DOI:
10.1039/C5RA09884J
(Paper)
RSC Adv., 2015,
5, 57647-57656
Molecular hydrogen binding affinities of metal cation decorated substituted benzene systems: insight from computational exploration†
Received
26th May 2015
, Accepted 25th June 2015
First published on 25th June 2015
Abstract
The binding affinity of hydrogen molecules towards Li+ and Mg2+ decorated C6H5X (X = −CH3, −NH2, −CN, −COOH) systems has been investigated theoretically with special emphasis on the nature of the interaction between metal cations and H2 molecules. Our calculations show that binding of H2 over C6H5X−M (where M = Li+, Mg2+) is improved on moving from Li+ to Mg2+. For both C6H5X−M complexes the electron donating substituents weaken the H2 binding energy considerably whereas electron withdrawing substituents slightly strengthen the interaction relative to the C6H6−M complex. The interaction of H2 molecules with the metal centers in Li+ and Mg2+ decorated C6H5X systems has been explored in the light of AIM formalism, NBO analysis and LMOEDA analysis. The polarization and the charge transfer together stabilize the system whereas the pairwise steric exchange interaction renders destabilization of the system. In the case of Mg2+ decorated systems, the amount of charge transfer from the bonding orbital of the hydrogen molecule to the antibonding lone pair orbital of the metal cation and thereby the polarization factor is much higher than that found in corresponding Li+ decorated systems.
1. Introduction
Use of hydrogen as a practical fuel in the transportation sector is one of the challenging issues because of its safe and efficient storage related difficulties. Extensive efforts are in progress for designing materials with high hydrogen adsorption/desorption characteristics under ambient conditions. Whereas a very weak interaction would lead to lower H2 sorption at reasonable temperature, a very strong interaction, on the other hand, might be irreversible or require extreme conditions to release H2.1 A quasi-molecular type of binding which is intermediate between physisorption and chemisorption is ideal for hydrogen storage under ambient temperatures and pressures.2
Microporous metal organic frameworks (MOFs) are one of the most promising class of hydrogen storage materials because their extremely high surface area helps in reversible H2 uptake and release and also in fast H2 desorption kinetics.3–11 The two main adsorptive sites for H2 on the MOF are the metal oxide center and the organic linker unit. This phenomena is proved by both experimental5,12,13 and various theoretical studies.11,14–17 The more positive H2 binding energy, the stronger hydrogen is bound. The strength of the H2 binding interaction should be in the range of ∼5–10 kcal mol−1 for reversible sorption and desorption of H2 at room temperature and at moderate pressures.1 This range of interaction may not be achievable by choosing only the neutral surface. There are number of ways through which the binding energy of H2 to MOFs can be increased. Dopants directly enhance the hydrogen binding energy of the carbon substrate, and in addition they can act as H2 binding centers. Because of the weak van der Waals interaction between H2 and the carbon support, the binding of hydrogen is very weak. To increase the binding, addition of electrostatic interaction is required and in order to cause this charging, the carbon systems could be doped with highly electropositive Group I and Group II metals. Alkali metal doping onto the organic linker part of MOFs and on carbon based systems have been studied widely and they are proved to be good candidates for this purpose. Han et al. first proposed Li-doped MOFs as a practical hydrogen storage material using an ab initio based GCMC simulation.18 Lochan et al. in their computational study established that strong H2 sorption site can be obtained in MOFs by incorporating open transition metal sites (Cr, Mo, V−, Mn+) on the organic linkers.19 Using DFT calculation Blomqvist et al. showed significant H2 adsorption properties of Li-decorated MOF-5.20 Mavrandonakis et al. showed that Li atom is preferably located on the organic linker part rather than the metal oxide part of MOF.21 However, Dalach and coworkers reported that in Zn2-based MOFs, the Li associates strongly with the metal oxide part and less with the organic linker part.22 In another theoretical work, Kolmann et al. investigated the interaction of one, two and three hydrogen molecules with Li+-doped benzene which acts as a model for lithium-doped carbon-based and metal organic framework materials.23 The interaction of hydrogen molecules with alkali and alkaline earth metal cations (Li+, Na+, K+, Be2+, Mg2+, Ca2+) in MH16 complexes was studied by Chandrakumar et al.24 By detailed energy decomposition analysis (EDA), they showed that both the electrostatic and charge transfer components play significant role in stabilizing those complexes. The polarization component also contributes significantly to the binding energy. In our earlier works we investigated the nature of interaction of different cation−H2 complexes by using symmetry adapted perturbation theory (SAPT).25,26 Srinivasu et al. in their work studied the effect of electron donating and electron withdrawing groups on hydrogen adsorption energy of Li+ doped and Na+ doped benzene ring respectively.27 They also concluded that ionic surface with a significant curvature would enhance the hydrogen adsorption significantly. In another work they further investigated the H2 adsorption properties of Na-doped CnHn systems (n = 4–6, 8).28 The effect of metal ions (Li+, Na+, Be2+, Mg2+ and Al3+) on hydrogen adsorption of benzene system was studied by Maark et al.29 The authors showed that Mg2+ ion-decorated MOF-5 has the most suitable binding energy for the practical H2 storage applications in comparison to other metal cations. Bodrenko et al. studied the hydrogen storage capacity in alkali and alkaline-earth metal decorated aromatic carbon ring based molecular materials.30 In a very recent work of Huang et al., the authors showed that alkali metal cation decorated carbon based molecular complexes with B- and N-substitution enhances the hydrogen storage capacity.31
In the present work, we have explored the viability of improving H2 binding energy by incorporating open metal sites over the organic linkers of MOF and on the carbon based nanomaterials. For this purpose we have used two metal cations, one is from alkali metal family (Li+) and the other one is from alkaline earth metal family (Mg2+). Inspired by the work of Maark et al., as discussed earlier, we choose C6H6–Mg2+ complex, where the benzene moiety acts as a model representing doped BDC linker of MOF-5 and doped polyaromatic carbon-based materials. We have also investigated how different substituents alter the interaction when several electron donating (−CH3, −NH2) and electron withdrawing (−CN, −COOH) groups are introduced in the benzene system. A comparative study between ab initio and different DFT functionals with various basis sets to determine the most appropriate density functional to describe H2 binding energy on the Li+ and Mg2+ decorated benzene and substituted benzene systems has been carried out. The most appropriate density functional, B2PLYPD in conjunction with 6-311+G(d,p) basis set is then used to investigate the binding energies for multiple hydrogen molecules. The interaction between the metal cation and hydrogen molecules is further studied critically employing natural bond orbital (NBO) analysis, topological analysis of charge density and localized molecular orbital energy decomposition analysis (LMOEDA).
2. Computational details
All electronic structure and frequency calculations, NBO analysis are performed using Gaussian 09,32 suite of quantum chemistry program. The DFT functionals considered for the systems under investigation are namely B3LYP,33,34 and B2PLYPD.35,36 Whereas hybrid B3LYP exchange correlation functional is the most widely used density functional, Grimme's dispersion corrected density functional B2PLYPD is known to work well for systems containing long range dispersion interactions. The double hybrid density functional B2PLYPD includes a second-order perturbation correction for nonlocal correlation effect which is comparable to the replacement of semi-local GGA exchange by the exact Hartree–Fock (HF) exchange, as in conventional hybrid functionals. The extension “D” stands for the long range dispersion-corrections that represent the van der Waals interaction. In order to provide benchmark results for comparison, we have also performed ab initio second-order Moller–Plesset perturbation theory (MP2)37–39 calculation.
All the ab initio and DFT calculations associated with the binding of one H2 molecule on the C6H6−M and C6H5X−M complexes are carried out using Pople's 6-31G(d),40 6-311+G(d,p),41,42 basis sets and Ahlrichs' TZVP triple-ζ basis set.43 The adsorption of multiple H2 molecules over these systems has been analyzed by only B2PLYPD functional in conjunction with 6-311+G(d,p) basis set as this level of calculation is expected to provide most promising results as proposed by Grimme and co-workers.35,36 The binding energy (ΔBE) of molecular hydrogen to the system is calculated from the energy difference between the H2 bonded complex, i.e., C6H5X−M with adsorbed hydrogen molecules and the monomers (C6H5X−M and relevant number of hydrogen molecules). The effect of basis-set superposition error (BSSE) has been estimated by the counterpoise (CP) procedure.44 Charge on the metal cation has been obtained at B2PLYPD level by applying natural population analysis (NPA). To assess the feasibility of adding hydrogen molecules one by one on the Li+ decorated system, we have calculated the incremental first-order energy difference by using the following equation
Δ1E(n) = E[(C6H5X–Li+)(H2)n] − E[(C6H5X–Li+)(H2)n−1] − E[H2] |
The above equation is also applicable for Mg
2+ containing systems.
The nature of interaction between the metal cation and hydrogen molecule of each optimized structure is calculated using the topological analysis of atoms in molecule formalism employing AIMAll program.45 For evaluating the strength of this interaction, two important parameters, namely electron density (ρ) and its Laplacian (∇2ρ) at the bond critical point (BCP) are calculated. In order to get a more vivid depiction about the nature of this type of bonding, we have employed natural bond orbital (NBO) analysis by using NBO 6.0 program,46 which is implemented in Gaussians 09. The donor–acceptor (bond–antibond) charge transfer interaction,47–49 say i → j, is evaluated by the delocalization correction energy (ΔECT) calculated using second-order perturbation theory
In the above equation,
F(
i,
j) is the off-diagonal Fock matrix element expressed in the NBO basis,
qi is the donor orbital occupancy and
εi and
εj are the respective orbital energies.
The steric calculations are carried out using Badenhoop–Weinhold procedure,50–52 implemented in NBO 6.0. The pairwise steric exchange interactions between H2 molecule and metal cation as well as among the H2 molecules correspond to the classical “steric hindrance” and is mainly responsible for destabilizing the system. Different contributions to the interaction energy are obtained by using the localized molecular orbital energy decomposition analysis (LMOEDA),53 by using GAMESS electronic structure code.54 The LMOEDA calculations are performed at the B2PLYPD level using the triple-ζ TZV basis set. The instantaneous interaction energy (ΔEint) between two fragments i.e. between hydrogen molecules and C6H5X−M complex under investigation can be decomposed into the terms described in the following equation:
ΔEint = Eelec + Eex-rep + Epol + Edisp |
The classical Coulomb interaction between the occupied orbitals on two fragments is described by electrostatic term Eelec, and Eex-rep is the repulsive exchange component resulting from the Pauli exclusion principle. Epol and Edisp correspond to polarization and dispersion terms respectively. The polarization term is expressed as the orbital relaxation energy that includes both polarization and charge transfer interactions.
3. Results and discussion
In this present investigation, we choose to employ three different basis sets namely 6-31G(d), 6-311+G(d,p) and TZVP with ab initio MP2 and density functionals B3LYP and B2PLYPD in order to calibrate the most suitable level of theory. Table S1 and S2 in the ESI† contain all the detailed hydrogen binding energy values of the investigated systems i.e. C6H6−M and C6H5X−M complexes. As hydrogen binds weakly to the metal center, an accurate treatment of dispersion correction to measure long range interaction is very vital for studying the present systems.27,28 Therefore as outlined in the computational section, Grimme's dispersion corrected density functional B2PLYPD in conjunction with 6-311+G(d,p) basis set is expected to be the most suitable combination in describing the energetics of all the systems under investigation. The following discussions are based on the B2PLYPD/6-311+G(d,p) geometries and energies, unless otherwise stated.
3.1 Interaction of hydrogen with C6H6–M complexes
While analyzing the interaction of hydrogen with C6H6–Li+ complex, we considered the possible maximum number of hydrogen molecules that can be adsorbed on the complex in accordance with the existing theoretical investigations.23,27,29 As illustrated in Fig. 1(a), Li+ ion can accommodate maximum three hydrogen molecules. This is also reflected in their Δ1E(n) values shown in Table 1. On addition of H2 molecules from one to three, the stability of the complexes increases with respect to their lower homologue. But 4li–ben, containing four hydrogen molecules, is destabilized by an amount of 0.31 kcal mol−1 energy as compared to 3li–ben, implying that further addition of hydrogen is energetically not favorable. As the number of hydrogen molecules increases from one to three, the arrangement of H2 around Li+ changes in order to accumulate additional H2 molecules having average Li+−H2 distance in the range 1.9 Å and 2.0 Å in all the cases. When the fourth hydrogen molecule is attached to the complex it moves away from the system by making a distance of ∼3.7 Å from the Li+ cation. According to Srinivasu et al.,27 smaller ionic radius and high effective charge density of the lithium ion is responsible for the steric crowding occurring among the hydrogen molecules around it and for this reason Li+ cannot accommodate more than three hydrogen molecules. As evident from the optimized geometries of 1li–ben, 2li–ben, 3li–ben and 4li–ben of Fig. 1(a), the increased Li+−H2 distance is an indication of gradual weakening of interaction with increased number of hydrogen molecules. The molecular hydrogen binding energy is −4.05 kcal mol−1 in 1li–ben complex while the NPA charge of Li+ is found to be 0.915 a.u. With the gradual increase of the number of hydrogen molecules the interaction becomes weaker, the value of average H2 binding energy being −2.88 kcal mol−1, −2.10 kcal mol−1 and −1.50 kcal mol−1 in 2li–ben, 3li–ben and 4li–ben, respectively. Charge on the metal cation accordingly decreases in these complexes in the range 0.884 a.u. to 0.857 a.u., which is consistent with the decreasing order of ΔBE value.
 |
| Fig. 1 Optimized geometries of C6H6–Li+(H2)n (a) and C6H5X–Li+(H2)n [(b)–(e)] complexes where n = 1–4. DLi–H2 represents the average bond length between Li+ and H2 molecules. All the bond lengths are given in Å unit. | |
Table 1 B2PLYPD BSSE corrected H2 binding energy (ΔBE) and incremental first-order energy difference [Δ1E(n)] values (kcal mol−1) of Li+ decorated systems and charge on the cation (a.u.)
System |
ΔBE |
Δ1E(n) |
Charge on cation |
1li–ben |
−4.05 |
— |
0.915 |
2li–ben |
−2.88 |
−1.70 |
0.884 |
3li–ben |
−2.10 |
−0.56 |
0.857 |
4li–ben |
−1.50 |
0.31 |
0.857 |
1li–aniline |
−3.71 |
— |
0.916 |
2li–aniline |
−2.79 |
−1.87 |
0.883 |
3li–aniline |
−2.14 |
−0.84 |
0.855 |
4li–aniline |
−1.50 |
0.43 |
0.855 |
1li–methyl |
−3.96 |
— |
0.916 |
2li–methyl |
−2.82 |
−1.68 |
0.885 |
3li–methyl |
−2.11 |
−0.69 |
0.858 |
4li–methyl |
−1.49 |
0.36 |
0.858 |
1li–cyano |
−4.24 |
— |
0.921 |
2li–cyano |
−3.42 |
−2.60 |
0.887 |
3li–cyano |
−2.63 |
−1.06 |
0.857 |
4li–cyano |
−1.89 |
0.33 |
0.858 |
1li–benzoic |
−4.15 |
— |
0.919 |
2li–benzoic |
−3.14 |
−2.13 |
0.886 |
3li–benzoic |
−2.38 |
−0.86 |
0.857 |
4li–benzoic |
−1.67 |
0.44 |
0.857 |
C6H6–Mg2+ complex on the other hand, can accommodate a number of maximum four hydrogen molecules (Fig. 2(a)). The Δ1E(n) values of these complexes depicted in Table 2 also support this observation. There is a negligible amount of energy decrement (∼−0.11 kcal mol−1) on going from 4mg–ben to 5mg–ben. Upon adsorption of up to four H2 molecules on the metal site, the average Mg2+−H2 distance increases in the range of 2.1 Å and 2.2 Å and the residual charge on the Mg2+ decreases (Table 2). On the other hand, addition of fifth hydrogen causes its distance ∼3.8 Å away from the metal center. The average binding energies of H2 to 1mg–ben, 2mg–ben, 3mg–ben, 4mg–ben and 5mg–ben are −13.60 kcal mol−1, −10.99 kcal mol−1, −9.47 kcal mol−1, −7.65 kcal mol−1 and −6.14 kcal mol−1 respectively. The increasing order of metal−H2 distance as well as the decreasing order of residual charge values with the addition of multiple hydrogen molecules signifies that they hold same correlation with the decreasing trend of binding energy values.
 |
| Fig. 2 Optimized geometries of C6H6–Mg2+(H2)n (a) and C6H5X–Mg2+(H2)n [(b)–(d)] complexes where n = 1–5. DMg–H2 represents the average bond length between Mg2+ and H2 molecules. All the bond lengths are given in Å unit. | |
Table 2 B2PLYPD BSSE corrected H2 binding energy (ΔBE) and incremental first-order energy difference [Δ1E(n)] values (kcal mol−1) of Mg2+ decorated systems and charge on the cation (a.u.)
System |
ΔBE |
Δ1E(n) |
Charge on cation |
1mg–ben |
−13.60 |
— |
1.786 |
2mg–ben |
−10.99 |
−8.40 |
1.742 |
3mg–ben |
−9.47 |
−6.43 |
1.701 |
4mg–ben |
−7.65 |
−2.19 |
1.672 |
5mg–ben |
−6.14 |
−0.11 |
1.675 |
1mg–aniline |
−12.27 |
— |
1.764 |
2mg–aniline |
−9.83 |
−7.40 |
1.732 |
3mg–aniline |
−8.60 |
−6.13 |
1.696 |
4mg–aniline |
−6.86 |
−1.65 |
1.671 |
5mg–aniline |
−5.49 |
−0.01 |
1.675 |
1mg–methyl |
−12.91 |
— |
1.789 |
2mg–methyl |
−10.55 |
−8.20 |
1.746 |
3mg–methyl |
−9.10 |
−6.20 |
1.706 |
4mg–methyl |
−7.31 |
−1.93 |
1.678 |
5mg–methyl |
−5.85 |
−0.02 |
1.683 |
1mg–cyano |
−14.00 |
— |
1.793 |
2mg–cyano |
−11.70 |
−9.40 |
1.744 |
3mg–cyano |
−10.17 |
−7.11 |
1.700 |
4mg–cyano |
−8.28 |
−2.61 |
1.670 |
5mg–cyano |
−6.68 |
−0.29 |
1.673 |
3.2 Interaction of hydrogen with C6H5X–M (X = −NH2, –CH3, –CN, –COOH) complexes
In this section we investigated the effect of different electron donating groups (−NH2, −CH3) and electron withdrawing groups (−CN, −COOH) on the hydrogen binding capability of the Li+ decorated substituted benzene rings. Due to extremely high affinity of Mg2+ for the oxide, it is quite impossible to model Mg2+ decorated benzoic acid. Therefore, except the COOH group, we have studied the effect of other three substituents on the C6H5X–Mg2+ complexes. The study of Srinivasu and co-workers on the Na+ and Li+ doped substituted benzene revealed that the interaction of molecular hydrogen with it is always found to be stronger than that of benzene irrespective of the nature of the substituents.27 Both Table 1 and Fig. 1(b)–(e) exemplify the effect of introducing the above mentioned substituents and also the effect of multiple hydrogen molecules adsorbed onto the C6H5X–Li+ complexes. Alike to the benzene system, the maximum number of hydrogen molecules adsorbed is three in all the four substituted benzene systems which is also revealed by their Δ1E(n) values in Table 2. With increase of the number of adsorbed hydrogen molecules, ΔBE as well as the residual charge on the Li+ cation decreases, whereas Li+−H2 distance increases accordingly. The varying trends of these geometrical parameters with the addition of multiple hydrogen molecules imply that these factors correlate themselves well with the decreasing binding energy values. As we move from an electron donating group (−NH2, −CH3) to an electron withdrawing group (−CN, −COOH), the charge on the alkali metal cation increases which in turn helps in increasing the binding energy of H2 to some extent. The average binding energy of hydrogen molecule in the complexes 1li–aniline, 1li–methyl, 1li–benzoic and 1li–cyano are −3.71 kcal mol−1, −3.96 kcal mol−1, −4.15 kcal mol−1 and −4.24 kcal mol−1 while the residual charges on the Li+ are 0.916 a.u., 0.916 a.u., 0.919 a.u. and 0.921 a.u., respectively. The corresponding optimized geometries of these complexes are shown in Fig. 1.
Table 2 summarizes the average binding energy (ΔBE) of hydrogen and the charge on the metal cation of corresponding C6H5X–Mg2+ complexes. The estimated ΔBE values indicate that the electron donating substituents like 1mg–aniline and 1mg–methyl weaken the H2 binding energy considerably relative to the benzene complex 1mg–ben while the electron withdrawing substituent 1mg–cyano slightly strengthens the interaction. The ΔBE values of 1mg–aniline, 1mg–methyl and 1mg–cyano are −12.27 kcal mol−1, −12.91 kcal mol−1 and −14.00 kcal mol−1 and the residual charges on cation are 1.764 a.u., 1.789 a.u. and 1.793 a.u., respectively. The number of maximum hydrogen molecules adsorbed remains the same relative to the C6H6–Mg2+ complex as confirmed by both the optimized geometries (Fig. 2(b)–(d)) and the incremental energy difference values Δ1E(n) (Table 2). Keeping the substituent unaltered and by varying the number of hydrogen molecules from one to five, the order of ΔBE decreases. This change is consistent with the other structural changes mentioned in Fig. 2.
3.3 LMOEDA analysis
The LMOEDA analysis is executed by taking the interaction between the C6H5X−M system and the hydrogen molecules. Table 3 displays the four contributing energy terms of the instantaneous interaction energy (ΔEint) which is outlined in the introduction section. Irrespective of the nature of the substituent, the electrostatic term Eelec is repulsive in nature when one hydrogen molecule gets adsorbed on C6H5X–Li+ complexes while for the multiple hydrogen adsorbed systems it becomes attractive in nature. For all the complexes investigated, the attractive term Epol is higher than the repulsive terms as well as other attractive terms. Therefore, it has the major contribution to the interaction energy. Among the two attractive terms (Epol and Edisp), the Epol provides 82.6% whereas the dispersion term contributes only 17.4% to the total attractive energy of 1li–ben complex. As evident from the Table 3 the extent of interaction energy increases with increase of the number of hydrogen molecules. In case of 2li–ben, 3li–ben and 4li–ben the attractive polarization energy contributes 70.0%, 67.6% and 67.6% to their respective total attraction energy. In these three complexes, the other two attractive terms, i.e., electrostatic energy term and dispersion term varies between 10% to 12% and 20–20.5%, respectively. On going from electron donating group to electron withdrawing group polarization energy does not vary considerably. The value of Epol for 1li–aniline, 1li–methyl, 1li–cyano and 1li–benzoic are −6.91 kcal mol−1, −7.11 kcal mol−1, −7.33 kcal mol−1 and −7.33 kcal mol−1, respectively.
Table 3 ΔEint and its contributions in kcal mol−1 for H2 adsorbed C6H6–Li+ and C6H5X–Li+ complexes
System |
ΔEint |
Eelec |
Eex-rep |
Epol |
Edisp |
1li–ben |
−3.44 |
0.30 |
4.13 |
−6.50 |
−1.37 |
2li–ben |
−5.17 |
−1.63 |
11.05 |
−11.36 |
−3.23 |
3li–ben |
−6.49 |
−2.83 |
16.97 |
−15.87 |
−4.76 |
4li–ben |
−6.97 |
−2.90 |
17.38 |
−16.46 |
−4.99 |
1li–aniline |
−3.33 |
0.66 |
4.35 |
−6.91 |
−1.43 |
2li–aniline |
−5.44 |
−1.98 |
12.37 |
−12.19 |
−3.64 |
3li–aniline |
−6.88 |
−3.28 |
18.34 |
−16.67 |
−5.27 |
4li–aniline |
−7.41 |
−3.68 |
19.67 |
−17.49 |
−5.91 |
1li–methyl |
−3.39 |
0.98 |
4.11 |
−7.11 |
−1.37 |
2li–methyl |
−5.01 |
−1.11 |
11.45 |
−12.00 |
−3.35 |
3li–methyl |
−6.36 |
−2.37 |
17.43 |
−16.47 |
−4.95 |
4li–methyl |
−6.84 |
−2.47 |
17.91 |
−17.08 |
−5.20 |
1li–cyano |
−3.76 |
0.66 |
4.32 |
−7.33 |
−1.41 |
2li–cyano |
−6.03 |
−1.18 |
11.06 |
−12.56 |
−3.35 |
3li–cyano |
−7.63 |
−2.25 |
16.82 |
−17.28 |
−4.92 |
4li–cyano |
−8.20 |
−2.39 |
17.45 |
−18.04 |
−5.22 |
1li–benzoic |
−3.61 |
0.85 |
4.27 |
−7.33 |
−1.40 |
2li–benzoic |
−5.82 |
−1.33 |
11.78 |
−12.77 |
−3.50 |
3li–benzoic |
−7.54 |
−2.64 |
17.98 |
−17.68 |
−5.20 |
4li–benzoic |
−8.32 |
−3.22 |
19.90 |
−18.98 |
−6.02 |
It is noteworthy that for all of the C6H5X–Mg2+ complexes, as described in Table 4, the electrostatic energy term is negative in nature and the absolute value of Epol is the largest one. The contribution of the polarization term to the total attraction energy is about 78.4% for the complex 1mg–ben and the electrostatic energy term has no such significant contribution in this respect. Similar to the Li+ decorated complexes here also the interaction energy increases with increase in the number of hydrogen molecules. The interaction energy values are −11.47 kcal mol−1, −18.48 kcal mol−1, −24.00 kcal mol−1, −26.81 kcal mol−1 and −28.06 kcal mol−1 for the complexes from 1mg–ben to 5mg–ben, respectively. Among the three attractive terms (Eelec, Epol and Edisp), percentage of polarization energy of these complexes varies from 71.5 to 78.4%, the electrostatic energy term varies from 10.7% to 15.7% and the dispersion term is only 10.9–13.1% of the total attraction energy. Similar to the lithium analogues, here also the Epol does not change considerably on going from electron donating to electron withdrawing group. The values of Epol are −14.15 kcal mol−1, −14.87 kcal mol−1 and −15.45 kcal mol−1 in 1li–aniline, 1li–methyl and in 1li–cyano, respectively.
Table 4 ΔEint and its contributions in kcal mol−1 for H2 adsorbed C6H6–Mg2+ and C6H5X–Mg2+ complexes
System |
ΔEint |
Eelec |
Eex-rep |
Epol |
Edisp |
1mg–ben |
−11.47 |
−1.98 |
7.01 |
−14.48 |
−2.02 |
2mg–ben |
−18.48 |
−5.04 |
14.93 |
−24.40 |
−3.97 |
3mg–ben |
−24.00 |
−7.17 |
21.81 |
−32.97 |
−5.67 |
4mg–ben |
−26.81 |
−8.18 |
28.14 |
−39.62 |
−7.15 |
5mg–ben |
−28.06 |
−8.79 |
29.06 |
−40.86 |
−7.47 |
1mg–aniline |
−10.51 |
−1.47 |
7.11 |
−14.15 |
−2.00 |
2mg–aniline |
−16.83 |
−4.56 |
15.29 |
−23.66 |
−3.90 |
3mg–aniline |
−22.38 |
−7.33 |
23.05 |
−32.28 |
−5.82 |
4mg–aniline |
−25.10 |
−8.81 |
29.77 |
−38.53 |
−7.53 |
5mg–aniline |
−26.35 |
−9.46 |
30.75 |
−39.78 |
−7.86 |
1mg–methyl |
−11.10 |
−1.21 |
6.98 |
−14.87 |
−2.00 |
2mg–methyl |
−17.86 |
−4.32 |
15.04 |
−24.63 |
−3.95 |
3mg–methyl |
−23.14 |
−6.57 |
22.17 |
−32.99 |
−5.75 |
4mg–methyl |
−25.93 |
−7.85 |
28.93 |
−39.64 |
−7.37 |
5mg–methyl |
−27.20 |
−8.45 |
29.75 |
−40.80 |
−7.70 |
1mg–cyano |
−11.95 |
−1.67 |
7.25 |
−15.45 |
−2.08 |
2mg–cyano |
−19.88 |
−4.91 |
15.34 |
−26.14 |
−4.17 |
3mg–cyano |
−25.96 |
−7.21 |
22.53 |
−35.26 |
−6.02 |
4mg–cyano |
−29.52 |
−8.16 |
28.68 |
−42.45 |
−7.59 |
5mg–cyano |
−31.07 |
−9.02 |
30.62 |
−44.37 |
−8.30 |
3.4 NBO analysis
The results of NBO analysis reflect a strong charge transfer interaction between the bonding orbital of hydrogen molecule and the antibonding lone pair orbital of lithium cation. Table 5 depicts the amount of charge transfer among the participating NBOs (‘BD’ for 2-centre bond, ‘LP’ for 1-centre valence lone pair). Forward donation from H2 to metal is described by the second-order stabilization energy (ΔECT) which varies between 8.85 kcal mol−1 and 11.72 kcal mol−1 for the C6H6–Li+ complexes. For lithium decorated aniline, methyl benzene, cyano benzene and benzoic acid systems this charge transfer energy value varies between 8.53 kcal mol−1 and 12.09 kcal mol−1. For each of the complexes when the fourth hydrogen molecule gets attached to the system the amount of charge transferred from H2 to metal drastically decreases. The values are 0.35 kcal mol−1, 0.22 kcal mol−1, 0.38 kcal mol−1, 0.50 kcal mol−1 and 0.56 kcal mol−1 for 4li–ben, 4li–aniline, 4li–methyl, 4li–cyano and 4li–benzoic systems respectively. The total pairwise steric exchange energy (ΔEsteric) consists of contribution arising from the interaction among the H2 molecules (ΔEH2–H2) as well as between the Li+ cation and H2 molecule (ΔELi–H2). Upon adsorption of up to three hydrogen molecules over C6H5X–Li+ systems the ΔELi–H2 value varies between 2.6 kcal mol−1 and ∼4.0 kcal mol−1 while ΔEH2–H2 term is in the range of ∼2.0–2.3 kcal mol−1. The absence of pairwise steric exchange energy between Li+ and the fourth H2 molecule in both unsubstituted and substituted benzene systems implies that the fourth hydrogen molecule is oriented away from the system. The value of ΔEsteric considerably increases with increase in the number of hydrogen molecules around the metal cation for all the unsubstituted and substituted benzene systems. The ΔEsteric energy terms being more or less same in both the 3li–ben and 4li–ben complexes indicate that the steric repulsion is relieved in 4li–ben by placing its fourth hydrogen molecule far apart from the system. The same correlation of ΔEsteric energy term is seen in the presence of both electron donating and electron withdrawing groups.
Table 5 Second order energy correction (ΔECT) and pairwise steric exchange energy (ΔEsteric) associated with H2 to Li+ charge transfer interaction in Li+ decorated complexes
System |
Type of interaction |
ΔECT |
ΔEsteric |
1li–ben |
 |
8.85 |
3.95 |
2li–ben |
 |
10.63 |
9.03 |
 |
10.63 |
3li–ben |
 |
11.72 |
15.50 |
 |
11.71 |
 |
11.72 |
4li–ben |
 |
11.50 |
15.91 |
 |
11.39 |
 |
0.35 |
 |
11.45 |
1li–aniline |
 |
8.53 |
3.95 |
2li–aniline |
 |
10.91 |
8.58 |
 |
10.26 |
3li–aniline |
 |
11.88 |
14.46 |
 |
11.37 |
 |
11.86 |
4li–aniline |
 |
11.70 |
15.42 |
 |
11.15 |
 |
0.22 |
 |
11.68 |
1li–methyl |
 |
8.90 |
3.69 |
2li–methyl |
 |
10.78 |
8.84 |
 |
10.42 |
3li–methyl |
 |
11.75 |
14.78 |
 |
11.73 |
 |
11.36 |
4li–methyl |
 |
11.48 |
15.49 |
 |
0.38 |
 |
11.07 |
 |
11.40 |
1li–cyano |
 |
8.91 |
3.87 |
2li–cyano |
 |
10.85 |
8.74 |
 |
10.87 |
3li–cyano |
 |
12.01 |
14.90 |
 |
11.98 |
 |
12.09 |
4li–cyano |
 |
11.65 |
15.45 |
 |
11.71 |
 |
11.70 |
 |
0.50 |
1li–benzoic |
 |
8.87 |
3.97 |
2li–benzoic |
 |
10.51 |
8.62 |
 |
10.84 |
3li–benzoic |
 |
12.08 |
14.55 |
 |
11.94 |
 |
11.67 |
4li–benzoic |
 |
11.71 |
16.09 |
 |
11.76 |
 |
11.38 |
 |
0.56 |
For C6H5X–Mg2+ complexes the amount of charge transferred from H2 bonding orbital to metal antibonding orbital is found to be much higher than the analogous Li+ decorated systems indicating stronger interaction in case of Mg2+ decorated complexes. The detailed NBO analysis of the participating donor (H2) and acceptor (Mg2+) NBOs in charge transfer interaction are illustrated in Table 6. The ΔECT values fluctuate in between 13.88 kcal mol−1 and 18.43 kcal mol−1 for all the unsubstituted and substituted Mg2+ containing systems. Due to the extremely larger distance of the fifth hydrogen molecule from the metal center, the charge transfer interaction drastically decreases. It is interesting to note that similar to the charge transfer interaction, ΔEsteric values in Mg2+ containing systems (Table 6) are much higher than the Li+ containing equivalent systems. Parallel to the Li+ decorated systems the ΔEsteric energy term increases with increase in the number of hydrogen molecules. The pairwise steric exchange energy terms between Mg–H2 (ΔEMg–H2) and H2–H2(ΔEH2–H2) varies between 3.2–6.7 kcal mol−1 and 0.1–3.4 kcal mol−1 respectively up to the addition of fourth hydrogen molecule on C6H6–Mg2+ and C6H5X–Mg2+ complexes. But as the number of attached H2 molecules changes from four to five the ΔEMg–H2 term between Mg2+ and the fifth hydrogen molecule vanishes indicating its inability to be remained cohesively attached to the Mg2+ cation.
Table 6 Second order energy correction (ΔECT) and pairwise steric exchange energy (ΔEsteric) associated with H2 to Mg2+ charge transfer interaction in Mg2+ decorated complexes
System |
Type of interaction |
ΔECT |
ΔEsteric |
1mg–ben |
 |
16.73 |
6.73 |
2mg–ben |
 |
17.81 |
13.22 |
 |
17.50 |
3mg–ben |
 |
17.78 |
21.11 |
 |
17.78 |
 |
17.78 |
4mg–ben |
 |
16.46 |
29.34 |
 |
16.45 |
 |
16.44 |
 |
16.47 |
5mg–ben |
 |
16.68 |
29.57 |
 |
15.68 |
 |
15.47 |
 |
16.78 |
 |
0.26 |
1mg–aniline |
 |
15.52 |
6.29 |
2mg–aniline |
 |
16.85 |
12.77 |
 |
16.25 |
3mg–aniline |
 |
17.04 |
15.82 |
 |
17.21 |
 |
16.93 |
4mg–aniline |
 |
14.52 |
26.82 |
 |
16.88 |
 |
16.63 |
 |
14.49 |
5mg–aniline |
 |
13.88 |
28.70 |
 |
16.75 |
 |
16.52 |
 |
14.11 |
 |
0.25 |
1mg–methyl |
 |
16.55 |
6.38 |
2mg–methyl |
 |
17.38 |
12.96 |
 |
16.83 |
3mg–methyl |
 |
17.53 |
20.49 |
 |
17.40 |
 |
16.98 |
4mg–methyl |
 |
15.70 |
27.57 |
 |
16.70 |
 |
16.55 |
 |
14.89 |
5mg–methyl |
 |
14.61 |
29.74 |
 |
16.77 |
 |
16.57 |
 |
14.15 |
 |
0.26 |
1mg–cyano |
 |
16.87 |
6.63 |
2mg–cyano |
 |
17.95 |
13.46 |
 |
17.73 |
3mg–cyano |
 |
17.86 |
20.87 |
 |
18.43 |
 |
18.22 |
4mg–cyano |
 |
16.25 |
28.49 |
 |
17.05 |
 |
17.13 |
 |
16.83 |
5mg–cyano |
 |
16.55 |
31.01 |
 |
16.84 |
 |
16.40 |
 |
16.74 |
 |
0.36 |
3.5 AIM analysis
AIM analysis is carried out in order to get further insight into the H2−metal cation interaction of the Li+ and Mg2+ decorated C6H5X systems. Table S3 and S4 in the ESI† depict the electron density and the Laplacian of charge density (∇2ρ) values at the respective hydrogen-metal cation BCP of each complex. The molecular graphs of 1li–ben and 1mg–ben are illustrated in Fig. 3. The charge shift type interaction between H2 and metal center is realized by small positive value of ∇2ρ at the respective BCP. The electron density value along the (H2)–Li+ BCPs and the ∇2ρ value in all the C6H6–Li+ and C6H5X–Li+ complexes lies within the range of 0.0120–0.0156 and 0.0664–0.0828, specifying that this interaction is stronger than the van der Waals interaction. In case of Mg2+ complexes, the value of electron density at the (H2)–Mg2+ BCPs varies between 0.0125 and 0.0225 while ∇2ρ value lies in between 0.0538 and 0.1048 indicating stronger charge shift type interaction among H2–Mg2+ complexes compared to the Li+-analogue.
 |
| Fig. 3 Electron density molecular graphs of (I) 1li–ben and (II) 1mg–ben. The green dot localized between the H2 and the metal cation indicates the location of bond critical point of the electron density. | |
4. Conclusions
In this work we have systematically explored and analyzed the adsorption and thereby the binding affinities of hydrogen molecules on alkali (Li+) and alkaline earth metal (Mg2+) decorated benzene. Between these two metal cations the later one is found to bind greater number of hydrogen molecules than the former one. The binding energy of hydrogen molecule is much higher in C6H6–Mg2+ complex than the C6H6–Li+ complex. Addition of electron donating and electron withdrawing groups to the benzene moiety is also found to affect the hydrogen adsorption. The topological analysis of the electron density at the (H2)−metal cation BCP (Li+, Mg2+) shows closed shell interaction. For this particular type of bond the contribution of polarization term is much more prominent than other attractive and repulsive terms of the instantaneous interaction energy, evident from the LMOEDA analysis. There is considerable amount of charge transfer from the bonding orbital of hydrogen molecule to the antibonding lone pair orbital of metal cation, found from the NBO analysis. In case of C6H5X–Mg2+ system, the amount of charge transfer is much higher than that found in C6H5X–Li+ system. This significant contribution of charge transfer in the interaction between H2 molecule and the metal cation is also reflected in polarization term value in LMOEDA analysis. Substantial steric exchange interactions are also realized among the bound H2 molecules and the metal center. Here also, these interactions are more pronounced in Mg2+ complexes compared to the analogous Li+ complexes. A strong correlation is found to exist between the charge transfer interaction and the total binding energy of the complex. In addition the contribution of polarization towards the binding energy is significantly high. The polarization and charge transfer interactions together help in stabilizing the system while the pairwise steric exchange interaction on the other hand, destabilizes the system, resulting decrease in binding energy of H2. Our work is expected to provide crucial insight into the interaction between metal cation and hydrogen molecules and thereby provides fundamental basis of designing carbon nanomaterials, metal organic frameworks etc. for efficient hydrogen storage.
Acknowledgements
TB, TD and KS are grateful to the Council of Scientific and Industrial Research (CSIR), Government of India, for providing them research fellowships. AKD is grateful to the Department of Science and Technology (DST), Govt. of India, for a research grant under scheme number: SB/S1/PC-79/2012.
References
- R. C. Lochan and M. Head-Gordon, Phys. Chem. Chem. Phys., 2006, 8, 1357 RSC.
- P. Jena, J. Phys. Chem. Lett., 2011, 2, 206 CrossRef CAS.
- S. S. Han, J. L. Mendoza-Cortés and W. A. Goddard III, Chem. Soc. Rev., 2009, 38, 1460 RSC.
- M. Eddaoudi, H. Li, T. Reineke, M. Fehr, D. Kelley, T. L. Groy and O. M. Yaghi, Top. Catal., 1999, 9, 105 CrossRef CAS.
- N. L. Rosi, J. Eckert, M. Eddaoudi, D. T. Vodak, J. Kim, M. O’Keeffe and O. M. Yaghi, Science, 2003, 300, 1127 CrossRef CAS PubMed.
- J. L. C. Rowsell, A. R. Millward, K. S. Park and O. M. Yaghi, J. Am. Chem. Soc., 2004, 126, 5666 CrossRef CAS PubMed.
- J. L. C. Rowsell and O. M. Yaghi, Angew. Chem., Int. Ed., 2005, 44, 4670 CrossRef CAS PubMed.
- J. L. C. Rowsell and O. M. Yaghi, J. Am. Chem. Soc., 2006, 128, 1304 CrossRef CAS PubMed.
- S. S. Han, W.-Q. Deng and W. A. Goddard III, Angew. Chem., Int. Ed., 2007, 46, 6289 CrossRef CAS PubMed.
- A. Kuc, T. Heine, G. Seifert and H. A. Duarte, Chem.–Eur. J., 2008, 14, 6597 CrossRef CAS PubMed.
- G. Garberoglio, A. I. Skoulidas and J. K. Johnson, J. Phys. Chem. B, 2005, 109, 13094 CrossRef CAS PubMed.
- J. L. C. Rowsell, J. Eckert and O. M. Yaghi, J. Am. Chem. Soc., 2005, 127, 14904 CrossRef CAS PubMed.
- Y. Kubota, M. Takata, R. Matsuda, R. Kitaura, S. Kitagawa, K. Kato, M. Sakata and T. C. Kobayashi, Angew. Chem., Int. Ed., 2005, 44, 920 CrossRef CAS PubMed.
- O. Hübner, A. Glöss, M. Fichtner and W. Klopper, J. Phys. Chem. A, 2004, 108, 3019 CrossRef.
- T. Sagara, J. Klassen, J. Ortony and E. Ganz, J. Chem. Phys., 2005, 123, 014701 CrossRef PubMed.
- Q. Yang and C. Zhong, J. Phys. Chem. B, 2005, 109, 11862 CrossRef CAS PubMed.
- S. Hamel and M. Côté, J. Chem. Phys., 2004, 121, 12618 CrossRef CAS PubMed.
- S. S. Han and W. A. Goddard III, J. Am. Chem. Soc., 2007, 129, 8422 CrossRef CAS PubMed.
- R. C. Lochan, R. Z. Khaliullin and M. Head-Gordon, Inorg. Chem., 2008, 47, 4032 CrossRef CAS PubMed.
- A. Blomqvist, C. M. Araújo, P. Srepusharawoot and R. Ahuja, Proc. Natl. Acad. Sci. U. S. A., 2007, 104, 20173 CrossRef CAS PubMed.
- A. Mavrandonakis, E. Tylianakis, A. K. Stubos and G. E. Froudakis, J. Phys. Chem. C, 2008, 112, 7290 CAS.
- P. Dalach, H. Frost, R. Q. Snurr and D. E. Ellis, J. Phys. Chem. C, 2008, 112, 9278 CAS.
- S. J. Kolmann, B. Chan and M. J. T. Jordan, Chem. Phys. Lett., 2008, 467, 126 CrossRef CAS PubMed.
- K. R. S. Chandrakumar and S. K. Ghosh, Chem. Phys. Lett., 2007, 447, 208 CrossRef CAS PubMed.
- S. Pakhira, C. Sahu and K. Sen, Struct. Chem., 2013, 24, 549 CrossRef CAS.
- K. Sen, S. Pakhira and C. Sahu, Mol. Phys., 2014, 112, 182 CrossRef CAS PubMed.
- K. Srinivasu, K. R. S. Chandrakumar and S. K. Ghosh, Phys. Chem. Chem. Phys., 2008, 10, 5832 RSC.
- K. Srinivasu, K. R. S. Chandrakumar and S. K. Ghosh, ChemPhysChem, 2009, 10, 427 CrossRef CAS PubMed.
- T. A. Maark and S. Pal, Int. J. Hydrogen Energy, 2010, 35, 12846 CrossRef CAS PubMed.
- I. V. Bodrenko, A. V. Avdeenkov, D. G. Bessarabov, A. V. Bibikov, A. V. Nikolaev, M. D. Taran and E. V. Tkalya, J. Phys. Chem. B, 2012, 116, 25286 CAS.
- H.-W. Huang, H.-Ju. Hsieh, I. Lin, Y.-J. Tong and H.-T. Chen, J. Phys. Chem. B, 2015, 119, 7662 CAS.
- M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci and G. A. Petersson, et al., Gaussian 09, Revision B.01, Gaussian, Inc., Wallingford, CT, 2010 Search PubMed.
- A. D. Becke, J. Chem. Phys., 1993, 98, 5648 CrossRef CAS PubMed.
- C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785 CrossRef CAS.
- T. Schwabe and S. Grimme, Phys. Chem. Chem. Phys., 2007, 9, 3397 RSC.
- F. Neese, T. Schwabe and S. Grimme, J. Chem. Phys., 2007, 126, 124115 CrossRef PubMed.
- C. Møller and M. S. Plesset, Phys. Rev., 1934, 46, 618 CrossRef.
- M. Head-Gordon, J. A. Pople and M. J. Frisch, Chem. Phys. Lett., 1988, 153, 503 CrossRef CAS.
- M. J. Frisch, M. Head-Gordon and J. A. Pople, Chem. Phys. Lett., 1990, 166, 275 CrossRef CAS.
-
(a) R. Ditchfield, W. J. Hehre and J. A. Pople, J. Chem. Phys., 1971, 54, 724 CrossRef CAS PubMed;
(b) W. J. Hehre, R. Ditchfield and J. A. Pople, J. Chem. Phys., 1972, 56, 2257 CrossRef CAS PubMed;
(c) P. C. Hariharan and J. A. Pople, Mol. Phys., 1974, 27, 209 CrossRef CAS PubMed;
(d) M. S. Gordon, Chem. Phys. Lett., 1980, 76, 163 CrossRef CAS;
(e) P. C. Hariharan and J. A. Pople, Theor. Chim. Acta, 1973, 28, 213 CrossRef CAS;
(f) J.-P. Blaudeau, M. P. McGrath, L. A. Curtiss and L. Radom, J. Chem. Phys., 1997, 107, 5016 CrossRef CAS PubMed;
(g) M. M. Francl, W. J. Pietro, W. J. Hehre, J. S. Binkley, D. J. DeFrees, J. A. Pople and M. S. Gordon, J. Chem. Phys., 1982, 77, 3654 CrossRef CAS PubMed;
(h) R. C. Binning Jr and L. A. Curtiss, J. Comput. Chem., 1990, 11, 1206 CrossRef PubMed;
(i) V. A. Rassolov, J. A. Pople, M. A. Ratner and T. L. Windus, J. Chem. Phys., 1998, 109, 1223 CrossRef CAS PubMed;
(j) V. A. Rassolov, M. A. Ratner, J. A. Pople, P. C. Redfern and L. A. Curtiss, J. Comput. Chem., 2001, 22, 976 CrossRef CAS PubMed.
- R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem. Phys., 1980, 72, 650 CrossRef CAS PubMed.
- A. D. McLean and G. S. Chandler, J. Chem. Phys., 1980, 72, 5639 CrossRef CAS PubMed.
- A. Schäfer, C. Huber and R. Ahlrichs, J. Chem. Phys., 1994, 100, 5829 CrossRef PubMed.
- S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553 CrossRef CAS PubMed.
- T. A. Keith, AIMAll, 14.06.21, TK Gristmill Software, Overland Park, KS, 2013, http://aim.tkgristmill.com Search PubMed.
- E. D. Glendening, J. K. Badenhoop, A. E. Reed, J. E. Carpenter, J. A. Bohmann, C. M. Morales, C. R. Landis and F. Weinhold, NBO6.0, Theoretical Chemistry Institute, University of Wisconsin, Madison, WI, USA, 2013 Search PubMed.
- F. Weinhold and C. R. Landis, Discovering Chemistry with Natural Bond Orbitals, John Wiley & Sons, Inc., Hoboken, New Jersey, 2012 Search PubMed.
- A. E. Reed, R. B. Weinstock and F. Weinhold, J. Chem. Phys., 1985, 83, 735 CrossRef CAS PubMed.
- A. E. Reed, L. A. Curtiss and F. Weinhold, Chem. Rev., 1988, 88, 899 CrossRef CAS.
- J. K. Badenhoop and F. Weinhold, J. Chem. Phys., 1997, 107, 5406 CrossRef CAS PubMed.
- J. K. Badenhoop and F. Weinhold, J. Chem. Phys., 1997, 107, 5422 CrossRef CAS PubMed.
- J. K. Badenhoop and F. Weinhold, Int. J. Quantum Chem., 1999, 72, 269 CrossRef CAS.
- P. Su and H. Li, J. Chem. Phys., 2009, 131, 014102 CrossRef PubMed.
- 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 Jr, J. Comput. Chem., 1993, 14, 1347 CrossRef CAS PubMed.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra09884j |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.