Next Article in Journal
Preparation of Electrospun Gelatin Mat with Incorporated Zinc Oxide/Graphene Oxide and Its Antibacterial Activity
Next Article in Special Issue
The Influence of Halogenated Hypercarbon on Crystal Packing in the Series of 1-Ph-2-X-1,2-dicarba-closo-dodecaboranes (X = F, Cl, Br, I)
Previous Article in Journal
Quantum Dot Labelling of Tepary Bean (Phaseolus acutifolius) Lectins by Microfluidics
Previous Article in Special Issue
Boron Chemistry for Medical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Complexes Between Adamantane Analogues B4X6 -X = {CH2, NH, O ; SiH2, PH, S} - and Dihydrogen, B4X6:nH2 (n = 1–4)

1
Instituto de Química-Física “Rocasolano”, CSIC, Serrano, 119, E-28006 Madrid, Spain
2
Instituto de Química Médica, CSIC, Juan de la Cierva, 3, E-28006 Madrid, Spain
*
Author to whom correspondence should be addressed.
Submission received: 6 January 2020 / Revised: 16 February 2020 / Accepted: 19 February 2020 / Published: 26 February 2020

Abstract

:
In this work, we study the interactions between adamantane-like structures B4X6 with X = {CH2, NH, O ; SiH2, PH, S} and dihydrogen molecules above the Boron atom, with ab initio methods based on perturbation theory (MP2/aug-cc-pVDZ). Molecular electrostatic potentials (MESP) for optimized B4X6 systems, optimized geometries, and binding energies are reported for all B4X6:nH2 (n = 1–4) complexes. All B4X6:nH2 (n = 1–4) complexes show attractive patterns, with B4O6:nH2 systems showing remarkable behavior with larger binding energies and smaller B···H2 distances as compared to the other structures with different X.

Graphical Abstract

1. Introduction

Hydrogen storage is becoming an important issue regarding the energetic needs of our modern world [1,2]. Different methods are designed for hydrogen storage [3,4,5], including cryogenics, high pressures, and chemical compounds that reversibly release dihydrogen upon heating [6]. Among the different systems described in the literature, the metal organic frameworks (MOF) have been the most successful ones as hydrogen storage [7,8,9,10,11].
Related to the computational study in the chemical trapping of dihydrogen, noncovalent interactions must be taken into account [12], given the relatively weak attracting force—mostly dispersive—derived from two neutral molecules, leading to general complexes with formula Z:nH2, where Z is a neutral molecule and n is the number of H2 molecules attached to the Z neutral system. A number of theoretical articles have been devoted to the interaction of dihydrogen with metallic systems [13,14,15,16,17,18,19,20,21,22].
The adamantane scaffold is widely used due to their steric [23,24], lipophilic [25] and rigid characteristics [26,27,28,29,30]. Derivatives that include heteroatoms have been synthesized, in addition to the well-known aza-(AZADO and hexamethylenetetramine) [31,32,33,34] and oxo-derivatives (tetrodotoxin) [35]. Other derivatives involving C/As/O (arsenicin A) [36], C/N/S (tetramethylenedisulfotetramine) [37], P/S (phosphorus pentasulfide) [38], P/N [39], and C/P/S [40] have also been described.
For the study of dihydrogen complexes, we propose the use of adamantane analog systems where each CH tetrahedral vertex in adamantane is substituted by a Boron atom, and the remaining CH2 moieties is substituted by divalent X groups, with X = {CH2, NH, O; SiH2, PH, S} thus leading to B4X6 tetrahedral molecules, as shown in Scheme 1.
The B4X6:nH2 complexes (n = 1–4) could be formed by approaching H2 molecules towards the B atom centered vertically along the four equivalent local Ĉ3 rotation axis. The existence of adamantane analogs B4X6 is not known, except for the 1-boraadamantane, where only one tetrahedral CH vertex is substituted by a B atom [41], with the remaining structure unaltered; this system has also been the target for a computational study for complex formation [42] with Lewis acids and superacids. Further substitutions of B atoms in tetrahedral sites have been studied from a theoretical point of view only [43]. Directly related to this work is the concept of the, σ-hole, proposed by Politzer and Murray [44,45,46], which refers to the electron-deficient outer lobe of a p orbital involved in a covalent bond, especially when one of the atoms is highly electronegative, that present positive values for the electrostatic potential [47,48]. On the other hand, when the deficient outer lobe of a p orbital involved in a covalent bond is perpendicular or axially oriented with respect of the molecular frame, the electrostatic nature of the interactions considered between B atoms in B4X6 systems and H2 molecules can be rationalized in terms of π-holes [49]. We should emphasize the relation between the Lewis acidity of trivalent B centers and the (non)planarity of the structure surrounding the B atom [50].

2. Results and Discussion

2.1. Molecular Electrostatic Potential (MESP) and π-Holes in B4X6 Systems

As stated above, we can rationalize the electrostatic nature of the interaction between B4X6 and H2 molecules in terms of π-holes, namely regions of positive electrostatic potential perpendicular or axially-oriented (as in the adamantane structure) with respect to a portion of the molecular framework, as shown in Figure 1. The empty/electron-deficient p lobe of Boron pointing outwards in the B4X6 systems is an electron (or surplus charge density) attractor, as shown by the positive values of the π-holes.
As shown in Figure 1b, the MESP of the B4O6 system shows areas of positive (blue) and negative (red) values corresponding to deficient electron density (negative charge attractor) and surplus electron density (positive charge attractor) areas, respectively. Clearly, the electron density deficiency area above the Boron atoms in B4X6 could attract the electron density of the σ bond of the H2 molecule. As shown in Figure 1b, the large electronegativity of the three oxygen atoms bound to boron must have a stronger effect on the attachment of H2 molecules as a function of the π-hole values:
π-hole (au): 0.131 (O) >> 0.058 (NH) > 0.047 (SiH2) > 0.034 (CH2) > 0.025 (PH, S).

2.2. Geometries and Energies of B4X6:nH2 Complexes (n = 1–4)

Figure 2 shows the optimized geometries of the isolated B4X6 systems at MP2/aug-cc-pVDZ level, corresponding to energy minima for all cases. The cartesian coordinates for the B4X6 optimized structures are gathered in Table S1, with the MP2 method and basis sets aug-cc-pVDZ and aug-cc-pVTZ, of double-ζ and triple-ζ quality respectively, including diffuse and polarization functions.
In the computations, we first obtain the energy profile of a frozen H2 molecule approaching this B atom (d distance) along the corresponding local C3 axis, as shown in Figure 3 for the B4X6:H2 complexes.
From Figure 3 we can clearly observe that all energy profiles are attractive for an H2 molecule down to 3 Å, and then three different curve patterns emerge: (i) for X = {CH2, NH, PH, S} the energy profile becomes repulsive when d < 3 Å (ii) for X = O, the energy minimum well is flatter and becomes repulsive shifting down to values of d ~ 1.7–2.0 Å; and finally (iii) for X = SiH2 the energy profile remains attractive down to 1.25 Å. The inset plot of Figure 3—upper right corner—shows an energy profile zoom-in of the region 2.5 Å < d < 3.3 Å in order to see more clearly the positions of the energy minima regions, for a given X. Clearly, the {CH2, NH}, and {PH, S} curves show similar energy minima regions: We turn from an attractive to a repulsive system at d ≤ 2.1 Å (CH2), 2.2 Å (NH), 2.48 Å (PH), and 2.55 Å (S). As stated above, a zoom-in of the energy profile for 2.5 Å ≤ d ≤ 3.3 Å is included in order to unveil the effect of approaching a H2 molecule to the B4X6 system where several curves have similar profiles. If we observe closely the curves from the zoom-in inset of Figure 3, the energy minima for CH2, NH, PH, and S are located as follows: dmin(CH2) ~ 2.73 Å, dmin(NH) ~ 2.77 Å, dmin(PH) ~ 3.05 Å, dmin(S) ~ 3.06 Å. For O and SiH2 there are no minima within this region since the curves are always attractive.
Once we choose the d which corresponds to the energy minimum in Figure 3, we relax the nuclear coordinates in the whole complexes hence determining the energy minimum structure for the B4X6:nH2 systems. Due to the different behavior of the B4(SiH2)6 system versus an H2 molecule—permanent attractive profile for d down to 1.25 Å—as compared to the other complexes—Figure 3—and the lack of an energy minimum geometry for the B4(SiH2)6:H2 complex – a geometry optimization shows a bond breaking in the H2 molecule and a rearrangement of the B4(SiH2)6 adamantane structure—this system will be analyzed further in another work. The optimized structures for all B4X6:nH2 complexes (n = 1–4) are depicted in Figure S2, except for B4O6:nH2 (n = 1–4), the latter shown in Figure 4. In Table 1 we gather the average B···H2 and H···H distances in the optimized geometries of the different B4X6:nH2 complexes, all corresponding to energy minima at the MP2/aug-cc-pVDZ level of theory.
As regards to the B···H2 distances (d) in the B4X6:nH2 complexes, as gathered in Table 1, three groups can be clearly distinguished: (i) X = {CH2, NH} with d distances of ~ 2.7 Å (ii) X = {PH, S} with d distances of ~ 3.0 Å and (iii) X = O, with shorter d distances down to ~ 1.6 Å. The case for B4O6 is quite remarkable. As more H2 molecules are attached to B4O6, the d(B···H2) distances are elongated steadily up to d ~ 1.85 Å, and the H···H molecules remain slightly stretched down to Δ ~ 0.016 Å. This behavior for the B4O6 systems is unique as compared to the other systems since in the latter the attachment of the H2 molecules is quite farther to the B atom and the H2 molecules remain practically unaltered. There is no clear tendency—as compared to B4O6—for the d distances as more H2 molecules are added for X = {CH2, NH, PH, S}, with tiny differences for the series 1 ≤ n ≤ 4.
Turning now to the H-H distances in the complexes, as shown in Table 1, in the energy minimum structures of the complexes, the H-H distances are very similar as compared to the isolated H2 molecule, 0.755 Å. However, there is an exception for the oxygen complexes: when one H2 molecule is attached to the B4O6 system, the H···H bond is elongated by Δ ~ 0.02 Å. As further H2 molecules are attached to the B4O6 system, this elongated H-H bond is shortened consecutively by ~ 0.004 Å, down to 0.763 Å in each H-H molecule of the B4O6:4H2 complex, though still 0.008 Å longer than in the isolated H-H molecule.
Finally, we show the computed binding energies of the H2 molecules for the different complexes B4X6:nH2 (n = 1–4), as seen in Table 2 and displayed in Figure 5, where we also include the CBS extrapolated values. As expected from the computed MESP and energy profiles in B4X6:H2 complexes, the larger binding energy for one H2 molecule corresponds to the B4O6 system, with ΔE ~ 29 kJ/mol (ΔECBS ~ 22 kJ/mol). For comparative purposes, the electronic binding energy of the water dimer is ΔE[(H2O)2] ~ 21 kJ/mol [51].
The maximum binding energy for the complexes corresponds to B4O6:4H2 with a value of 79 kJ/mol (CBS extrapolation 60 kJ/mol). However, the binding energy of one H2 molecule attached to the other B4X6 systems is remarkably smaller in comparison, especially when the CBS extrapolation is added. The addition of more H2 molecules to the complexes shows practically additive relations for all X. As displayed in Figure 5, when extrapolated to the CBS limit, we can see several features regarding the binding energies in B4X6:nH2 (n = 1–4) complexes as compared to the MP2/aug-cc-pVDZ energies: (1) the CBS extrapolated binding energies are smaller for a given X and n (2) the (absolute value of the) slope of ΔE versus n (number of H2) molecules decreases for CBS extrapolated values (3) both CBS extrapolated and non-extrapolated binding energies follow a similar linear trend, except for X = O, the latter with clearly larger (CBS) binding energies, from 20 kJ/mol (n = 1) to 60 kJ/mol (n = 4).
We should notice that for X = O, though the CBS extrapolated slope is smaller than the non-extrapolated one, yet this slope is larger (in absolute value) as compared to the other Xs hence the peculiar behavior of B4O6:nH2 as compared to complexes with different Xs. We should also emphasize the small differences (less than ~ 5 kJ/mol ) between binding energies for different Xs for a given number n of attached H2 molecules, with the exception of X = O with larger binding energies.

3. Computational Method

All geometries of the B4X6 systems and the corresponding B4X6:nH2 (n = 1–4) complexes were optimized with second-order Møller-Plesset perturbation theory (MP2) [52] and a double-ζ basis set including polarization and diffuse functions [53], such as aug-cc-pVDZ. The interactions between B4X6 systems and H2 molecules are clearly of noncovalent nature, weaker than conventional chemical bonds, given the closed-shell nature of the species involved and the lack of any further singlet coupling between unpaired electrons. We search for stable complexes B4X6:nH2 and dispersive corrections are important given the neutral and spin-zero nature of the involved systems, hence the use of MP2 theory in computations. This theory improves on the Hartree–Fock (a mean-field—molecular-orbital—theory of electronic structure) method by adding electron correlation effects by means of Rayleigh–Schrödinger perturbation theory (RS–PT).
The quantum-chemical computations in this work were carried out at the MP2 level of theory with the scientific software Gaussian09 (Gaussian Inc, Wallingford, CT, USA) [54], and the molecular electrostatic potential (MESP) for the B4X6 systems was computed with the DAMQT program [55,56], also at the MP2 level of theory. Frequency computations were performed in order to check the energy minimum nature in all B4X6 systems and B4X6:nH2 complexes (n = 1–4). The binding energies for the B4X6:nH2 complexes are computed as ΔE = E(B4X6:nH2) – E(B4X6) – n·E(H2), and reported in kJ/mol. Further geometry optimizations of all B4X6 complexes were carried out with the MP2/aug-cc-pVTZ computational model—with a triple-ζ basis set—in order to check the validity of the optimized geometries (see Supplementary Information). A single-point energy profile (MP2) versus B···H2 distance in the complex B4(CH2)6:H2 was computed using both basis sets: aug-cc-pVDZ and aug-cc-pVTZ (see Supplementary Information), double-ζ and triple-ζ respectively. As shown in Figure S1, the results show similar profiles along the local Ĉ3 axis of rotation on the B atom and therefore we can confirm the validity of the MP2/aug-cc-pVDZ computational model for geometries and binding energies.
In order to assess the dependency of the binding energies on basis set incompleteness, we also computed the binding energies of all complexes in the extrapolated complete basis set (CBS) limit. The CBS energy has been calculated by extrapolation of the HF energies calculated at aug-cc-pVkZ, with k = D, T and Q, and Equation (1) and the correlation part with Equation (2). The sum of the two components (HF and correlation) (Equation (3)) provides the MP2(CBS) energy.
E H F , k = E H F , l i m + A e B k
with k = 2, 3 and 4 for aug-cc-pVDZ, aug-cc-VTZ and aug-cc-pVQZ basis sets, respectively [57,58].
E c o o r , k = E c o r r , l i m + A k 3
with k = 3 and 4 for aug-cc-VTZ and aug-cc-pVQZ basis sets, respectively [59]. Finally, we have
E M P 2 ( C B S ) = E H F , l i m + E c o r r , l i m

4. Conclusions

From the results obtained in this work we can conclude with the following points:
1) The MESP in the adamantane-like structures B4X6, with X ={CH2, NH, O ; SiH2, PH, S}, show π-holes above the B atom with electron (density) attraction forces largest for B4O6 and lowest for B4(PH)6 and B4S6.
2) The energy profiles of one H2 molecule approaching along a C3v axes the B atom of B4X6 systems show attractive patterns up to certain values for all systems where it turns to repulsive below ~ 2.1 Å for B4(CH2)6 and B4(NH)6 and below ~ 2.5 Å for B4(PH)6 and B4S6, except for B4(SiH2)6 where the profile is always attractive with H2 bond breaking and cage rearrangement. For B4O6, there is a flat energy minimum region within 1.7–2.0 Å.
3) The attraction strength for electron density towards boron atoms in B4X6 is also shown in the energy profiles of the B4X6:H2 complexes as a function of the B···H2 distance d. The d distances in the energy minima structures coincide with the predicted distances from the energy profiles.
4) The binding energies of the B4X6:nH2 complexes—n = 1–4, follow a similar linear additive pattern (in magnitude and direction) for all X, except X = O, with larger binding energies. CBS extrapolation shows a significant decrease in the binding energies.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/1420-3049/25/5/1042/s1, Figure S1: Plot of ΔE (kJ/mol) versus d (Å) for the complex [B4(CH2)6]···H2 for a range of distances 1.25 Å ≤ d ≤ 4.0 Å and a step Δd = 0.25 Å. MP2/aug-cc-pVDZ (red solid line) and MP2/aug-cc-pVTZ (green dashed line) computations, Figure S2: Optimized structures of the B4X6:nH2 complexes (n = 1-4), X = {CH2 ; NH, PH; S} with MP2/aug-cc-pVDZ computations. All geometries corrrespond to energy minima, Table S1: Cartesian coordinates (Å) of MP2/aug-cc-pVDZ (left column) and MP2/aug-cc-pVTZ (right column) optimized geometries corresponding to energy minima of the adamantane analogue systems B4X6 with X = {CH2, SiH2; NH, PH; O, S}. All molecules have Td symmetry except B4X6 with X = {NH, PH} which have C1 symmetry.

Author Contributions

Conceptualization, J.E., I.A. and J.M.O.-E..; methodology, I.A. and J.M.O.-E.; software, I.A. and J.M.O.-E.; validation, I.A. and J.M.O.-E.; formal analysis, I.A. and J.M.O.-E.; investigation, I.A. and J.M.O.-E.; resources, I.A. J.M.O.-E., and J.E.; writing—original draft preparation, J.M.O.-E. and I.A.; writing—review and editing, J.M.O.-E. and I.A.; supervision, J.E., I.A. and J.M.O.-E.; project administration, I.A.; funding acquisition, I.A. and J.E. All authors have read and agreed to the published version of the manuscript.

Funding

We are grateful to Spanish MICINN for financial support through project CTQ2018-094644-B-C22 and Comunidad de Madrid (P2018/EMT-4329 AIRTEC-CM).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zell, T.; Langer, R. Hydrogen storage; Walter de Gruyter: Berlin, Germany, 2019. [Google Scholar]
  2. Moradi, R.; Groth, K.M. Hydrogen storage and delivery: Review of the state of the art technologies and risk and reliability analysis. Int. J. Hydro. Energy 2019, 44, 12254–12269. [Google Scholar] [CrossRef]
  3. Preuster, P.; Alekseev, A.; Wasserscheid, P. Hydrogen storage technologies for future energy systems. Annu. Rev. Chem. Biomol. 2017, 8, 445–471. [Google Scholar] [CrossRef] [PubMed]
  4. Mohan, M.; Sharma, V.K.; Kumar, E.A.; Gayathri, V. Hydrogen storage in carbon materials—A review. Energy Stor. 2019, 1, 1–26. [Google Scholar] [CrossRef]
  5. Khan, N.; Dilshad, S.; Khalid, R.; Kalair, A.R.; Abas, N. Review of energy storage and transportation of energy. Energy Stor. 2019, 1, 1–49. [Google Scholar] [CrossRef]
  6. Schneemann, A.; White, J.L.; Kang, S.; Jeong, S.; Wan, L.F.; Cho, E.S.; Heo, T.W.; Prendergast, D.; Urban, J.J.; Wood, B.C.; et al. Nanostructured metal hydrides for hydrogen storage. Chem. Rev. 2018, 118, 10775–10839. [Google Scholar] [CrossRef] [PubMed]
  7. Suh, M.P.; Park, H.J.; Prasad, T.K.; Lim, D.-W. Hydrogen storage in metal-organic frameworks. Chem. Rev. 2012, 112, 782–835. [Google Scholar] [CrossRef]
  8. Kapelewski, M.T.; Runčevski, T.; Tarver, J.D.; Jiang, H.Z.H.; Hurst, K.E.; Parilla, P.A.; Ayala, A.; Gennett, T.; Fitzgerald, S.A.; Brown, C.M.; et al. Record High Hydrogen Storage Capacity in the Metal–Organic Framework Ni2(m-dobdc) at Near-Ambient Temperatures. Chem. Mater. 2018, 30, 8179–8189. [Google Scholar] [CrossRef]
  9. Zhu, B.; Zou, R.; Xu, Q. Metal–organic framework based catalysts for hydrogen evolution. Adv. Energy Mater. 2018, 8, 1801193. [Google Scholar] [CrossRef]
  10. Yan, Y.; da Silva, I.; Blake, A.J.; Dailly, A.; Manuel, P.; Yang, S.; Schröder, M. High volumetric hydrogen adsorption in a porous anthracene-decorated metal–organic framework. Inorg Chem. 2018, 57, 12050–12055. [Google Scholar] [CrossRef] [Green Version]
  11. Koizumi, K.; Nobusada, K.; Boero, M. Hydrogen storage mechanism and diffusion in metal–organic frameworks. Phys. Chem. Chem. Phys. 2019, 21, 7756–7764. [Google Scholar] [CrossRef]
  12. Hobza, P.; Müller-Dethlefs, K. Non-Covalent Interactions. Theory and Experiment; Royal Society of Chemistry: Cambridge, UK, 2009. [Google Scholar]
  13. Chakraborty, A.; Giri, S.; Chattaraj, P.K. Analyzing the efficiency of Mn–(C2H4) (M = Sc, Ti, Fe, Ni; n = 1, 2) complexes as effective hydrogen storage materials. Struct. Chem. 2011, 22, 823–837. [Google Scholar] [CrossRef] [Green Version]
  14. Ma, L.-J.; Jia, J.; Wu, H.-S. Computational investigation of hydrogen storage on scandium–acetylene system. Int. J. Hydro. Energy 2015, 40, 420–428. [Google Scholar] [CrossRef]
  15. Konda, R.; Titus, E.; Chaudhari, A. Adsorption of molecular hydrogen on inorganometallic complexes B2H4M (M.=Li, Be, Sc, Ti, V). Struct. Chem. 2018, 29, 1593–1599. [Google Scholar] [CrossRef]
  16. Konda, R.; Kalamse, V.; Deshmukh, A.; Chaudhari, A. Closoborate-transition metal complexes for hydrogen storage. RSC Adv. 2015, 5, 99207–99216. [Google Scholar] [CrossRef]
  17. Naumkin, F.Y.; Wales, D.J. Hydrogen trapped in Ben cluster cages: The atomic encapsulation option. Chem. Phys. Lett. 2012, 545, 44–49. [Google Scholar] [CrossRef]
  18. Meng, Y.; Han, Y.; Zhu, H.; Yang, Z.; Shen, K.; Suo, B.; Lei, Y.; Zhai, G.; Wen, Z. Two dimetallocenes with vanadium and chromium: Electronic structures and their promising application in hydrogen storage. Int. J. Hydro. Energy 2015, 40, 12047–12056. [Google Scholar] [CrossRef]
  19. Grabowski, S.; Ruiperez, F. Dihydrogen bond interactions as a result of H2 cleavage at Cu, Ag and Au centres. Phys. Chem. Chem. Phys. 2016, 18, 12810–12818. [Google Scholar] [CrossRef]
  20. Alkorta, I.; Elguero, J.; Solimannejad, M.; Grabowski, S.J. Dihydrogen bonding vs metal−σ interaction in complexes between H2 and metal hydride. J. Phys. Chem. A 2011, 115, 201–210. [Google Scholar] [CrossRef]
  21. Alkorta, I.; Montero-Campillo, M.M.; Elguero, J.; Yáñez, M.; Mó, O. Complexes between H2 and neutral oxyacid beryllium derivatives. The role of angular strain. Mol. Phys. 2019, 117, 1142–1150. [Google Scholar] [CrossRef]
  22. Alkorta, I.; Montero-Campillo, M.M.; Elguero, J.; Yáñez, M.; Mó, O. Complexes between neutral oxyacid beryllium salts and dihydrogen: A possible way for hydrogen storage? Dalton Trans. 2018, 47, 12516–12520. [Google Scholar] [CrossRef]
  23. Kayal, S.; Manna, U.; Das, G. Steric influence of adamantane substitution in tris-urea receptor: Encapsulation of sulphate and fluoride-water cluster. J. Chem. Sci. 2018, 130, 1–6. [Google Scholar] [CrossRef] [Green Version]
  24. Hooley, R.J.; Shenoy, S.R.; Rebek, J., Jr. Electronic and steric effects in binding of deep cavitands. Org. Lett. 2008, 10, 5397–5400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Wanka, L.; Iqbal, K.; Schreiner, P.R. The lipophilic bullet hits the targets: Medicinal chemistry of adamantane derivatives. Chem. Rev. 2013, 113, 3516–3604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Fort, R.C. Adamantane: The chemistry of diamond molecules; Marcel Dekker: New York, NY, USA, 1976. [Google Scholar]
  27. Spilovska, K.; Zemek, F.; Korabecny, J.; Nepovimova, E.; Soukup, O.; Windisch, M.; Kuca, K. Adamantane –A lead structure for drugs in clinical practice. Curr. Med. Chem. 2016, 23, 3245–3266. [Google Scholar] [CrossRef]
  28. Schwertfeger, H.; Fokin, A.A.; Schreiner, P.R. Diamonds are a chemist’s best friend: Diamondoid chemistry beyond adamantane. Angew. Chem. Int. Ed. 2008, 47, 1022–1036. [Google Scholar] [CrossRef]
  29. Bagrii, E.I.; Nekhaev, A.I.; Maksimov, A.L. Oxidative functionalization of adamantanes. Petrol. Chem. 2017, 57, 183–197. [Google Scholar] [CrossRef]
  30. Štimac, A.; Šekutor, M.; Mlinarić-Majerski, K.; Frkanec, L.; Frkanec, R. Adamantane in drug delivery systems and surface recognition. Molecules 2017, 22, 297. [Google Scholar] [CrossRef] [Green Version]
  31. Speckamp, W.N.; Dijkink, J.; Huisman, H.O. 1-Aza-adamantanes. J. Chem. Soc. D: Chem. Comm. 1970, 4, 197–198. [Google Scholar] [CrossRef]
  32. Shibuya, M.; Tomizawa, M.; Suzuki, I.; Iwabuchi, Y. 2-Azaadamantane N-Oxyl (AZADO) and 1-Me-AZADO:  highly efficient organocatalysts for oxidation of alcohols. J. Am. Chem. Soc. 2006, 128, 8412–8413. [Google Scholar] [CrossRef]
  33. Kirillo, A.M. Hexamethylenetetramine: An old new building block for design of coordination polymers. Coordin. Chem. Rev. 2011, 255, 1603–1622. [Google Scholar] [CrossRef]
  34. Dreyfors, J.M.; Jones, S.B.; Sayed, Y. Hexamethylenetetramine: A review. Am. Ind. Hyg. Assoc. J. 1989, 50, 579–585. [Google Scholar] [CrossRef] [PubMed]
  35. Bane, V.; Lehane, M.; Dikshit, M.; O’Riordan, A.; Furey, A. Tetrodotoxin: Chemistry, toxicity, source, distribution and detection. Toxins 2014, 6, 693–755. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Mancini, I.; Guella, G.; Frostin, M.; Hnawia, E.; Laurent, D.; Debitus, C.; Pietra, F. On the first polyarsenic organic compound from Nature: Arsenicin A from the new caledonian marine sponge echinochalina bargibanti. Chem. Eur. J. 2006, 12, 8989–8994. [Google Scholar] [CrossRef]
  37. Whitlow, K.S.; Belson, M.; Barrueto, F.; Nelson, L.; Henderson, A.K. Tetramethylenedisulfotetramine: Old agent and new terror. Ann. Emerg. Med. 2005, 45, 609–613. [Google Scholar] [CrossRef] [PubMed]
  38. Available online: https://en.wikipedia.org/wiki/Phosphorus_pentasulfide (accessed on 12 December 2019).
  39. Shi, Y.X.; Xu, K.; Clegg, J.K.; Ganguly, R.; Hirao, H.; Friščić, T.; García, F. The first synthesis of the sterically encumbered adamantoid phosphazane P4(NtBu)6: Enabled by mechanochemistry. Angew. Chem. Int. Ed. 2016, 55, 12736–12740. [Google Scholar] [CrossRef] [PubMed]
  40. Kutsumura, N.; Ohshita, R.; Horiuchi, J.; Tateno, K.; Yamamoto, N.; Saitoh, T.; Nagumo, Y.; Kawai, H.; Nagase, H. Synthesis of heterocyclic compounds with adamantane-like cage structures consisting of phosphorus, sulfur, and carbon. Tetrahedron 2017, 73, 5214–5219. [Google Scholar] [CrossRef]
  41. Mikhailov, B.M. Reactions of allylboranes with unsaturated compounds. Pure Appl. Chem. 1980, 52, 691–704. [Google Scholar] [CrossRef]
  42. El-Hamdi, M.; Solà, M.; Poater, J.; Timoshkin, A.Y. Complexes of adamantane-based group 13 Lewis acids and superacids: Bonding analysis and thermodynamics of hydrogen splitting. J. Comput. Chem. 2016, 37, 1355–1362. [Google Scholar] [CrossRef] [Green Version]
  43. García, J.C.; Justo, J.F.; Machado, W.V.M.; Assali, L.V.C. Functionalized adamantane: Building blocks for nanostructure self-assembly. Phys. Rev. B 2009, 80, 125421. [Google Scholar] [CrossRef] [Green Version]
  44. Murray, J.S.; Lane, P.; Politzer, P. A predicted new type of directional noncovalent interaction. Int. J. Quantum Chem. 2007, 107, 2286–2292. [Google Scholar] [CrossRef]
  45. Mohajeri, A.; Pakiari, A.H.; Bagheri, N. Theoretical studies on the nature of bonding in σ-hole complexes. Chem. Phys. Lett. 2009, 467, 393–397. [Google Scholar] [CrossRef]
  46. Politzer, P.; Murray, J.; Concha, M. σ-hole bonding between like atoms; a fallacy of atomic charges. J. Mol. Model. 2008, 14, 659–665. [Google Scholar] [CrossRef] [PubMed]
  47. Sánchez-Sanz, G.; Trujillo, C.; Solimannejad, M.; Alkorta, I.; Elguero, J. Orthogonal interactions between nitryl derivatives and electron donors: Pnictogen bonds. Phys. Chem. Chem. Phys. 2013, 15, 14310–14318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Solimannejad, M.; Ramezani, V.; Trujillo, C.; Alkorta, I.; Sánchez-Sanz, G.; Elguero, J. Competition and interplay between σ-hole and π-hole interactions: A computational study of 1:1 and 1:2 complexes of nitryl halides (O2NX) with ammonia. J. Phys. Chem. A 2012, 116, 5199–5206. [Google Scholar] [CrossRef] [PubMed]
  49. Murray, J.; Lane, P.; Clark, T.; Riley, K.; Politzer, P. σ-holes, π-holes and electrostatically-driven interactions. J. Mol. Model. 2012, 18, 541–548. [Google Scholar] [CrossRef]
  50. Saida, A.B.; Chardon, A.; Osi, A.; Tumanov, N.; Wouters, J.; Adjieufack, A.I.; Champagne, B.; Berionni, G. Pushing the Lewis acidity boundaries of boron compounds with non-planar triarylboranes derived from triptycenes. Angew. Chem. Int. Ed. 2019, 58, 1–6. [Google Scholar]
  51. Feyereisen, M.W.; Feller, D.; Dixon, D.A. Hydrogen bond energy of the water dimer. J. Phys. Chem. 1996, 100, 2993–2997. [Google Scholar] [CrossRef]
  52. Møller, C.; Plesset, M.S. Note on an approximation treatment for many-electron systems. Phys. Rev. 1934, 46, 618–622. [Google Scholar] [CrossRef] [Green Version]
  53. Kendall, R.A.; Dunning, T.H., Jr.; Harrison, R.J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef] [Green Version]
  54. Gaussian 09, Revision, E.01; Gaussian Inc.: Wallingford, CT, USA, 2019.
  55. Kumar, A.; Yeole, S.D.; Gadre, S.R.; López, R.; Rico, J.F.; Ramirez, G.; Ema, I.; Zorrilla, D. DAMQT 2.1.0: A new version of the DAMQT package enabled with the topographical analysis of electron density and electrostatic potential in molecules. J. Comput. Chem. 2015, 36, 2350–2359. [Google Scholar] [CrossRef]
  56. López, R.; Fernández Rico, J.; Ramírez, G.; Ema, I.; Zorrilla, D. DAMQT: A package for the analysis of electron density in molecules. Comput. Phys. Commun. 2009, 180, 1654–1660. [Google Scholar] [CrossRef]
  57. Feller, D. The use of systematic sequences of wave functions for estimating the complete basis set, full configuration interaction limit in water. J. Chem. Phys. 1993, 98, 7059. [Google Scholar] [CrossRef]
  58. Williams, T.G.; DeYonker, N.J.; Wilson, A.K. Hartree-Fock complete basis set limit properties for transition metal diatomics. J. Chem. Phys. 2008, 128, 044101. [Google Scholar] [CrossRef] [PubMed]
  59. Halkier, A.; Helgaker, T.; Jorgensen, P.; Klopper, W.; Olsen, J. Basis-set convergence of the energy in molecular Hartree–Fock calculations. Chem. Phys. Lett. 1999, 302, 437–446. [Google Scholar] [CrossRef]
Scheme 1. Substitution of CH and CH2 groups by B and X respectively in (a) adamantane leading to the (b) B4X6 systems studied in this work. One of the four equivalent local Ĉ3 rotation axis is also shown.
Scheme 1. Substitution of CH and CH2 groups by B and X respectively in (a) adamantane leading to the (b) B4X6 systems studied in this work. One of the four equivalent local Ĉ3 rotation axis is also shown.
Molecules 25 01042 sch001
Figure 1. (a) π-hole on the boron atom (a.u.) and (b) molecular electrostatic potential on the 0.001 au electron density isosurface for the B4O6 molecule. Red and blue colors indicate Molecular Electrostatic Potential (MESP) values < −0.015 and > +0.015 au, respectively. The location of the π-holes is indicated with a black dot.
Figure 1. (a) π-hole on the boron atom (a.u.) and (b) molecular electrostatic potential on the 0.001 au electron density isosurface for the B4O6 molecule. Red and blue colors indicate Molecular Electrostatic Potential (MESP) values < −0.015 and > +0.015 au, respectively. The location of the π-holes is indicated with a black dot.
Molecules 25 01042 g001
Figure 2. MP2/aug-cc-pVDZ optimized geometries of adamantane-like B4X6 systems, (a) X = CH2, (b) X = SiH2, (c) X = NH, (d) X = PH, (e) X = O, (f) X = S. All geometries correspond to energy minima. Cartesian coordinates gathered in Table S1.
Figure 2. MP2/aug-cc-pVDZ optimized geometries of adamantane-like B4X6 systems, (a) X = CH2, (b) X = SiH2, (c) X = NH, (d) X = PH, (e) X = O, (f) X = S. All geometries correspond to energy minima. Cartesian coordinates gathered in Table S1.
Molecules 25 01042 g002aMolecules 25 01042 g002b
Figure 3. Energy profiles of ΔE = E(B4X6:H2) – E(H2), in kJ/mol, versus d, in Ångström, for different X at MP2/aug-cc-pVDZ computational level. The geometries of B4X6 and H2 are kept frozen along the d coordinate. The inset plot on the upper right corner corresponds to a zoom-in region of ΔE versus d within the range 2.5 ≤ d ≤ 3.3 Å, showing the corresponding energy minima for CH2, NH, PH and S systems.
Figure 3. Energy profiles of ΔE = E(B4X6:H2) – E(H2), in kJ/mol, versus d, in Ångström, for different X at MP2/aug-cc-pVDZ computational level. The geometries of B4X6 and H2 are kept frozen along the d coordinate. The inset plot on the upper right corner corresponds to a zoom-in region of ΔE versus d within the range 2.5 ≤ d ≤ 3.3 Å, showing the corresponding energy minima for CH2, NH, PH and S systems.
Molecules 25 01042 g003
Figure 4. Optimized geometries for the B4O6:nH2 complexes (n = 1–4) with MP2/aug-cc-pVDZ computations. All geometries corrrespond to energy minima.
Figure 4. Optimized geometries for the B4O6:nH2 complexes (n = 1–4) with MP2/aug-cc-pVDZ computations. All geometries corrrespond to energy minima.
Molecules 25 01042 g004
Figure 5. Binding energies of the B4X6:nH2 complexes (n = 1–4) as function of attached H2 molecules. MP2/aug-cc-pVDZ computations (solid lines) and extrapolated MP2/CBS limit (dashed lines).
Figure 5. Binding energies of the B4X6:nH2 complexes (n = 1–4) as function of attached H2 molecules. MP2/aug-cc-pVDZ computations (solid lines) and extrapolated MP2/CBS limit (dashed lines).
Molecules 25 01042 g005
Table 1. Average B···H2 (d) and H···H distances (Å) in the B4X6:nH2 complexes, X = {CH2 ; NH, PH; O, S} with MP2/aug-cc-pVDZ computations. The H-H distance in the isolated H2 molecule is 0.755 Å, computed at the MP2/aug-cc-pVDZ level of theory.
Table 1. Average B···H2 (d) and H···H distances (Å) in the B4X6:nH2 complexes, X = {CH2 ; NH, PH; O, S} with MP2/aug-cc-pVDZ computations. The H-H distance in the isolated H2 molecule is 0.755 Å, computed at the MP2/aug-cc-pVDZ level of theory.
B···H2H-H
X1:11:21:31:41:11:21:31:4
CH22.7372.7402.7452.7500.7550.7550.7540.754
NH2.7772.7272.7052.7270.7550.7550.7550.755
PH3.0803.0453.0463.0490.7550.7550.7550.755
O1.6241.6871.7661.8500.7740.7700.7660.763
S3.0713.0703.0683.0680.7550.7550.7550.755
Table 2. Binding energies (kJ/mol) in optimized (B4X6:nH2) complexes, X = {CH2, NH, PH, O, S} with MP2/aug-cc-pVDZ computations and the MP2 complete basis set (CBS) limit obtained by extrapolation of the HF energies calculated at aug-cc-pVkZ, with k = D, T and Q, following Equations (1)–(3). ΔE(1:n) for a given X corresponds to the binding energy of the complex B4X6:nH2.
Table 2. Binding energies (kJ/mol) in optimized (B4X6:nH2) complexes, X = {CH2, NH, PH, O, S} with MP2/aug-cc-pVDZ computations and the MP2 complete basis set (CBS) limit obtained by extrapolation of the HF energies calculated at aug-cc-pVkZ, with k = D, T and Q, following Equations (1)–(3). ΔE(1:n) for a given X corresponds to the binding energy of the complex B4X6:nH2.
XΔE(1:1)ΔE(1:2)ΔE(1:3)ΔE(1:4)
MP2 CBS MP2 CBS MP2 CBS MP2 CBS
CH2−6.6 −2.8−13.3 −5.4−20.1 −8.0 −26.8 −10.7
NH−5.8 −2.5−12.9 −6.1−20.0 −9.7 −25.8 −12.2
PH−6.2 −1.6−13.3 −3.9−19.9 −6.1−26.4 −7.7
O−28.6 −22.1−49.5 −37.6−65.5 −49.1−79.2 −58.9
S−7.9 −3.3−15.8 −6.7−23.4 −10.1−31.6 −13.5

Share and Cite

MDPI and ACS Style

Oliva-Enrich, J.M.; Alkorta, I.; Elguero, J. Complexes Between Adamantane Analogues B4X6 -X = {CH2, NH, O ; SiH2, PH, S} - and Dihydrogen, B4X6:nH2 (n = 1–4). Molecules 2020, 25, 1042. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25051042

AMA Style

Oliva-Enrich JM, Alkorta I, Elguero J. Complexes Between Adamantane Analogues B4X6 -X = {CH2, NH, O ; SiH2, PH, S} - and Dihydrogen, B4X6:nH2 (n = 1–4). Molecules. 2020; 25(5):1042. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25051042

Chicago/Turabian Style

Oliva-Enrich, Josep M., Ibon Alkorta, and José Elguero. 2020. "Complexes Between Adamantane Analogues B4X6 -X = {CH2, NH, O ; SiH2, PH, S} - and Dihydrogen, B4X6:nH2 (n = 1–4)" Molecules 25, no. 5: 1042. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25051042

Article Metrics

Back to TopTop