The effect of electric field on hydrogen storage for B/N-codoped graphyne

Lihong Zhanga, Ning Wanga, Shengli Zhangb and Shiping Huang*a
aState Key Laboratory of Organic-Inorganic Composites, Beijing University of Chemical Technology, Beijing 100029, China. E-mail: huangsp@mail.buct.edu.cn; Fax: +86-10-64427616
bSchool of Materials Science and Engineering, Nanjing University of Science and Technology, Nanjing, Jiangsu, 210094, China

Received 29th July 2014 , Accepted 8th October 2014

First published on 8th October 2014


Abstract

Hydrogen adsorption on a B/C/N sheet under different external electric fields is investigated by first-principles calculations. Through the analyses of structural properties of the B/C/N system, we find that NbBf, BbNo, BaNe, and NaBg are more probable to be synthesized. Through molecular dynamics calculations, it was found that the structures for B/N doped graphyne are stable. For NbBf, BbNo, BaNe, and NaBg, the most stable positions for hydrogen adsorption are the H1 sites. For a single H2 adsorbed on a B/C/N sheet, the adsorption energy increases greatly as the electric field increases, and the maximum adsorption energy is 0.506 eV when the electric field is 0.035 a.u. It is also found that the adsorption energy of H2 adsorbed on NbBf under electric field increases faster than H2 adsorbed on other sheets. The interaction between H2 molecule and B/C/N sheet is the Kubas interaction under an external electric field.


1. Introduction

Due to rising standards of living and growing populations, energy consumption is expected to increase dramatically. Hydrogen is very attractive as a clean energy source because of its efficiency, abundance and environmental friendliness.1–4 However, how to store hydrogen safely and transport efficiently is a crucial problem.3,5

Among various hydrogen storage materials, carbon-based nanostructures, including nanotubes,6 fullerenes7 and graphenes, have attracted considerable attention because of their remarkable properties, such as good reversibility, high capacity, and fast kinetics.8 However, hydrogen adsorption on carbon materials is limited by the van der Waals (vdW) interaction, which is too weak to store hydrogen in moderate conditions. Recent studies showed that nanostructures composed of light elements such as B and N offer many advantages to store hydrogen.9–11 For example, the BN layer is slightly more resistant to oxidation than graphene, and is thus more suitable for application at room temperature, where graphene would be oxidized.12 Furthermore, similar to carbon nanotubes, BN nanotubes are also regarded as possible hydrogen storage media. It has been experimentally found that the BN nanotubes can store as much as 2.6 wt% of hydrogen at 10 MPa.13 Collapsed BN nanotubes exhibit a higher hydrogen storage capacity with 4.2 wt% of hydrogen.14 Moreover, Zhou et al. pointed out that an effective method to promote hydrogen adsorption is to add an external electric field.9 The new concept is based on the fact that the polarization of H2 molecule caused by the electric field associated with point ions allows exposed metal cations to store hydrogen in a quasi-molecular form.15,16 Subsequently, Liu et al. put forward that electric field can induce a reversible switch for hydrogen adsorption and desorption based on Li-doped carbon nanotube and Li-doped graphene.17,18 Enhanced hydrogen adsorption on carbonaceous sorbent under an electric field has been demonstrated in experiment by Shi et al.19

Currently, there are some theoretical reports examining graphyne. Graphyne is a new carbon-based form that consists of planar carbon sheets containing sp and sp2 bonds,20–23 which can be regarded as the big hexagonal rings joined together by the acetylenic linkages (C–C[triple bond, length as m-dash]C–C) rather than the C–C[double bond, length as m-dash]C–C in graphene. Experimentally, although large graphyne structures have not been synthesized yet, graphdiyne (expanded graphynes) films and graphdiyne tubes have been already obtained.24

In this study, we present theoretical study of the structure of 2D graphyne and its structural analogs involving BN rings (BN-yne). The latter is composed of BN hexagonal rings linked by C-chains. In this work, we investigate the stability of B/C/N systems, structural characteristic and electronic properties of hydrogen adsorption on a B/C/N sheet using density functional theory (DFT) calculations. First, we study the possible synthesis for B/C/N system based on the formation energy of B/C/N sheet. Then, we investigate stability of the B/C/N sheets through a dynamic calculation. Finally, we focus on the structural and electronic characteristics of hydrogen adsorption on the B/C/N sheet under different electric fields.

2. Computational methods

All geometry optimizations are carried out using DFT calculations within the local density approximation (LDA) as implemented in the DMol3 package.25,26 We applied the Perdew and Wang correlation functional (PWC)27 and relaxed the geometries. Previous studies have demonstrated that LDA can predict the physisorption energies of H2 on the surface of graphite and carbon nanotubes accurately28–30 and can be suitable for charged carbon nanostructures with electric fields.17,31 Full structural optimizations were obtained using a convergence tolerance energy of 1 × 10−5 Hartree, a maximum force of 2 × 10−3 Hartree per Å, and a maximum displacement of 5 × 10−3 Å. Moreover, a double numerical-polarized basis set (DNP) is employed. The DNP basis set included a double quality basis set with a p-type polarization function added to hydrogen and d-type polarization functions added to heavier atoms. Self-consistent field (SCF) calculations were performed with a convergence of 1 × 10−6 Hartree on the total energy. The direct inversion of iterative subspace (DIIS) approach was used to accelerate SCF convergence. A (2 × 2) B/N-codoped graphyne supercell was established with lattice parameters of a = b = 13.7887 Å, c = 28.7497 Å, α = β = 90°, γ = 120°. The vacuum space of 20 Å was used in the direction perpendicular to the B/C/N sheet in order to avoid the interaction between neighboring layers. The electric field was applied in the z direction, as shown in Fig. 1. The total density of states (DOS) and partial density of states (PDOS) of B/N-doped graphyne with and without electric field in the super cells were calculated with a k-point of (12 × 12 × 1). The Ortmann, Bechstedt and Schmidt (OBS) method was employed to correct the dispersion effect,32 and the effect of basis set superposition error (BSSE) was considered with the use of counterpoise (CP) correction.33
image file: c4ra07761j-f1.tif
Fig. 1 Single H2 adsorbed on the NbBf sheet, vertical electric field is applied in the z direction.

3. Results and discussion

The optimized structure of pristine graphyne is shown in Fig. 2. In pristine graphyne, the bonds of C atoms have three types, i.e., C–C single bond, C[double bond, length as m-dash]C double bond, and C[triple bond, length as m-dash]C triple bond (marked as number 1, 2, and 3 in Fig. 2, respectively). The bond lengths of C–C, C[double bond, length as m-dash]C, and C[triple bond, length as m-dash]C are 1.410 Å, 1.430 Å, and 1.227 Å, respectively. The calculated bond lengths are in good agreement with previous reports.28,34 For B- or N-doped graphyne, there are two inequivalent sites for one B or N atom randomly substituting one C atom in graphyne, labeled as a and b in Fig. 2. After the structure optimization, the geometries are obtained and their structural parameters are presented in Table 1. The formation energy (EF) can be defined as follows:
 
EF = EDG + nECEPG − ∑EDopant (1)
where EDG is the total energy of single doped or codoped graphyne, EPG is the total energy of pristine graphyne, and ∑EDopant is the total energy of dopant B, N, or B/N pairs. EC is the energy of one carbon atom in pristine graphyne, and n is the number of carbon atoms substituted by dopants. In the case of B doping at a site, the bond lengths of three B–C bonds are 1.528, 1.528, and 1.492 Å, longer than corresponding dC–C in pristine graphyne. For N doping, the bond lengths of the three N–C bonds are 1.423, 1.423, and 1.357 Å, which are equal or shorter than corresponding dC–C in pristine graphyne. In the case of B (N) doping at the b site, the bond lengths of B–C are 1.485 and 1.344 Å, longer than the corresponding dC–C in pristine graphyne; the bond lengths of N–C are 1.366 and 1.187 Å, shorter than the corresponding dC–C in pristine graphyne. In a word, the bond lengths of B–C (N–C) are longer (shorter) than the bond lengths of corresponding C–C bonds in the case of both B(N) doping at the a site and b site, due to different atomic radius of B and N atoms. It can be seen from the formation energy (in Table 1) that B or N doped at the a site and b site of graphyne are of great possibility to be synthesized, especially for N-doped graphyne.

image file: c4ra07761j-f2.tif
Fig. 2 The optimized structure of pristine graphyne. Numbers 1, 2, and 3 donate C–C, C[double bond, length as m-dash]C, and C[triple bond, length as m-dash]C bonds, respectively; Letters a, b, …, p represent 16 substitution positions for B(N) doping and B/N codoping.
Table 1 The bond lengths of C–B (C–N), and the formation energies EF for B or N doping
  B doping N doping
a site b site a site b site
dC–B(N) (Å) 1.528, 1.492 1.485, 1.344 1.423, 1.357 1.366, 1.187
EF (eV) 0.23 0.19 0.15 0.19


As for B/N codoping, we considered sixteen substitution sites. We divided the codoping configurations into two groups, as shown in Fig. 2. One is ax or xa (x = b, c, …, p), which represents that B (or N) atom substitutes C atom in a site and N (or B) atom substitutes C atom in b, c, d, e, f, g, h, i, j, k, l, m, n, o, and p sites; the other is by or yb (y = a, c, d, …, p), which indicates that B (or N) atom substitutes C atom in b site and N (or B) atom substitutes C atom in a, c, d, e, f, g, h, i, j, k, l, m, n, o, and p sites. As given in Table 2, the dC–B and dC–N for B/N codoping are almost the same as those of single B and N doping, indicating that the configuration maintains the planar structure.35 The formation energies changed from 0.190 to 0.265 eV for ax and from 0.150 to 0.233 eV for xa, indicating that N doping at a site is easier to synthesize than B doping at a site; the formation energies vary from 0.141 to 0.271 eV for bx and from 0.188 to 0.271 eV for xb, revealing that B doping at the b site is easier to synthesize than N doping at b site.

Table 2 The C-dopants distance (dC–B and dC–N), and the formation energy (EF). When B or N atom bond with several C atoms, the dC–B or dC–N indicate the shortest bond length. (The definitions of all symbols can be seen in Fig. 2)
ax ab ac ad ae af ag ah ai aj ak al am an ao ap
dC–B (Å) 1.521 1.512 1.504 1.497 1.490 1.490 1.488 1.491 1.489 1.488 1.491 1.492 1.493 1.492 1.494
dC–N (Å) 1.179 1.172 1.344 1.355 1.187 1.186 1.355 1.348 1.185 1.179 1.352 1.406 1.405 1.415 1.405
EF (eV) 0.265 0.237 0.196 0.190 0.228 0.227 0.193 0.191 0.240 0.235 0.233 0.233 0.196 0.200 0.197

xa ba ca da ea fa ga ha ia ja ka la ma na oa pa
dC–B (Å) 1.348 1.342 1.504 1.490 1.346 1.345 1.489 1.495 1.346 1.343 1.490 1.526 1.521 1.526 1.520
dC–N (Å) 1.415 1.315 1.344 1.348 1.354 1.353 1.355 1.354 1.352 1.355 1.352 1.353 1.355 1.350 1.354
EF (eV) 0.189 0.156 0.196 0.191 0.152 0.150 0.193 0.190 0.162 0.153 0.232 0.233 0.196 0.200 0.196

by ba bc bd be bf bg bh bi bj bk bl bm bn bo bp
dC–B (Å) 1.348 1.513 1.342 1.347 1.344 1.344 1.345 1.346 1.342 1.344 1.343 1.343 1.348 1.351 1.348
dC–N (Å) 1.415 1.337 1.315 1.352 1.185 1.187 1.352 1.354 1.187 1.187 1.355 1.411 1.416 1.425 1.417
EF (eV) 0.189 0.271 0.156 0.162 0.190 0.192 0.150 0.152 0.188 0.189 0.153 0.153 0.153 0.141 0.153

yb ab cb db eb fb gb hb ib jb kb lb mb nb ob pb
dC–B (Å) 1.521 1.512 1.514 1.488 1.343 1.344 1.490 1.490 1.344 1.344 1.487 1.527 1.532 1.531 1.530
dC–N (Å) 1.179 1.337 1.173 1.184 1.187 1.187 1.186 1.188 1.185 1.187 1.180 1.179 1.182 1.175 1.182
EF (eV) 0.265 0.271 0.237 0.240 0.188 0.192 0.227 0.228 0.190 0.189 0.234 0.235 0.230 0.222 0.230


To check the stability of B/C/N system, we use the ab initio molecular dynamics (MD) simulation for B/C/N sheets. In the simulations, the temperature gradually increases from 1 to 900 K in 1000 time steps, and the time step is 1 fs. It is found that the structure does not undergo an evident deformation even at the temperature of 900 K.

Based on the formation energy, we choose four B/N-codoping configurations to investigate one H2 adsorbed on different sites of B/C/N sheets. There are three types of adsorption sites on B/C/N sheets: hollow site (H), top site (T), and bridge site (B). For NbBf, there are two hollow sites, five top sites, and five bridge sites, as shown in Fig. 3(a). The calculated results of adsorption energies are summarized in Fig. 3(b). The adsorption energy of H2 on B/C/N sheets is defined as follows:

 
image file: c4ra07761j-t1.tif(2)
where EH2–B/C/N is the total energy of H2 adsorbed on the B/C/N sheets, EB/C/N is the total energy of B/C/N sheets, and EH2 is the total energy of free-standing H2 molecule. F indicates the intensity of electric field; when F = 0, eqn (2) refers to the Ead without the electric field; when F ≠ 0, eqn (2) refers to the Ead under the electric field. Dispersion and BSSE corrections are introduced to the calculation, and these correction data are listed in the ESI Table S1. It can be seen that the largest adsorption energy of H2 adsorbed on NbBf is 0.407 eV at the H1 site. Then, we calculate the adsorption energies of H2 adsorbed on different sites of BbNo, BaNe, and NaBg sheets, and the most favorable positions are the H1 site with the largest adsorption energies of 0.400, 0.357, and 0.341 eV, respectively.


image file: c4ra07761j-f3.tif
Fig. 3 (a) Different adsorption sites on a NbBf sheet, including two hollow sites (H), five top sites (T), and five bridge sites (B). The gray, blue, and yellow balls are carbon, nitrogen, and boron atoms, respectively. (b) The adsorption energy of H2 adsorbed on different sites of NbBf sheet with DFT-D and BSSE correction.

Then we studied the relative stability of different configurations for H2 molecule on the B/C/N sheet in an external electric field vertical to the sheet. Similar to the field-free case, three types of adsorption sites (top site, bridge site, and hollow site) were considered. For each site, various initial orientations of the H2 molecule were studied. After optimization, it was seen that the angle between H2 molecule and B/N-codoped graphyne plane increases with the increase of electric field, as shown in Fig. 4. The H2 molecule prefers to align along the z direction, i.e. parallel to the electric field, when the intensity of the electric field is strong enough, for example, 0.035 a.u. in the case of single H2 absorbed on the NbBf sheet. For NbBf, BbNo, BaNe, and NaBg, the preferable positions are H1 sites; for Na and Nb, the preferable positions are T4 and H2 sites.


image file: c4ra07761j-f4.tif
Fig. 4 The configurations of single H2 absorbed on the NbBf sheet, with the increase of electric field. The vertical coordinate represents the angle between H2 molecule and B/N-codoped graphyne plane.

We also investigated the effect of the electric field on hydrogen adsorption. As shown in Fig. 5, the hydrogen adsorption energy and the H–H bond length are plotted as a function of the magnitude of the electric field. Both dispersion and BSSE corrections are considered in the calculation, as listed in the ESI Table S2. It can be found that the adsorption energy increases dramatically under the electric field, indicating that the electric field can easily modulate the adsorption energy of H2 adsorbed on the B/C/N sheet. It is also found that the adsorption energy of H2 adsorbed on NbBf under an electric field increases faster than H2 adsorbed on other sheets. Through Mulliken charge analysis in Table 3, we find that the charge of the B/C/N sheets increases with increasing the electric field strength, whereas the increase for electric quantity of NbBf is faster than other sheet. This may be considered for why the adsorption energy of H2 adsorbed on NbBf increases faster than H2 adsorbed on other sheets. Due to the polarization interaction induced by the electric field and the polar bonds on the B/C/N sheets, the H–H bond length increased greatly with increasing electric field strength. When the electric field reached 0.038 a.u., the H–H bond length became 0.870 Å for H2 adsorbed on a NbBf sheet, and an elongation of nearly 13% occurred. It has been reported that the Kubas interaction between a transition metal atom and H2 can elongate the bond length by 10–25%.36,37 The interaction energy between H2 molecule and B/N-codoped graphyne are in the range of 0.243–0.407 eV in absence of an electric field. However, the interaction energy was in the range of 0.407–0.506 eV in presence of an electric field. These results illustrate the interaction between H2 molecule and B/N-codoped graphyne is the Kubas interaction. From the analysis of molecular orbitals, it was found that the bonding orbitals are mainly between the π orbital of the B/N-codoped graphyne and the σ* anti-bonding orbital of the hydrogen molecule, and this interaction can be explained by the Kubas interaction.35 When the electric field intensity is 0.039 a.u., we found that the H2 molecule was dissociated. Therefore, the molecular hydrogen adsorption can be achieved only under a certain electric field.


image file: c4ra07761j-f5.tif
Fig. 5 (a) The adsorption energy and (b) the H–H bond length as a function of the magnitude of the electric field.
Table 3 The change of the Mulliken charge for the B/C/N sheets under the electric field
E-field (a.u.) The Mulliken charge for B/C/N sheets (|e|)
BbNo BaNe NaBg NbBf
0 0.012 0.010 0.007 0.010
0.02 0.049 0.026 0.044 0.048
0.035 0.143 0.077 0.143 0.155


To investigate the effect of the electric field on the electronic structure, we analyzed the partial density of states (PDOS) for a single H2 adsorbed on B/C/N sheets under different electric fields. Fig. 6 shows the PDOS of the hydrogen atom, boron atom and nitrogen atom for H2 adsorbed on NbBf. We found that the H-s orbital moved toward the Fermi level as the electric field strength was increased. This indicates that the H2 elongated by the electric field was less stable than the free-H2 molecule.


image file: c4ra07761j-f6.tif
Fig. 6 The partial density of states for one single H2 adsorbed on NbBf under different electric fields. (a) The electric field is 0.000 a.u. (b) The electric field is 0.020 a.u. (c) The electric field is 0.038 a.u. The Fermi energy is set to zero.

4. Conclusions

DFT calculations were carried out to explore the B/N-codoped graphyne for hydrogen storage. Through the analyses of the structure for B/N-doped graphyne, it is found that NbBf, BbNo, BaNe, and NaBg are more possible to be synthesized. For NbBf, after a single H2 adsorption, it is found that the adsorption energy is 0.41 eV. In order to enhance the adsorption energy of H2, a vertical electric field is applied on a B/C/N sheet. It is shown that the H2 molecule is polarized under an external electric field. It can be seen that the adsorption energy increases dramatically with increasing electric field intensity. The bond length of H–H increases as the electric field increases within a certain range. The hydrogen molecules are not dissociated and are all stored in a molecular form.

Acknowledgements

This work is supported by the National Natural Science Foundation of China (Grant no. 21376013).

References

  1. J. Alper, Science, 2003, 299, 1686–1687 CrossRef CAS PubMed.
  2. R. D. Cortright, R. R. Davda and J. A. Dumesic, Nature, 2002, 418, 964–967 CrossRef CAS PubMed.
  3. L. Schlapbach and A. Züttel, Nature, 2001, 414, 353–358 CrossRef CAS PubMed.
  4. Y. Lu, R. Jin and W. Chen, Nanoscale, 2011, 3, 2476–2480 RSC.
  5. R. Coontz and B. Hanson, Science, 2004, 305, 957 CrossRef CAS.
  6. C. Liu, Y. Y. Fan, M. Liu, H. T. Cong, H. M. Cheng and M. S. Dresselhaus, Science, 1999, 286, 1127–1129 CrossRef CAS PubMed.
  7. H. Kruse and S. Grimme, J. Phys. Chem. C, 2009, 113, 17006–17010 CAS.
  8. A. C. Dillon and M. J. Heben, Appl. Phys. A: Mater. Sci. Process., 2001, 72, 133–142 CrossRef CAS.
  9. J. Zhou, Q. Wang, Q. Sun, P. Jena and X. S. Chen, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 2801–2806 CrossRef CAS PubMed.
  10. M. Khazaei, M. S. Bahramy, N. S. Venkataramanan, H. Mizuseki and Y. Kawazoe, J. Appl. Phys., 2009, 106, 094303 CrossRef.
  11. L. P. Zhang, P. Wu and M. B. Sullivan, J. Phys. Chem. C, 2011, 115, 4289–4296 CAS.
  12. J. Wang, C. H. Lee and Y. K. Yap, Nanoscale, 2010, 2, 2028–2034 RSC.
  13. R. Ma, Y. Bando, H. Zhu, T. Sato, C. Xu and D. Wu, J. Am. Chem. Soc., 2002, 124, 7672–7673 CrossRef CAS PubMed.
  14. C. C. Tang, Y. Bando, T. Sato, K. Kurashima, X. X. Ding, Z. W. Gan and S. R. Qi, Appl. Phys. Lett., 2002, 80, 4641–4643 CrossRef CAS.
  15. J. Niu, B. K. Rao and P. Jena, Phys. Rev. Lett., 1992, 68, 2277 CrossRef CAS PubMed.
  16. M. Yoon, S. Yang, C. Hicke, E. Wang, D. Geohegan and Z. Zhang, Phys. Rev. Lett., 2008, 100, 206806 CrossRef PubMed.
  17. W. Liu, Y. H. Zhao, J. Nguyen, Y. Li, Q. Jiang and E. J. Lavernia, Carbon, 2009, 47, 3452–3460 CrossRef CAS.
  18. W. Liu, Y. H. Zhao, Y. Li, Q. Jiang and E. J. Lavernia, J. Phys. Chem. C, 2009, 113, 2028–2033 CAS.
  19. S. Shi, J. Y. Hwang, X. Li, X. Sun and B. I. Lee, Int. J. Hydrogen Energy, 2010, 35, 629–631 CrossRef CAS.
  20. R. H. Baughman, H. Eckhardt and M. Kertesz, J. Chem. Phys., 1987, 87, 6687–6699 CrossRef CAS.
  21. N. Narita, S. Nagai, S. Suzuki and K. Nakao, Phys. Rev. B: Condens. Matter Mater. Phys., 1998, 58, 11009 CrossRef CAS.
  22. L. D. Pan, L. Z. Zhang, B. Q. Song, S. X. Du and H. J. Gao, Appl. Phys. Lett., 2011, 98, 173102 CrossRef.
  23. M. Long, L. Tang, D. Wang, Y. Li and Z. Shuai, ACS Nano, 2011, 5, 2593–2600 CrossRef CAS PubMed.
  24. G. Li, Y. Li, H. Liu, Y. Guo, Y. Li and D. Zhu, Chem. Commun., 2010, 46, 3256–3258 RSC.
  25. B. Delley, J. Chem. Phys., 1990, 92, 508–517 CrossRef CAS.
  26. B. Delley, J. Chem. Phys., 2000, 113, 7756–7764 CrossRef CAS.
  27. J. P. Perdew and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 45, 13244–13249 CrossRef.
  28. Y. Okamoto and Y. Miyamoto, J. Phys. Chem. B, 2001, 105, 3470–3474 CrossRef CAS.
  29. J. I. Martínez, I. Cabria, M. J. López and J. A. Alonso, J. Phys. Chem. C, 2009, 113, 939–941 Search PubMed.
  30. Z. M. Ao, Q. Jiang, R. Q. Zhang, T. T. Tan and S. Li, J. Appl. Phys., 2009, 105, 074307 CrossRef.
  31. Y. V. Shtogun and L. M. Woods, J. Phys. Chem. C, 2009, 113, 4792–4796 CAS.
  32. F. Ortmann, F. Bechstedt and W. G. Schmidt, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 205101 CrossRef.
  33. S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553–566 CrossRef CAS.
  34. I. Cabria, M. J. López and J. A. Alonso, J. Chem. Phys., 2008, 128, 144704 CrossRef CAS PubMed.
  35. X. Deng, M. Si and J. Dai, J. Chem. Phys., 2012, 137, 201101 CrossRef PubMed.
  36. B. Kiran, A. K. Kandalam and P. Jena, J. Chem. Phys., 2006, 124, 224703 CrossRef PubMed.
  37. J. H. Guo, W. D. Wu and H. Zhang, Struct. Chem., 2009, 20, 1107–1113 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra07761j

This journal is © The Royal Society of Chemistry 2014