In silico studies on the origin of selective uptake of carbon dioxide with cucurbit[7]uril amorphous material

Debashis Sahu and Bishwajit Ganguly*
Computation and Simulation Unit, Analytical Discipline & Centralized Instrument Facility, and Academy of Scientific and Innovative Research, CSIR-Central Salt and Marine Chemicals Research Institute, Bhavnagar, Gujarat 364002, India. E-mail: ganguly@csmcri.org; Fax: +91-278-2567562

Received 8th July 2015 , Accepted 18th August 2015

First published on 19th August 2015


Abstract

The efficient capture and storage of flue gases is of current interest due to environmental problems. We report the adsorption of flue gases (CO2, N2 and CH4) on amorphous solid Cucurbit[7]uril (CB[7]) computationally. The DFT calculations revealed that CO2 can be adsorbed more strongly inside the cavity of CB[7] compared to N2 and CH4 molecules. The glycoluril units of CB[7] are the preferential sites for the adsorption of CO2 gas molecules. The cooperative binding of CO2 molecules inside the cavity of CB[7] has been observed. The geometrical analysis reveals that the carbon atom of CO2 is in close proximity to the nitrogen atom of the glycoluril of CB[7] and the CO2 oxygen atom is in close contact to the carbonyl carbon of the glycoluril unit. The calculated results show that four CO2 and four CH4 molecules can reside inside the CB[7] cavity. However, five N2 gas molecules can be accommodated inside the CB[7] cavity. The energy decomposition analysis (EDA) performed with the adsorbed CO2 on the wall of CB[7] shows that the dispersive force is playing an important role for the uptake of CO2 inside the cavity. The process of desorption was also examined with the desorption enthalpies (ΔHDE) calculated per gas molecule, which suggests that both adsorption and desorption processes are kinetically feasible. The origin of the interactions between the amorphous solid CB[7] and the flue gases can help to design materials to maximize the capture and separation of such gases.


Introduction

Molecular gas recognition is an emerging area of research that has various applications in gas sensing, gas storage, gas purification, gas conversion, and as biosensors.1,2 Gas encapsulation within the cavities of synthetic organic molecules has been explored with different systems in solution and/or in the solid state.1 The threat of global warming is one of the most pressing environmental concerns in recent times.3 Power plants and steel mills are the main stationary source for the emission of CO2 in nature.4 The 40–50% of total CO2 emission is coming out of such stationary sources.4 In the steel making process, CO gas, in particular, is used as a reducing agent, which is converted into CO2 gas and emitted from the plants.4 This is a challenge in developing new materials for the storage, separation and recycling of the important gases such as CO2 and H2 in industrial plants.5 In recent years, metal–organic frameworks (MOFs)6–10 or covalent organic frameworks (COFs)11,12 have been reported in the literature, which showed remarkable CO2 sorption capacity and selectivity with an exceptionally high surface area. Such MOFs and COFs can be modified easily to enhance the storage capacity. However, these materials are sensitive to moisture (a major component in flue gas) that dampens the efficient capture of CO2.4,13 A few organic molecules like calixarenes,14 cucurbit[n]uril,1,4,13,15,16 carbon nanotubes and activated carbon5 have been explored as porous materials for gas storage and separation. These organic porous materials can be used as appropriate candidate for such adsorbents that can be easily synthesized from the readily available organic molecules. Recently, organic host cucurbit[n]uril (CB[n], n = 6, 7) is used as good alternatives for the CO2 adsorbents4,13,15 and can also be used effectively for the separation of CO2 from different gas mixtures.17 This pumpkin shape cucurbiturils (CBs) have become an extremely attractive class of supramolecular synthetic receptors18,19 with different molecular applications.20,21 Cucurbiturils (CBs) are cyclic oligomers of n glycoluril units linked by 2n methylene groups.22 The CBs possess two hydrophilic carbonylated rims and a hydrophobic nano-cavity.23 The synthetic macro-polycyclic receptors CB[n] have excellent ability to bind various organic, inorganic, biological molecules and ions in the aqueous phase as well as in the solid state.24,25 The amorphous solids CB[6] and CB[7] have been explored as sorbents for the flue gases such as CO2, N2 and CH4.13 CB[7] has shown remarkable selectivity of CO2 over N2 and CH4.13 At 297 K and 1 bar pressure, the CO2 storage capacity of amorphous CB[7] is 2.3 mmol g−1 (50 cm3 g−1) and 1.2 mmol g−1 (25.2 cm3 g−1) for CB[6], respectively.13 In this article, we have examined the origin and magnitude of the possible interactions of the flue gases with CB[7] employing DFT (M06-2X) functional level of theory.26,27 The interactions have been segregated by EDA and non-covalent interactions (NCI) surface analyses.

Computational details

Full geometrical optimizations have been carried out in the gas phase employing the M06-2X level28–30 with standard 6-31G(d) basis set. The hybrid meta-functional M06-2X has been considered as an excellent DFT functional considering the non-covalent interactions31 and it predicts the accurate valence and Rydberg electronic excitation energies for main group chemistry.28 Frequency calculations were performed at the same level of theory, to confirm that each stationary point is a local minimum (with zero imaginary frequency). The single-point calculations with M06-2X/6-311+G(d,p) have also been performed on M06-2X/6-31G(d) optimized geometries. The calculations were performed with the placement of gas molecules either at the center of the host molecule or else near to the surface walls. The calculated results provide the optimum positions of gas molecules with respect to the host cavity site. Similar study was also performed with the gas molecules outside the surface of CB[7]. Generally, Minnesota functional M06-2X shows good results for short- and medium-range dispersions however, this functional is not suitable for long-range dispersion in larger systems.32–35 The calculated results obtained with M06-2X functional was further compared with MO6-2X-D3 results for representative systems, where the long-range dispersion is considered using DFT-D3 treatment.32 Furthermore, the importance of dispersion corrections was revealed through the calculations performed with the standard other functional like PBEPBE36 using 6-31G(d) basis sets. NMR chemical shifts were calculated by the GIAO method37 at the M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory to correlate the observed experimental NMR shifts. All the calculations have been performed with GAUSSIAN 09 suite of program.38

The adsorption enthalpy (ΔH) is calculated as:

 
ΔH = [HS(gas)n − (HS + nHgas)] (1)

The average binding enthalpy per gas molecule (Hb) is calculated as:

 
Hb = (1/n)[HS(gas)n − (HS + nHgas)] (2)

Similarly, the adsorption energy (ΔE)39,40 is calculated as:

 
ΔE = [ES(gas)n − (ES + nEgas)] (3)

The desorption enthalpy (ΔHDE)39–41 is calculated as:

 
ΔHDE = Hgas + (1/m)[HS(gas)nmHS(gas)n] (4)
where, ‘S’ denotes the host molecule (CB[7]) and ‘gas’ denotes flue gas (CO2/N2/CH4). [n = total no. of gas molecule and m = no. of the desorbed gas molecules.]

The wave function analysis in this study is performed in the multifunctional MULTIWFN-3.2.1 suite of program.42 To study the electron transfer of the trapped gas molecules inside the CB[7] cavity, we have plotted the difference map of electron density by using the MULTIWFN-3.2.1 program.42 The Energy Decomposition Analysis (EDA)43 has been performed at the same level of theory using GAMESS software.44 In the EDA analysis, the total interaction energy is divided into the five parts, viz. electrostatic, exchange, repulsion, polarization and DFT dispersion energies.45

Results and discussion

We have examined the origin and nature of the binding affinity of gas molecules (CO2, N2 and CH4) with CB[7] using DFT calculations. The physisorption of the gas molecules was performed with the single unit of CB[7] for computational simplicity. The gas molecules (CO2, N2 and CH4) were adsorbed sequentially inside the cavity of the CB[7] and the binding enthalpies (ΔH) were calculated with M06-2X/6-31G(d) level of theory. The adsorption process was initiated with only one gas molecule inside the cavity of CB[7] (Fig. 1). The M06-2X/6-31G(d) calculated results showed that the CO2 and CH4 molecules reside near the glycoluril-wall of the CB[7] cavity, however, the N2 molecule is present at the center of the CB[7] cavity due to the repulsive interactions with the peripheral wall of the CB[7] (Fig. 1). The calculated structure shows that the carbon atom of CO2 is in close proximity to the nitrogen atom of the glycoluril of CB[7] (Fig. 1). Further, the CO2 oxygen atom is also in close contact to the carbonyl carbon of glycoluril unit. Such interactions were observed in the 13C MAS NMR spectrum study.13 The 13C MAS NMR spectrum showed that there is an upfield shift of Δδc = −1.8 ppm for the stored CO2 and is presumably due to the interaction of gas molecule with the negatively charged carbonyl oxygens and or electron donating nitrogen atoms of the host molecule.13 The NMR (GIAO) calculations performed with CB[7] and CO2 molecule, we have also observed the similar upfield shift (Δδc = −1.7 ppm) for the stored CO2 inside the cavity of the CB[7] at M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory. The M06-2X/6-31G(d) calculated binding enthalpy (ΔH) of single CO2 is −8.4 kcal mol−1, which is in good agreement with the experimental enthalpy at zero coverage.13 The M06-2X/6-31G(d) binding enthalpies (ΔH) calculated for single N2 and CH4 are 0.0 kcal mol−1 and −4.6 kcal mol−1, respectively (Fig. 1). The calculated free energy results also support the higher adsorption of CO2 inside the CB[7] cavity compared to the other two gas molecules (see ESI, Table S1). The role of temperature on the adsorption of gases with CB[7] has also been considered. The increase in temperature can make the adsorption of CO2 be more selective than N2 and CH4 gas molecules with CB[7] host molecule assuming the pressure remains constant in such cases (see ESI, Table S1). The volume of host molecule to adsorb the gas molecules may also be important. The binding ability of CB[6] host molecule with single CO2 is better than CB[7], however, the volume of the host systems can dictate the gas adsorption capacity. The experimental results reveal the higher CO2 adsorption capacity of CB[7] host molecule.13 The computed binding energies (ΔE) with higher basis set with M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory also showed similar trend of interaction of these gas molecules with CB[7] (Fig. 1).
image file: c5ra13394g-f1.tif
Fig. 1 M06-2X/6-31G(d) calculated binding enthalpies (ΔH) of one CO2, N2 and CH4 inside the cavity of CB[7]. The computed binding energies (ΔE) at the M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory are given in the parentheses. All the energies are in kcal mol−1 and distances are in Å. (Colour of atoms: grey – C, white – H, blue – N and red – O.)

We have extended our computational experiment with two gas molecules placed inside the CB[7] cavity. The calculated M06-2X/6-31G(d) results show that the gas molecules are close to the glycoluril units of the host molecule in parallel fashion in all cases (Fig. 2). Here, the binding enthalpy of two CO2 molecules inside the cavity of CB[7] is −17.6 kcal mol−1, which is higher than the corresponding N2 (−9.9 kcal mol−1) and methane gas molecules (−10.1 kcal mol−1). Interestingly, the N2 molecules prefer to reside near the glycoluril unit of the host molecule and such a situation arises due to the electrostatic repulsions between the two N2 molecules inside the cavity of CB[7] (Fig. 2). The interaction of glycoluril units of CB[7] with the CH4 molecules is similar to that obtained with the single CH4 molecule (Fig. 1 and 2). The average binding enthalpies (Hb) of CO2, N2 and CH4 with CB[7] are −8.8, −4.9 and −5.0 kcal mol−1, respectively, which are comparable to the binding affinities of single gas molecule with CB[7] (Fig. 1 and 2). Since the long-range dispersion is not considered in the M06-2X functional, we have examined the long range dispersion by using D3-zero damping parameters for M06-2X functional.32–35 The dispersion corrected M06-2X-D3 binding enthalpies are slightly higher than the uncorrected M06-2X binding enthalpies however, the trends is similar in both cases (see ESI, Table S2). M06-2X-D3 calculated results predict the binding enthalpy even closer to the reported enthalpy of adsorption of CO2 molecules at zero coverage.13 The calculations performed with the standard PBEPBE/6-31G(d) level showed much lower binding enthalpies of the gas molecules with CB[7] compared to the M06-2X results (see ESI, Table S2).


image file: c5ra13394g-f2.tif
Fig. 2 M06-2X/6-31G(d) calculated binding enthalpies (ΔH) of two CO2, N2 and CH4 inside the cavity of CB[7] and their corresponding average binding enthalpies (Hb). The computed binding energies (ΔE) at the M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory are given in the parentheses. All the energies are in kcal mol−1 and distances are in Å. (Colour of atoms: grey – C, white – H, blue – N and red – O.)

We have extended our computational study with three gas molecules placed inside the CB[7] cavity. The binding enthalpy (ΔH) of three CO2 molecules with the host molecule is −26.5 kcal mol−1, which is higher than N2 (−13.7 kcal mol−1) and CH4 (−15.7 kcal mol−1), respectively (Fig. 3). The average binding enthalpies (Hb) of CO2, N2 and CH4 remain to be similar as observed with the previous cases (Fig. 1–3). The gas molecules entrapped inside the cavity of CB[7] showed cooperative interactions among them. The gas molecules were taken out from the cavity of CB[7] and calculations were performed with orientations as obtained in the host cavity. The cooperative binding energy of three CO2 molecules is −4.3 kcal mol−1, whereas, such interactions with N2 and CH4 are −0.9, −1.0 kcal mol−1 respectively (Fig. 4). Similar cooperative interactions for CO2 molecules were also reported for the MOF [Zn2(Atz)2(ox), Atz: 3-amino-1,2,4-triazole; ox: oxalate] in the literature.46 These results further suggest the cooperative effect among CO2 molecules is also responsible for the enhanced binding energy compared to N2 and CH4 inside the cavity of CB[7].


image file: c5ra13394g-f3.tif
Fig. 3 M06-2X/6-31G(d) calculated binding enthalpies (ΔH) of three CO2, N2 and CH4 inside the cavity of CB[7] and the corresponding average binding enthalpies (Hb). The computed binding energies (ΔE) at the M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory are given in the parentheses. All the energies are in kcal mol−1 and distances are in Å. (Colour of atoms: grey – C, white – H, blue – N and red – O.)

image file: c5ra13394g-f4.tif
Fig. 4 Cooperative intermolecular interactions of the three CO2, N2 and CH4 inside the cavity of CB[7]. M06-2X/6-31G(d) calculated energies are in kcal mol−1 and distances are in Å.

The similar trend of binding enthalpies are also observed for the four gas molecules inside the CB[7] cavity. The calculated binding enthalpy (ΔH) of the CO2 is −34.2 kcal mol−1 compared to N2 (−19.6 kcal mol−1) and CH4 (−21.5 kcal mol−1) (Fig. 5) and, the average binding enthalpies (Hb) of CO2, N2 and CH4 are −8.6, −4.9 and −5.4 kcal mol−1 respectively (Fig. 5). M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) calculated binding energy (ΔE) results show the similar trend as obtained with M06-2X/6-31G(d) calculated binding enthalpies (ΔH) (Fig. 5). The CO2 molecules are close to the glycoluril wall of the CB[7] like the previous cases (Fig. 1–3) and interacting strongly with each other (Fig. 6). The cooperative enhancement effect of CO2 molecules is also responsible for the strong CO2 adsorption (Fig. 6).46 The N2 molecules oriented inside the cavity of CB[7] also showed some attractive interactions, presumably due to the induced dipoles created in presence of these gas molecules. Desorption of gas molecules from the host molecule is also important for practical purposes. We have performed the calculations to examine the desorption (inverse of the adsorption process)39,40 of the gas molecules from the CB[7]. The M06-2X/6-31G(d) calculated desorption enthalpies (ΔHDE) for CO2, N2 and CH4 are 9.5, 3.9 and 5.5 kcal mol−1 respectively. These results suggest that both adsorption and desorption processes are kinetically feasible.


image file: c5ra13394g-f5.tif
Fig. 5 M06-2X/6-31G(d) calculated binding enthalpies (ΔH) of four CO2, N2 and CH4 inside the cavity of CB[7] and the corresponding average binding enthalpies (Hb). The computed binding energies (ΔE) at the M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory are given in the parentheses. All the energies are in kcal mol−1 and distances are in Å. (Colour of atoms: grey – C, white – H, blue – N and red – O.)

image file: c5ra13394g-f6.tif
Fig. 6 Cooperative intermolecular interactions of the four CO2, N2 and CH4 inside the cavity of CB[7]. M06-2X/6-31G(d) calculated energies are given in kcal mol−1 and distances are in Å.

The cooperative effect among the gas molecules and their interactions with the CB[7] was further exemplified with non-covalent interactions (NCI) isosurface plots (Fig. 7). NCI isosurfaces are generally used to understand the nature of interaction in different types of systems.47 The NCI results reveal that the flue gas molecules have attractive van der Waals interaction14,47–49 among themselves and also with the host walls as represented in colour-coding scheme.


image file: c5ra13394g-f7.tif
Fig. 7 Noncovalent interactions (NCI) isosurface plots for the trapped gas molecules (CO2, N2 and CH4) inside the CB[7] cavity.

The nicely packed gas molecules inside the cavity of CB[7] can exhibit electron transfer between the host and the trapped gas molecules (CO2, N2 and CH4). We have calculated the difference map of electron density of the CB[7] and the trapped gas molecules to understand the electron transfer between the host and guest systems in MULTIWFN-3.2.1 program42 (Fig. 8). The electron transfer is significantly higher in the case of CO2 compared to N2 and CH4 trapped gas molecules inside the CB[7].


image file: c5ra13394g-f8.tif
Fig. 8 The difference map of electron density of the CB[7] and the trapped gas molecules. The red and yellow isosurfaces represent the region in which electron density is increased and decreased after gas molecules trapped to CB[7], respectively.

To examine further, we have incorporated gas molecules inside the cavity of CB[7], and it appears that the fifth gas molecule for CO2 and CH4 moves away from the cavity of the host molecule (Fig. 9). However, N2 being smaller in size, five such molecules can reside inside the cavity of a single CB[7].


image file: c5ra13394g-f9.tif
Fig. 9 M06-2X/6-31G(d) optimized structures of the five gas molecules (CO2, N2 and CH4) with CB[7] (side view) with their corresponding binding enthalpies (ΔH) and average binding enthalpies (Hb). The computed binding energies at the M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory are given in the parentheses. All the energies are in kcal mol−1 and distances are in Å. (Colour of atoms: grey – C, white – H, blue – N and red – O.)

The experimental results support that four CO2 gas molecules can be accommodated inside the cavity of CB[7].13 Further to note that the maximum sorption capacity of amorphous CB[7] for CO2 is 118 cm3 g−1 (5.3 mmol g−1) at 1 bar and 196 K, which corresponds to 6.1 CO2 molecules per CB[7]. These results suggest that the two CO2 molecules are present in the interstitials voids of the host molecules.13 We have placed two CO2 molecules outside the cavity of CB[7], while keeping the four CO2 molecules inside the host molecule (Fig. 10). The M06-2X/6-31G(d) calculated results showed that all six CO2 molecules are bound strongly with the host system compared to the other two gas molecules (N2 and CH4) (Fig. 10).


image file: c5ra13394g-f10.tif
Fig. 10 M06-2X/6-31G(d) optimized structures of the four gas molecules (CO2, N2 and CH4) inside the cavity of CB[7] while two gas molecules at the outside the cavity of CB[7] with their corresponding binding enthalpies (ΔH) and average binding enthalpies (Hb). The computed binding energies at the M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory are given in the parentheses. All the energies are in kcal mol−1 and distances are in Å. (Colour of atoms: grey – C, white – H, blue – N and red – O.)

The study was performed with the adsorption of mixed gas molecules inside the cavity of CB[7]. The adsorption of N2 and CH4 gas molecules in the presence of two CO2 molecules is weaker compared to the adsorption of another CO2 molecule at the same level of theory. Therefore, it appears that the selectivity of CO2 adsorption with CB[7] is higher even in mixed environment of flue gas molecules (see ESI, Fig. S2).

The origin and magnitude of CO2 interaction with CB[7] was examined with Energy Decomposition Analysis (EDA).43 The EDA analysis was modelled with three glycoluril units of CB[7] and the CO2 molecule. This geometry has been taken from the adsorbed CO2 molecule inside the cavity of CB[7] (Fig. 1). The M06-2X/6-31G(d) calculated results with the model geometry reveal that the total interaction energy is −5.9 kcal mol−1, which is further segregated into five energy components i.e., electrostatic (−7.0 kcal mol−1), exchange (−6.4 kcal mol−1), repulsion (29.6 kcal mol−1), polarization (−1.4 kcal mol−1) and DFT dispersion energies (−20.5 kcal mol−1). EDA calculations are performed with the model systems, (Fig. 11) which show that the role of dispersive interaction is predominant in nature towards the adsorption of CO2 on the cavity of CB[7]. This result is in agreement with the reports on CO2 adsorption with MOF systems [Zn2(Atz)2(ox), Atz: 3-amino-1,2,4-triazole; ox: oxalate] reported in the literature.46 The interaction energy of CO2 with the CB[7] moiety was further calculated with basis set superposition error (BSSE) correction via counterpoise procedure (CP) method with M06-2X/6-31G(d) level of theory. The calculated corrected complexation energy of CO2 with CB[7] moiety is −5.9 kcal mol−1, which is in reasonable agreement with the uncorrected BSSE interaction energy (−5.9 kcal mol−1).


image file: c5ra13394g-f11.tif
Fig. 11 CO2 molecule interacting towards the inner side of the CB[7] moiety.

Therefore, the higher binding enthalpy of CO2 with the walls of CB[7] arises due to the dispersion interactions. The strong quadruple moment of the CO2 molecule may help to associate with other CO2 gas molecules inside the cavity of CB[7], which in turn can augment the adsorption process.46,50 The preferred binding of CH4 than N2 in all the cases can be explained based on the polarizability data.46 The higher polarizability of CH4 (12.5 Bohr3) compared to N2 (8.4 Bohr3) calculated with M06-2X/6-31G(d) level of theory rationalizes the observed results. It can be mentioned here that the binding enthalpies of the CO2, N2 and CH4 outside the cavity of CB[7] are generally lower compared to the inside the cavity of CB[7] cases (see ESI, Fig. S1).

Conclusions

This work demonstrates the origin of the physisorption of the gas species (CO2, N2 and CH4) by CB[7] in a systematic study. The DFT (M06-2X) calculated results show that four CO2 and CH4 molecules can reside inside the CB[7] cavity, while five N2 gas molecules can be accommodated in the cavity of host molecule. The M06-2X/6-31G(d) calculations suggest that the binding enthalpy of CO2 with CB[7] is higher compared to the other flue gases (N2 and CH4). The binding enthalpy of four CO2 molecules is −34.2 kcal mol−1 inside the CB[7] cavity, which is much higher than the N2 (−19.6 kcal mol−1) and CH4 (−21.5 kcal mol−1), respectively. The DFT-D3 corrected M06-2X functional results indicate that the binding enthalpy of the adsorbed CO2 gas molecule inside the CB[7] cavity much closer to the reported enthalpy of adsorption of CO2 molecules at zero coverage.13 The 13C NMR (GIAO) calculations at M06-2X/6-311+G(d,p)//M06-2X/6-31G(d) level of theory level of theory shows that the up-field shift of the carbon atom of CO2 while interacting with the glycoluril wall of CB[7], which corroborates the experimentally observed results. The Energy Decomposition Analysis (EDA) reveals that the dispersive force is predominant for the adsorption of CO2 with the glycoluril units of CB[7]. The preferential adsorption of CH4 over N2 attributes to the higher polarizability of the CH4 over N2. These results may help to explore the potentials of cucurbiturils for separation and storage of gas molecules.

Acknowledgements

CSIR-CSMCRI Communication No. 149/2014. The authors thank DBT, DST, CSIR (MSM, SIP) New Delhi, India and DAE-BRNS Mumbai, India for financial support. D. S. is thankful to UGC, New Delhi, India, for awarding a senior research fellowship. D. S. is also thankful to AcSIR for enrolment in the Ph.D. program. The authors thankfully acknowledge the computer resources provided by NCL, Pune, India, and CMMACS, Bangalore, India. The authors are also thankful to Dr Rabindranath Lo for his valuable suggestions in this work. We thank the reviewers' for their valuable comments/suggestions that have helped us to improve the paper.

References

  1. L. Fusaro, E. Locci, A. Lai and M. Luhmer, J. Phys. Chem. B, 2008, 112, 15014–15020 CrossRef CAS.
  2. D. M. Rudkevich, Angew. Chem., Int. Ed., 2004, 43, 558–571 CrossRef CAS.
  3. D. M. D'Alessandro, B. Smit and J. R. Long, Angew. Chem., Int. Ed., 2010, 49, 6058–6082 CrossRef PubMed.
  4. H. Kim, Y. Kim, M. Yoon, S. Lim, S. M. Park, G. Seo and K. Kim, J. Am. Chem. Soc., 2010, 132, 12200–12202 CrossRef CAS PubMed.
  5. J. Tian, P. K. Thallapally, S. J. Dalgarno, P. B. McGrail and J. L. Atwood, Angew. Chem., 2009, 48, 5492–5495 CrossRef CAS PubMed.
  6. D. Britt, H. Furukawa, B. Wang, T. G. Glover and O. M. Yaghi, Proc. Natl. Acad. Sci. U. S. A., 2009, 106, 20637–20640 CrossRef CAS PubMed.
  7. C. A. Fernandez, S. K. Nune, R. K. Motkuri, P. K. Thallapally, C. Wang, J. Liu, G. J. Exarhos and B. P. McGrail, Langmuir, 2010, 26, 18591–18594 CrossRef CAS PubMed.
  8. J.-R. Li, Y.-G. Ma, M. C. McCarthy, J. Sculley, J.-M. Yu, H.-K. Jeong, P. B. Balbuena and H.-C. Zhou, Coord. Chem. Rev., 2011, 255, 1791–1823 CrossRef CAS.
  9. S. Ma and L. Meng, Pure Appl. Chem., 2011, 83, 167–188 CAS.
  10. J. J. Gassensmith, H. Furukawa, R. A. Smaldone, R. S. Forgan, Y. Y. Botros, O. M. Yaghi and J. F. Stoddart, J. Am. Chem. Soc., 2011, 133, 15312–15315 CrossRef CAS.
  11. M. Mastalerz, Angew. Chem., Int. Ed., 2008, 47, 445–447 CrossRef CAS PubMed.
  12. M. G. Schwab, B. Fassbender, H. W. Spiess, A. Thomas, X. Feng and K. Müllen, J. Am. Chem. Soc., 2009, 131, 7216–7217 CrossRef CAS PubMed.
  13. J.-A. Tian, S.-Q. Ma, P. K. Thallapally, D. Fowler, B. P. McGrail and J. L. Atwood, Chem. Commun., 2011, 47, 7626–7628 RSC.
  14. J. L. Atwood, L. J. Barbour and A. Jerga, Science, 2002, 296, 2367–2369 CrossRef CAS PubMed.
  15. J. Tian, J. Liu, J. Liu and P. K. Thallapally, CrystEngComm, 2013, 15, 1528–1531 RSC.
  16. S. Pan, S. Mondal and P. K. Chattaraj, New J. Chem., 2013, 37, 2492–2499 RSC.
  17. Y. Jin, B. A. Voss, R. D. Noble and W. Zhang, Angew. Chem., Int. Ed., 2010, 49, 6348–6351 CrossRef CAS PubMed.
  18. E. Masson, X. X. Ling, R. Joseph, L. Kyeremeh-Mensah and X. Y. Lu, RSC Adv., 2012, 2, 1213–1247 RSC.
  19. J. W. Lee, S. Samal, N. Selvapalam, H.-J. Kim and K. Kim, Acc. Chem. Res., 2003, 36, 621–630 CrossRef CAS PubMed.
  20. J. Lagona, P. Mukhopadhyay, S. Chakrabarti and L. Isaacs, Angew. Chem., Int. Ed., 2005, 44, 4844–4870 CrossRef CAS PubMed.
  21. D. Sahu and B. Ganguly, Tetrahedron Lett., 2013, 54, 5246–5249 CrossRef CAS PubMed.
  22. D. H. Macartney, Isr. J. Chem., 2011, 51, 600–615 CrossRef CAS PubMed.
  23. B. C. Pemberton, R. Raghunathan, S. Volla and J. Sivaguru, Chem.–Eur. J., 2012, 18, 12178–12190 CrossRef CAS PubMed.
  24. O. Danylyuk and V. P. Fedin, Cryst. Growth Des., 2012, 12, 550–555 CAS.
  25. A. Suvitha, N. S. Venkataramanan, H. Mizuseki, Y. Kawazoe and N. Ohuchi, J. Inclusion Phenom. Macrocyclic Chem., 2009, 66, 213–218 CrossRef.
  26. W. J. Hehre, J. Chem. Phys., 1972, 56, 2257–2261 CrossRef CAS PubMed.
  27. M. M. Francl, J. Chem. Phys., 1982, 77, 3654–3665 CrossRef CAS PubMed.
  28. Y. Zhao and D. G. Truhlar, Acc. Chem. Res., 2008, 41, 157–167 CrossRef CAS PubMed.
  29. Y. Zhao and D. G. Truhlar, J. Chem. Theory Comput., 2008, 4, 1849–1868 CrossRef CAS.
  30. E. G. Hohenstein, S. T. Chill and C. D. Sherrill, J. Chem. Theory Comput., 2008, 4, 1996–2000 CrossRef CAS.
  31. Y. Zhao and D. Truhlar, Theor. Chem. Acc., 2008, 120, 215–241 CrossRef CAS.
  32. L. Goerigk, H. Kruse and S. Grimme, ChemPhysChem, 2011, 12, 3421–3433 CrossRef CAS PubMed.
  33. S. Grimme, Chem.–Eur. J., 2012, 18, 9955–9964 CrossRef CAS PubMed.
  34. D. Roy, M. Marianski, N. T. Maitra and J. J. Dannenberg, J. Chem. Phys., 2012, 137, 134109 CrossRef PubMed.
  35. R. Sedlak, T. Janowski, M. Pitoňák, J. Ŕezáč, P. Pulay and P. Hobza, J. Chem. Theory Comput., 2013, 9, 3364–3374 CrossRef CAS PubMed.
  36. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS.
  37. K. Wolinski, J. F. Hinton and P. Pulay, J. Am. Chem. Soc., 1990, 112, 8251–8260 CrossRef CAS.
  38. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, and D. J. Fox, Gaussian 09, Revis. B01, Gaussian, Inc., Wallingford, CT, 2010 Search PubMed.
  39. S. Pan, S. Banerjee and P. K. Chattaraj, J. Mex. Chem. Soc., 2012, 56, 229–240 CAS.
  40. S. Pan, S. Giri and P. K. Chattaraj, J. Comput. Chem., 2011, 33, 425–434 CrossRef.
  41. A. K. Biswas, R. Lo, M. K. Si and B. Ganguly, Phys. Chem. Chem. Phys., 2014, 16, 12567–12575 RSC.
  42. T. Lu and F. Chen, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed.
  43. M. von Hopffgarten and G. Frenking, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 43–62 CrossRef CAS PubMed.
  44. M. W. Schmidt, K. K. Baldridge, J. A. Boatz, S. T. Elbert, M. S. Gordon, J. H. Jensen, S. Koseki, N. Matsunaga, K. A. Nguyen, S. Su, T. L. Windus, M. Dupuis and J. A. Montgomery, J. Comput. Chem., 1993, 14, 1347–1363 CrossRef CAS PubMed.
  45. N. Thellamurege and H. Hirao, Molecules, 2013, 18, 6782–6791 CrossRef CAS PubMed.
  46. 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.
  47. S. Pan, S. Mandal and P. K. Chattaraj, J. Phys. Chem. B, 2015 DOI:10.1021/acs.jpcb.5b01396.
  48. Q. Sun, M. Wang, Z. Li, Y. Ma and A. Du, Chem. Phys. Lett., 2013, 575, 59–66 CrossRef CAS PubMed.
  49. G. Román-Pérez, M. Moaied, J. M. Soler and F. Yndurain, Phys. Rev. Lett., 2010, 105, 145901 CrossRef.
  50. P. Chowdhury, C. Bikkina and S. Gumma, J. Phys. Chem. C, 2009, 113, 6616–6621 CAS.

Footnote

Electronic supplementary information (ESI) available: Fig. S1 and S2 and Tables S1 and S2. See DOI: 0.1039/c5ra13394g

This journal is © The Royal Society of Chemistry 2015