17.3
Imidazole-s 14.0 13.1 12.1
Indole-s 17.5 14.6 (15.5)



image file: c5cp00673b-f2.tif
Fig. 2 RI-scs-MP2/aVTZ structures of selected functional molecules involved in the CO2-interaction.

image file: c5cp00673b-f3.tif
Fig. 3 Designed systems for CO2 capture. The values are BEs (−ΔEe) in kJ mol−1 at the M06-2X/aVDZ level.

3.5. Applications of functional molecules to CO2 capture

In CO2 capture, not only does CO2-BE affect the selectivity but also the molecular weight of the absorbent affecting the weight capacity is an important factor. NMe3, formic acid, formamide, guanidine, 7-azaindole, TBD, and melamine show large CO2-BEs and their molecular weights are 59, 46, 45, 59, 118, 139, and 126 g mol−1, respectively. However, melamine has three CO2-binding sites and then the weight of melamine per CO2 is 42 g mol−1, which shows an extremely impressive weight capacity. Formic acid and formamide have small molecular weights and large CO2-BEs. Formamide has a larger CO2-BE than formic acid. The CO2 interacting systems of amides were experimentally studied in the vapor–liquid equilibrium state.81 However, formamide is a liquid in the standard state. Since the amide–amide interaction is very strong (−ΔEe = 60.2 kJ mol−1 at the M06-2X level), CO2 cannot be dissolved in the formamide solvent.

As exemplary host systems using imidazole, formamide and imine, we can designed imidazole-4-amide and imidazole-2-imine, as shown in Fig. 3. The CO2-BEs of imidazole-4-amide and imidazole-2-imine (−ΔEe = 22.4 and 23.6 kJ mol−1, respectively), are larger than those of imidazole, formamide and imine (19.7, 20.7, and 18.2 kJ mol−1) and that of imidazole-2-carboxylic acid (21.5 kJ mol−1). Thus, the imidazole-4-amide and imidazole-2-imine molecules show impressive CO2 BEs. Indeed, an analogue of imidazole-4-amide was already synthesized and reported as a polymer.82 Dacarbazine as a derivative of imidazole-4-amide is a well-known chemical. The imidazole-2-imine moiety is also found in many imidazole derivatives.

Indole was also successfully used in the polymer form to capture CO2 gas.19 In our study, guanidine, 7-azaindole, TBD and melamine show large CO2-BEs. Among them, 7-azaindole has a similar structure to indole. It is not easy to synthesize the 7-azaindole functional group into polymers due to the difficult oxidation reaction of 7-azaindole. However, if 7-azaindole is used as a functional unit of the materials for CO2 capture, such materials could show high selectivity for CO2 capture due to the large CO2-BE of 7-azaindole (−ΔEe = 24.4 kJ mol−1 at the M06-2X/aVDZ level). The CO2-BE of indole is 17.5 kJ mol−1 at the M06-2X level. As mentioned in the previous section, CO2 can interact with up to four 7-azaindole molecules (Fig. 3), as the experiment showed the CO2 adsorption enthalpy of −49 kJ mol−1 at zero coverage of indole,19 which indicates the interactions with 3 to 4 indole molecules. In this tetra-coordination, the CO2-BE is calculated to be 53.8 kJ mol−1 (−ΔEe) at the M06-2X/aVDZ level. The large adsorption enthalpy is critical to the high capacity and selectivity for CO2 in the gas mixture.

A guanidine-functional polymer was reported to show good performance at high temperature.83 Several applications of melamine were reported to show high capacity for CO2 capture.84,85 However, in most cases the central triazine ring was used as one or two stacking CO2-binding sites which have three N atoms,86 which resulted in relatively weak CO2-BE. Melamine-terminal materials could show better performance due to the effective electrostatic CO2-binding as shown in Fig. 1 and 2. As shown in a model system (Fig. 3), CO2-two melamines, CO2-three melamines and CO2-four melamines show large CO2-BEs (47.9, 69.5 and 85.0 kJ mol−1, respectively) at the M06-2X/aVDZ level. Such large values arise from the maximized electrostatic interactions between CO2 and melamine molecules associated with four-fold triple bindings comprised of two Oδ−⋯H+Nδ− electrostatic H-bonds and one Cδ+⋯Nδ− electrostatic bond. These model systems show much higher CO2-BEs. The multi-N-containing molecules (guanidine, 7-azaindole, TBD, and melamine) with large CO2-BEs could be used as multi-binding sites for a CO2 molecule in devising absorbent materials with large CO2 adsorption enthalpies.

4. Concluding remarks

Tautomerizable multi-N-containing strong bases (guanidine, 7-azaindole, TBD, and melamine) show considerably strong electrostatic interactions with CO2 due to their strong amphoteric properties. Among them, melamine shows the largest CO2-electrostatic BE. The stronger binding between these functional molecules with CO2 could imply better selectivity than the amine species. Among the various aromatic systems considered, indole shows the largest dispersion interaction energy with CO2. The chemical units with large CO2-BEs could be applied to devising functional materials for efficient CO2 capture. Furthermore, CO2 by tetra coordination of melamines gives a very large CO2-BE (85.0 kJ mol−1). Thus, multi-N-containing molecules (guanidine, 7-azaindole, TBD, and melamine) with large CO2-BEs could be used as multi-binding sites for a CO2 molecule. The present results could provide useful information for the development of promising functionalized materials for CO2 capture/sequestration.

Acknowledgements

This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education, Science and Technology (2011-0011222) and NRF (National Honor Scientist Program: 2010-0020414). It was also supported by the Research Fund of UNIST (project no. 1.140001.01). It was also supported by KISTI (KSC-2014-C3-020) and KISTI (KSC-2014-C3-030).

References

  1. A. A. Lacis, G. A. Schmidt, D. Rind and R. A. Ruedy, Science, 2010, 330, 356–359 CrossRef CAS PubMed.
  2. J. D. Shakun, P. U. Clark, F. He, S. A. Marcott, A. C. Mix, Z. Liu, B. Otto-Bliesner, A. Schmittner and E. Bard, Nature, 2012, 484, 49–54 CrossRef CAS PubMed.
  3. M. Garner, A. Attwood, D. S. Baldwin, A. James and M. R. Munafo, Neuropsychopharmacology, 2011, 36, 1557–1562 CrossRef CAS PubMed.
  4. D. M. D'Alessandro, B. Smit and J. R. Long, Angew. Chem., Int. Ed., 2010, 49, 6058–6082 CrossRef PubMed.
  5. C. Lastoskie, Science, 2010, 330, 595–596 CrossRef CAS PubMed.
  6. H. Li, P. H. Opgenorth, D. G. Wernick, S. Rogers, T.-Y. Wu, W. Higashide, P. Malati, Y.-X. Huo, K. M. Cho and J. C. Liao, Science, 2012, 335, 1596 CrossRef CAS PubMed.
  7. P. Markewitz, W. Kuckshinrichs, W. Leiter, J. Linssen, P. Zapp, R. Bongartz, A. Schreiber and T. E. Muller, Energy Environ. Sci., 2012, 5, 7281–7305 CAS.
  8. C.-H. Huang and C.-S. Tan, Aerosol Air Qual. Res., 2014, 14, 480–499 CAS.
  9. P. Nugent, Y. Belmabkhout, S. D. Burd, A. J. Cairns, R. Luebke, K. Forrest, T. Pham, S. Ma, B. Space, L. Wojtas, M. Eddaoudi and M. J. Zaworotko, Nature, 2013, 495, 80–84 CrossRef CAS PubMed.
  10. D.-X. Xue, A. J. Cairns, Y. Belmabkhout, L. Wojtas, Y. Liu, M. H. Alkordi and M. Eddaoudi, J. Am. Chem. Soc., 2013, 135, 7660–7667 CrossRef CAS PubMed.
  11. R. Poloni, B. Smit and J. B. Neaton, J. Am. Chem. Soc., 2012, 134, 6714–6719 CrossRef CAS PubMed.
  12. L. Du, Z. Lu, K. Zheng, J. Wang, X. Zheng, Y. Pan, X. You and J. Bai, J. Am. Chem. Soc., 2013, 135, 562–565 CrossRef CAS PubMed.
  13. H. Amrouche, S. Aguado, J. Perez-Pellitero, C. Chizallet, F. Siperstein, D. Farrusseng, N. Bats and C. Nieto-Draghi, J. Phys. Chem. C, 2011, 115, 16425–16432 CAS.
  14. J. A. Thompson, N. A. Brunelli, R. P. Lively, J. R. Johnson, C. W. Jones and S. Nair, J. Phys. Chem. C, 2013, 117, 8198–8207 CAS.
  15. R. Babarao, R. Custelcean, B. P. Hay and D. Jiang, Cryst. Growth Des., 2012, 12, 5349–5356 CAS.
  16. F. Rezaei, R. P. Lively, Y. Labreche, G. Chen, Y. Fan, W. J. Koros and C. W. Jones, ACS Appl. Mater. Interfaces, 2013, 5, 3921–3931 CAS.
  17. Y. Kuwahara, D.-Y. Kang, J. R. Copeland, N. A. Brunelli, S. A. Didas, P. Bollini, C. Sievers, T. Kamegawa, H. Yamashita and C. W. Jones, J. Am. Chem. Soc., 2012, 134, 10757–10760 CrossRef CAS PubMed.
  18. E. Girard, T. Tassaing, S. Camy, J.-S. Condoret, J.-D. Marty and M. Destarac, J. Am. Chem. Soc., 2012, 134, 11920–11923 CrossRef CAS PubMed.
  19. M. Saleh, H. M. Lee, K. C. Kemp and K. S. Kim, ACS Appl. Mater. Interfaces, 2014, 6, 7325–7333 CAS.
  20. H. Choi, Y. C. Park, Y.-H. Kim and Y. S. Lee, J. Am. Chem. Soc., 2011, 133, 2084–2087 CrossRef CAS PubMed.
  21. F. A. Chowdhury, H. Yamada, T. Higashii, K. Goto and M. Onoda, Ind. Eng. Chem. Res., 2013, 52, 8323–8331 CrossRef CAS.
  22. Y. Matsuzaki, H. Yamada, F. A. Chowdhury, T. Higashii and M. Onoda, J. Phys. Chem. A, 2013, 117, 9274–9281 CrossRef CAS PubMed.
  23. B. Han, C. Zhou, J. Wu, D. J. Tempel and H. Cheng, J. Phys. Chem. Lett., 2011, 2, 522–526 CrossRef CAS.
  24. D. Y. Kim, H. M. Lee, S. K. Min, Y. Cho, I.-C. Hwang, K. Han, J. Y. Kim and K. S. Kim, J. Phys. Chem. Lett., 2011, 2, 689–694 CrossRef CAS.
  25. X. Luo, Y. Guo, F. Ding, H. Zhao, G. Cui, H. Li and C. Wang, Angew. Chem., Int. Ed., 2014, 53, 7053–7057 CrossRef CAS PubMed.
  26. I. Niedermaier, M. Bahlmann, C. Papp, C. Kolbeck, W. Wei, S. K. Calderon, M. Grabau, P. S. Schulz, P. Wasserscheid, H.-P. Steinruck and F. Maier, J. Am. Chem. Soc., 2014, 136, 436–441 CrossRef CAS PubMed.
  27. E. F. da Silva, H. Lepaumier, A. Grimstvedt, S. J. Vevelstad, A. Einbu, K. Vernstad, H. F. Svendsen and K. Zahlsen, Ind. Eng. Chem. Res., 2012, 51, 13329–13338 CrossRef CAS.
  28. C. S. Srikanth and S. S. C. Chuang, J. Phys. Chem. C, 2013, 117, 9196–9205 CAS.
  29. X. Ge, S. L. Shaw and Q. Zhang, Environ. Sci. Technol., 2014, 48, 5066–5075 CrossRef CAS PubMed.
  30. Y. S. Yu, H. F. Lu, T. T. Zhang, Z. X. Zhang, G. X. Wang and V. Rudolph, Ind. Eng. Chem. Res., 2013, 52, 12622–12634 CrossRef CAS.
  31. N. Gargiulo, F. Pepe and D. Caputo, J. Nanosci. Nanotechnol., 2014, 14, 1811–1822 CrossRef CAS PubMed.
  32. S. Choi, T. Watanabe, T.-H. Bae, D. S. Sholl and C. W. Jones, J. Phys. Chem. Lett., 2012, 3, 1136–1141 CrossRef CAS.
  33. R. Vaidhyanathan, S. S. Iremonger, G. K. H. Shimizu, P. G. Boyd, S. Alavi and T. K. Woo, Angew. Chem., Int. Ed., 2012, 51, 1826–1829 CrossRef CAS PubMed.
  34. R. Haldar, S. K. Reddy, V. M. Suresh, S. Mohapatra, S. Balasubramanian and T. K. Maji, Chem. – Eur. J., 2014, 20, 4347–4356 CrossRef CAS PubMed.
  35. S. Wang, W.-C. Li, L. Zhang, Z.-Y. Jin and A.-H. Lu, J. Mater. Chem. A, 2014, 2, 4406–4412 Search PubMed.
  36. S. Biswas, D. E. P. Vanpoucke, T. Verstraelen, M. Vandichel, S. Couck, K. Leus, Y.-Y. Liu, M. Waroquier, V. V. Speybroeck, J. F. M. Denayer and P. V. D. Voort, J. Phys. Chem. C, 2013, 117, 22784–22796 CAS.
  37. E. N. Lay, V. Taghikhani and C. Ghotbi, J. Chem. Eng. Data, 2006, 51, 2197–2200 CrossRef CAS.
  38. K. Shirono, T. Morimatsu and F. Takemura, J. Chem. Eng. Data, 2008, 53, 1867–1871 CrossRef CAS.
  39. K. M. de Lange and J. R. Lane, J. Chem. Phys., 2011, 135, 064304 CrossRef PubMed.
  40. J. Makarewicz, J. Chem. Phys., 2010, 132, 234305 CrossRef PubMed.
  41. A. Torrisi, C. Mellot-Draznieks and R. G. Bell, J. Chem. Phys., 2009, 130, 194703 CrossRef PubMed.
  42. A. Torrisi, C. Mellot-Draznieks and R. G. Bell, J. Chem. Phys., 2010, 132, 044705 CrossRef PubMed.
  43. L. Chen, F. Cao and H. Sun, Int. J. Quantum Chem., 2013, 113, 2261–2266 CrossRef CAS.
  44. K. D. Vogiatzis, A. Mavrandonakis, W. Klopper and G. E. Froudakis, ChemPhysChem, 2009, 10, 374–383 CrossRef CAS PubMed.
  45. Y. Zhao and D. G. Truhlar, J. Phys. Chem. A, 2006, 110, 5121–5129 CrossRef CAS PubMed.
  46. M. Gerenkamp and S. Grimme, Chem. Phys. Lett., 2004, 392, 229–235 CrossRef CAS PubMed.
  47. S. Grimme, J. Chem. Phys., 2003, 118, 9095–9102 CrossRef CAS PubMed.
  48. M. Kolaski, A. Kumar, N. J. Singh and K. S. Kim, Phys. Chem. Chem. Phys., 2011, 13, 991–1001 RSC.
  49. T. Helgaker, W. Klopper, H. Koch and J. Noga, J. Chem. Phys., 1997, 106, 9639–9646 CrossRef CAS PubMed.
  50. S. K. Min, E. C. Lee, H. M. Lee, D. Y. Kim, D. Kim and K. S. Kim, J. Comput. Chem., 2008, 29, 1208–1221 CrossRef CAS PubMed.
  51. R. Vaidhyanathan, S. S. Iremonger, G. K. H. Shimizu, P. G. Boyd, S. Alavi and T. K. Woo, Science, 2010, 330, 650–653 CrossRef CAS PubMed.
  52. A. Heßelmann, G. Jansen and M. Schütz, J. Chem. Phys., 2005, 122, 014103 CrossRef PubMed.
  53. E. C. Lee, D. Kim, P. Jurečka, P. Tarakeshwar, P. Hobza and K. S. Kim, J. Phys. Chem. A, 2007, 111, 3446–3457 CrossRef CAS PubMed.
  54. TURBOMOLE V6.4 2012; a development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, Karlsruhe, Germany, 2007, available from http://www.turbomole.com.
  55. H.-J. Werner, P. J. Knowles, R. Lindh, F. R. Manby, M. Schutz, P. Celani, T. Korona, G. Rauhut, R. D. Amos, A. Bernhardsson, A. Berning, D. L. Cooper, M. J. O. Deegan, A. J. Dobbyn, F. Eckert, C. Hampel, G. Hetzer, A. W. Lloyd, S. J. McNicholas, W. Meyer, M. E. Mura, A. Nicklass, P. Palmieri, R. Pitzer, U. Schumann, H. Stoll, A. J. Stone, R. Tarroni and T. Thorsteinsson, MOLPRO, a package of ab initio programs, version 2006.1, Institut für Theoretische Chemie, Universität Stuttgart, Stuttgart, Germany, 2006 Search PubMed.
  56. D.-S. Zhang, Z. Chang, Y.-F. Li, Z.-Y. Jiang, Z.-H. Xuan, Y.-H. Zhang, J.-R. Li, Q. Chen, T.-L. Hu and X.-H. Bu, Sci. Rep., 2013, 3, 3312 Search PubMed.
  57. J. J. Nogueira, S. A. Vazquez, U. Lourderaj, W. L. Hase and E. Martinez-Nunez, J. Phys. Chem. C, 2010, 114, 18455–18464 CAS.
  58. Z. Xiang, S. Leng and D. Cao, J. Phys. Chem. C, 2012, 116, 10573–10579 CAS.
  59. M. A. Hussain, Y. Soujanya and G. N. Sastry, Environ. Sci. Technol., 2011, 45, 8582–8588 CrossRef CAS PubMed.
  60. J. Y. Lee, B. H. Hong, D. Y. Kim, D. R. Mason, J. W. Lee, Y. Chun and K. S. Kim, Chem. – Eur. J., 2013, 19, 9118–9122 CrossRef CAS PubMed.
  61. J. Y. Lee, B. H. Hong, W. Y. Kim, S. K. Min, Y. Kim, M. V. Jouravlev, R. Bose, K. S. Kim, I.-C. Hwang, L. J. Kaufman, C. W. Wong, P. Kim and K. S. Kim, Nature, 2009, 460, 498–501 CrossRef CAS.
  62. Y. Cho, W. J. Cho, I. S. Youn, G. Lee, N. J. Singh and K. S. Kim, Acc. Chem. Res., 2014, 47, 3321–3330 CrossRef CAS PubMed.
  63. S. K. Min, W. Y. Kim, Y. Cho and K. S. Kim, Nat. Nanotechnol., 2011, 6, 162–165 CrossRef CAS PubMed.
  64. N. J. Singh, S. K. Min, D. Y. Kim and K. S. Kim, J. Chem. Theory Comput., 2009, 5, 515–529 CrossRef CAS.
  65. Y. F. Chen, J. Y. Lee, R. Babarao, J. Li and J. W. Jiang, J. Phys. Chem. C, 2010, 114, 6602–6609 CAS.
  66. S. S. Han, D. Kim, D. H. Jung, S. Cho, S.-H. Choi and Y. Jung, J. Phys. Chem. C, 2012, 116, 20254–20261 CAS.
  67. C. M. Tenney and C. M. Lastoskie, Environ. Prog., 2006, 25, 343–354 CrossRef CAS.
  68. S. Vaupel, B. Brutschy, P. Tarakeshwar and K. S. Kim, J. Am. Chem. Soc., 2006, 128, 5416–5426 CrossRef CAS PubMed.
  69. E. C. Lee, B. H. Hong, J. Y. Lee, J. C. Kim, D. Kim, Y. Kim, P. Tarakeshwar and K. S. Kim, J. Am. Chem. Soc., 2005, 127, 4530–4537 CrossRef CAS PubMed.
  70. K. Muller-Dethlefs and P. Hobza, Chem. Rev., 2000, 100, 143–167 CrossRef PubMed.
  71. J.-H. Choi and J.-H. Cho, J. Am. Chem. Soc., 2006, 128, 3890–3891 CrossRef CAS PubMed.
  72. C.-C. Hwang, J. J. Tour, C. Kittrell, L. Espinal, L. B. Alemany and J. M. Tour, Nat. Commun., 2014, 5, 3961 CAS.
  73. S. L. Qiao, Z. K. Du, W. Huang and R. Q. Yang, J. Solid State Chem., 2014, 212, 69–72 CrossRef CAS PubMed.
  74. V. Chandra, S. U. Yu, S. H. Kim, Y. S. Yoon, D. Y. Kim, A. H. Kwon, M. Meyyappan and K. S. Kim, Chem. Commun., 2012, 48, 735–737 RSC.
  75. D. Kajiya and K. Saitow, J. Phys. Chem. B, 2010, 114, 16832–16837 CrossRef CAS PubMed.
  76. R. Dey, R. Haldar, T. K. Maji and D. Ghoshal, Cryst. Growth Des., 2011, 11, 3905–3911 CAS.
  77. M. G. Rabbani and H. M. El-Kaderi, Chem. Mater., 2011, 23, 1650–1653 CrossRef CAS.
  78. M. S. Shannon, J. M. Tedstone, S. P. O. Danielsen, M. S. Hindman and J. E. Bara, Energy Fuels, 2013, 27, 3349–3357 CrossRef CAS.
  79. M. Prakash, K. Mathivon, D. M. Benoit, G. Chambaud and M. Hochlaf, Phys. Chem. Chem. Phys., 2014, 16, 12503–12509 RSC.
  80. S. Altarawneh, S. Behera, P. Jena and H. M. El-Kaderi, Chem. Commun., 2014, 50, 3571–3574 RSC.
  81. H.-S. Byun, N.-H. Kim and C. Kwak, Fluid Phase Equilib., 2003, 208, 53–68 CrossRef CAS.
  82. S.-Y. Duan, X.-M. Ge, N. Lu, F. Wu, W. Yuan and T. Jin, Int. J. Nanomed., 2012, 7, 3813–3822 CAS.
  83. M. A. Alkhabbaz, R. Khunsupat and C. W. Jones, Fuel, 2014, 121, 79–85 CrossRef CAS PubMed.
  84. K. Ahmad, O. Nowla, E. M. Kennedy, B. Z. Dlugogorski, J. C. Mackie and M. Stockenhuber, Energy Technol., 2013, 1, 345–349 CrossRef CAS.
  85. J.-X. Hu, H. Shang, J.-G. Wang, L. Luo, Q. Xiao, Y.-J. Zhong and W.-D. Zhu, Ind. Eng. Chem. Res., 2014, 53, 11828–11837 CrossRef CAS.
  86. M. Saleh, S. B. Baek, H. M. Lee and K. S. Kim, J. Phys. Chem. C, 2015, 119, 5395–5402 CAS.

This journal is © the Owner Societies 2015
Click here to see how this site uses Cookies. View our privacy policy here.