Next Article in Journal
A Detailed Study of Electronic and Dynamic Properties of Noble Gas–Oxygen Molecule Adducts
Next Article in Special Issue
Theoretical Study of Hydrogen Production from Ammonia Borane Catalyzed by Metal and Non-Metal Diatom-Doped Cobalt Phosphide
Previous Article in Journal
Novel Complexes of 3-[3-(1H-Imidazol-1-yl)propyl]-3,7-diaza-bispidines and β-Cyclodextrin as Coatings to Protect and Stimulate Sprouting Wheat Seeds
Previous Article in Special Issue
Relationship between the Molecular Geometry and the Radiative Efficiency in Naphthyl-Based Bis-Ortho-Carboranyl Luminophores
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

B3Al4+: A Three-Dimensional Molecular Reuleaux Triangle

1
Nanocluster Laboratory, Institute of Molecular Science, Shanxi University, Taiyuan 030006, China
2
Department of Chemistry, University of Helsinki, A. I. Virtasen Aukio 1, P.O. Box 55, FIN-00014 Helsinki, Finland
3
Instituto de Ciencias Químicas Aplicadas, Facultad de Ingeniería, Universidad Autónoma de Chile, Av. Pedro de Valdivia 425, Santiago 7500912, Chile
4
Fachbereich Chemie, Philipps-Universitt Marburg Hans-Meerwein-Straße, 35043 Marburg, Germany
5
Departamento de Física Aplicada, Centro de Investigación y de Estudios Avanzados, Unidad Mérida. Km 6 Antigua Carretera a Progreso. Apdo., Postal 73, Cordemex, Merida 97310, Mexico
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(21), 7407; https://doi.org/10.3390/molecules27217407
Submission received: 22 September 2022 / Revised: 19 October 2022 / Accepted: 23 October 2022 / Published: 1 November 2022
(This article belongs to the Special Issue New Boron Chemistry: Current Advances and Future Prospects)

Abstract

:
We systematically explore the potential energy surface of the B3Al4+ combination of atoms. The putative global minimum corresponds to a structure formed by an Al4 square facing a B3 triangle. Interestingly, the dynamical behavior can be described as a Reuleaux molecular triangle since it involves the rotation of the B3 triangle at the top of the Al4 square. The molecular dynamics simulations, corroborating with the very small rotational barriers of the B3 triangle, show its nearly free rotation on the Al4 ring, confirming the fluxional character of the cluster. Moreover, while the chemical bonding analysis suggests that the multicenter interaction between the two fragments determines its fluxionality, the magnetic response analysis reveals this cluster as a true and fully three-dimensional aromatic system.

Graphical Abstract

1. Introduction

In 2010, Wang and coworkers detected an anionic cluster of nineteen boron atoms in the gas phase through a photoelectron spectroscopy experiment [1]. The structure has a planar pentagonal inner B6 core encircled by a B13 ring (Figure 1a). A couple of years later, Merino and coworkers found that the inner B6 fragment rotates almost freely with respect to the periphery [2], so they labeled this type of system as a Wankel-type rotor [2,3,4,5,6,7,8]. Over time, other boron clusters were also reported to exhibit similar behavior, including B13+ as the most prominent case [9,10]. In fact, the dynamic behavior of B13+ was experimentally verified by Asmis et al. in 2016 using cryogenic ion vibrational spectroscopy [11,12]. The main reason for this fascinating fluxionality is attributed to the multicenter bonds, often found in boron clusters [13]. Moreover, this property is not exclusive to only planar forms; compasses [14,15], drums [16], stirrers, sphere-shaped clusters [17], and other doped boron architectures have similar dynamic behavior [18,19,20,21]. Other related examples are the global minima of B7M2 and B8M2 (M = Zn, Cd, Hg), which can be described as an M2 dimer spinning freely on a boron wheel resembling a magnetic stirrer placed on a baseplate [22]. The reader interested in more details on fluxionality in boron clusters is referred to [12].
Inspired by these fluxional boron systems, we analyze the dynamical properties of the global minimum of B3Al4+ (Figure 1b). This cluster has an Al4 square facing a B3 triangle. Although this system was recently reported by Wen et al. [23], they overlooked its fluxional properties. Moreover, upon further exploration of the nature of the bonding and aromaticity, we find that this system has all the characteristics to define it as a true and fully three-dimensional aromatic system.

2. Computational Details

The PES was systematically explored using the Coalescence Kick (CK) program and a modified genetic algorithm implemented in GLOMOS [24,25]. Initial screening in singlet and triplet states was performed for both programs at the PBE0/LAN2DZ level [26,27]. In the range of 30 kcal mol−1 above the putative global minimum, the lowest isomers were minimized and characterized at the PBE0/def2-TZVP level [28]. Final energies were computed at the CCSD(T) [29]/def2-TZVP level, including the zero-point energy correction (ZPE) at the PBE0/def2-TZVP level. Thus, the energy discussion is based on the CCSD(T)/def2-TZVP//PBE0/def2-TZVP results. Natural bond orbital analysis was conducted at the PBE0 level to obtain natural atomic charges (using the NBO6 program [30]) and Wiberg bond indices (WBI) [31]. The Born–Oppenheimer molecular dynamics [32] (BOMD) were carried out at the PBE0/6-31G(d) level to ascertain the fluxional behavior. The aromatic character was evaluated by computing the magnetic response to a uniform external magnetic field. This is achieved by calculating the magnetically induced current density [33,34,35,36] (Jind) and the induced magnetic field [37,38,39] (Bind) using the GIMIC [33,34,35,36] and Aromagnetic [40] programs, respectively. Since magnetic properties computed with the BHandHLYP [41] functional yield results in good agreement with those obtained by CCSD(T) computations [42], Jind and Bind were calculated at the BHandHLYP/def2-TZVP level using gauge-including atomic orbitals (GIAOs) [43,44]. All these calculations were performed with Gaussian 16 [45]. Furthermore, the adaptive natural density partitioning [46] (AdNDP) analysis was also carried out to understand the nature of the interactions using the Multiwfn program [47].

3. Discussion

The putative global minimum of B3Al4+ is a singlet with Cs symmetry (1), in which an Al4 square interacts with a B3 triangle, forming a polyhedron (see Figure 2). The B-B distances range from 1.60–1.63 Å, creating an isosceles triangle with a B-B bond aligned with an Al-Al bond. On the other hand, the Al4 ring is a trapezoid or quasi-square, with Al-Al bond distances lying between 2.60–2.77 Å. The distance between the two fragments is 1.73 Å.
The energetically closest isomer is only 5.1 kcal mol−1 above the global minimum and consists of an Al4 tetrahedron. The third isomer differs from the putative global minimum by exchanging a B atom for an Al atom and is 8.1 kcal/mol higher in energy. Note that all isomers in Figure S1 have a B3 unit. As shown in Figure S2, the lowest vibrational frequency of 1 is 69 cm−1, which corresponds to the rotation of the B3 fragment. Following this smooth rotation mode, we identify the transition state (TS) belonging to the Cs point group. The structural variations between the TS and the global minimum are negligible. The rotation barrier is only 0.1 kcal/mol (including the zero-point energy correction), which means that the B3 ring can undergo almost free rotation on top of the Al4 ring.
Figure 3 shows the structural evolution of B3Al4+ with the B3 triangle rotating clockwise. Starting from the global minimum, fragment B3 must rotate 15° to reach the maximum of the rotation barrier. A new minimum is achieved by rotating B3 another 15°, and a complete rotation lap is obtained by repeating this process twelve times. BOMD simulations at 300 K for 25 ps and starting from the global minimum structure show that the B3 triangle rotates with respect to the Al4 ring, just like a three-dimensional (3D) molecular Reuleaux triangle. The average distance between the B3 triangle and the Al4 quasi-square remains constant (see Movie 1 in the supplementary material).
A chemical bonding analysis is crucial to understanding the dynamical behavior of B3Al4+. The natural atomic charges and WBIs for 1 and TS are shown in Figure S3. The charge distribution is such that the B3 fragment has a charge of −2.0 |e| and the Al4 moiety of 3.0|e|. Instead, the WBI values show that the B-B bonds (WBI around 1.1) have a higher covalent character than the Al-Al bonds (WBI between 0.32 and 0.65), but even more fascinating is that the WBI values for the B-Al bonds are of the same magnitude as the latter (0.49 and 0.62), indicating that there is significant orbital overlap (an electron delocalization) between both fragments. We also employed AdNDP, an extension of the NBO analysis. AdNDP analysis recovers not only Lewis bond elements (lone pairs and 2c-2e bonds) but also multicenter bonds (nc-2e, n ≥ 3). For B3Al4+, the analysis identifies three 3c-2e σ-bonds with occupation numbers (ON) of 1.99–1.83 |e| and one π-bond (ON = 1.71 |e|) located mainly throughout the B3 fragment. This electron distribution is like that reported for the aromatic cyclopropenyl cation (C3H3+). On the other hand, the Al4 ring has three 4c-2e σ-bonds with ON of 1.99–1.88 |e|. The B3 triangle and the Al4 trapezoid interact via three 7c-2e bonds (Figure 4). Thus, the absence of 2c-2e bonds between the two fragments favors free rotation [48]. The bonding pattern for the TS is similar to that of the global minimum, as shown in Figure S4. The orbital compositions of canonical molecular orbitals (CMOs) for 1 and TS are listed in Tables S1 and S2, which support the bonding pattern provided by AdNDP.
As shown in Figure S5, the B3 ring has six σ- and two π-electrons, i.e., in principle, it satisfies Hückel’s aromaticity rule for both the σ- and π-clouds. Now, while the Al4 ring has six electrons, which we could classify as σ, the region between the two rings is also connected via six electrons, so both the Al4 fragment and the area between the rings satisfy the Hückel’s rule. The reality is that the orbital separation is complicated, and all the bonds in B3Al4+ have a multicentric character, i.e., they are fully delocalized in the small cage.
Aromaticity of B3Al4+ is confirmed by its magnetic response to an external magnetic field perpendicular to the Al4 ring. The Bind analysis reveals strong shielding values (<−50 ppm) of the z-component of Bind (Bindz) along the region between the Al4 and B3 rings. This can be explained by the presence of an entirely diatropic current density in this region (Figure 5a). Thus, the interaction of both rings creates a ring current in the region between the two rings instead of being two separate ring currents, implying that B3Al4+ has true three-dimensional aromaticity (Figure 5b). Integration of Jind in a plane that intersects the Al-Al and B-B bonds yields a ring–current strength (Jind) of 33.3 nA/T. Changes in the current density can be obtained by calculating the spatial derivative of the ring–current strength along the vertical z-axis (dJind/dz) [49]. This indicates that there are no paratropic contributions along the z-profile. Additionally, the ring–current profile shows that the strongest current density flux occurs just below the boron triangle, confirming the three-dimensional delocalization along the z-axis (Figure S6 in Supporting Information). Thus, B3Al4+ is a true 3D aromatic fluxional cluster.

4. Summary

After systematically exploring the potential energy surface of clusters with formula B3Al4+, we confirm that the putative global minimum corresponds to a structure formed by an Al4 quasi-square facing a B3 triangle. Most interestingly, the global minimum of B3Al4+ exhibits fluxionality. However, its dynamic behavior involves the rotation of the B3 triangle on the Al4 ring, embodying a molecular Reuleaux triangle. Corroborated by BOMD, a complete rotation cycle includes twelve repeating steps comprising the global minimum and a transition state separated by only 0.1 kcal/mol, indicating that the rotation between both fragments is virtually free. Furthermore, the AdNDP analysis suggests that the interaction between the B and Al fragments is via three 7c-2e bonds, favoring such dynamical behavior. Regarding aromaticity, the analysis based on the magnetic response reveals a strong shielding along the z-axis due to an entirely diatropic current density flowing between fragments leading to a three-dimensional double (σ + π)-aromaticity. So, we find a peculiar three-dimensional fluxional 3D aromatic cluster that resembles a molecular Reuleaux triangle.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27217407/s1, Figure S1: Twenty low-lying isomers of B3Al4+; Figure S2: Displacement vectors of soft vibrational mode of the global minimum and TS; Figure S3: Computed bond distances, Wiberg bond indices, and natural atomic charges for both the global minimum and TS of B3Al4+; Figure S4: Bonding pattern, according to AdNDP analysis, in the transition state related to the rotation of the B3 fragment in B3Al4+; Figure S5: Bonding model for B3Al4+; Figure S6: The vertical ring-current profile along the z-axis; Tables S1 and S2: Orbital composition analysis for occupied canonical molecular orbitals for both the global minimum and transition state of B3Al4+; Video S1: Molecular dynamics of B3Al4+; Cartesian coordinates of both the global minimum and the transition state.

Author Contributions

Conceptualization, J.-C.G. and G.M.; validation, L.-X.B., S.P. and X.Z.; writing, S.P., M.O.-I., D.S., J.-C.G. and G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (22173053) and the Natural Science Foundation of Shanxi Province (20210302123439). This work was funded in Mexico by Conacyt (Proyecto Sinergia 1561802). The work in Finland was supported by the Academy of Finland (Project numbers 314821 and 340583), the Magnus Ehrnrooth Foundation, Waldemar von Frenckell’s Foundation, and the Swedish Cultural Foundation in Finland.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Huang, W.; Sergeeva, A.P.; Zhai, H.-J.; Averkiev, B.B.; Wang, L.-S.; Boldyrev, A.I. A concentric planar doubly π-aromatic B19 cluster. Nat. Chem. 2010, 2, 202–206. [Google Scholar] [CrossRef] [PubMed]
  2. Jiménez-Halla, J.O.C.; Islas, R.; Heine, T.; Merino, G. B19: An aromatic Wankel motor. Angew. Chem. Int. Ed. Engl. 2010, 49, 5668–5671. [Google Scholar] [CrossRef]
  3. Moreno, D.; Pan, S.; Zeonjuk, L.L.; Islas, R.; Osorio, E.; Martínez-Guajardo, G.; Chattaraj, P.K.; Heine, T.; Merino, G. B182−: A quasi-planar bowl member of the Wankel motor family. Chem. Commun. 2014, 50, 8140–8143. [Google Scholar] [CrossRef]
  4. Wang, Y.-J.; Zhao, X.-Y.; Chen, Q.; Zhai, H.-J.; Li, S.-D. B11: A moving subnanoscale tank tread. Nanoscale 2015, 7, 16054–16060. [Google Scholar] [CrossRef]
  5. Wang, Y.-J.; You, X.-R.; Chen, Q.; Feng, L.-Y.; Wang, K.; Ou, T.; Zhao, X.-Y.; Zhai, H.-J.; Li, S.-D. Chemical bonding and dynamic fluxionality of a B15+ cluster: A nanoscale double-axle tank tread. Phys. Chem. Chem. Phys. 2016, 18, 15774–15782. [Google Scholar] [CrossRef]
  6. Yang, Y.; Jia, D.; Wang, Y.-J.; Zhai, H.-J.; Man, Y.; Li, S.-D. A universal mechanism of the planar boron rotors B11, B13+, B15+, and B19: Inner wheels rotating in pseudo-rotating outer bearings. Nanoscale 2017, 9, 1443–1448. [Google Scholar] [CrossRef]
  7. Liu, L.; Moreno, D.; Osorio, E.; Castro, A.C.; Pan, S.; Chattaraj, P.K.; Heine, T.; Merino, G. Structure and bonding of IrB12: Converting a rigid boron B12 platelet to a Wankel motor. RSC Adv. 2016, 6, 27177–27182. [Google Scholar] [CrossRef]
  8. Wang, Y.-J.; Feng, L.-Y.; Zhai, H.-J. Starting a subnanoscale tank tread: Dynamic fluxionality of boron-based B10Ca alloy cluster. Nanoscale Adv. 2019, 1, 735–745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Martínez-Guajardo, G.; Sergeeva, A.P.; Boldyrev, A.I.; Heine, T.; Ugalde, J.M.; Merino, G. Unravelling phenomenon of internal rotation in B13+ through chemical bonding analysis. Chem. Commun. 2011, 47, 6242–6244. [Google Scholar] [CrossRef] [PubMed]
  10. Merino, G.; Heine, T. And yet it rotates: Thestarter for a molecular Wankel motor. Angew. Chem. Int. Ed. Engl. 2012, 51, 10226–10227. [Google Scholar] [CrossRef]
  11. Fagiani, M.R.; Song, X.; Petkov, P.; Debnath, S.; Gewinner, S.; Schöllkopf, W.; Heine, T.; Fielicke, A.; Asmis, K.R. Structure and fluxionality of B13+ probed by infrared photodissociation spectroscopy. Chem. Int. Ed. Engl. 2017, 56, 501–504. [Google Scholar] [CrossRef]
  12. Pan, S.; Barroso, J.; Jalife, S.; Heine, T.; Asmis, K.R.; Merino, G. Fluxional boron clusters: From theory to reality. Acc. Chem. Res. 2019, 52, 2732–2744. [Google Scholar] [CrossRef]
  13. Jalife, S.; Liu, L.; Pan, S.; Cabellos, J.L.; Osorio, E.; Lu, C.; Heine, T.; Donald, K.J.; Merino, G. Dynamical behavior of boron clusters. Nanoscale 2016, 8, 17639–17644. [Google Scholar] [CrossRef]
  14. Wang, Y.-J.; Feng, L.-Y.; Guo, J.-C.; Zhai, H.-J. Dynamic Mg2B8 cluster: A nanoscale compass. Chem. Asian J. 2017, 12, 2899–2903. [Google Scholar] [CrossRef]
  15. Zhang, X.-Y.; Guo, J.-C. Dynamic fluxionality of ternary Mg2BeB8 cluster: A nanocompass. J. Mol. Model. 2020, 26, 30. [Google Scholar] [CrossRef]
  16. Li, W.-L.; Jian, T.; Chen, X.; Li, H.-R.; Chen, T.-T.; Luo, X.-M.; Li, S.-D.; Li, J.; Wang, L.-S. Observation of a metal-centered B2-Ta@B18 tubular molecular rotor and a perfect Ta@B20 boron drum with the record coordination number of twenty. Chem. Commun. 2017, 53, 1587–1590. [Google Scholar] [CrossRef]
  17. Martínez-Guajardo, G.; Cabellos, J.L.; Díaz-Celaya, A.; Pan, S.; Islas, R.; Chattaraj, P.K.; Heine, T.; Merino, G. Dynamical behavior of borospherene: A nanobubble. Sci. Rep. 2015, 5, 11287. [Google Scholar] [CrossRef] [Green Version]
  18. Guo, J.-C.; Feng, L.-Y.; Wang, Y.-J.; Jalife, S.; Vásquez-Espinal, A.; Cabellos, J.L.; Pan, S.; Merino, G.; Zhai, H.-J. Coaxial triple-layered versus helical Be6B11 clusters: Dual structural fluxionality and multifold aromaticity. Angew. Chem. Int. Ed. Engl. 2017, 56, 10174–10177. [Google Scholar] [CrossRef]
  19. Feng, L.-Y.; Guo, J.-C.; Li, P.-F.; Zhai, H.-J. Boron-based binary Be6B102− cluster: Three-layered aromatic sandwich, electronic transmutation, and dynamic structural fluxionality. Phys. Chem. Chem. Phys. 2018, 20, 22719–22729. [Google Scholar] [CrossRef]
  20. Wang, Y.-J.; Feng, L.-Y.; Zhai, H.-J. Sandwich-type Na6B7 and Na8B7+ clusters: Charge-transfer complexes, four-fold π/σ aromaticity, and dynamic fluxionality. Phys. Chem. Chem. Phys. 2019, 21, 18338–18345. [Google Scholar] [CrossRef]
  21. Barroso, J.; Pan, S.; Merino, G. Structural transformations in boron clusters induced by metal doping. Chem. Soc. Rev. 2022, 51, 1098–1123. [Google Scholar] [CrossRef]
  22. Yu, R.; Barroso, J.; Wang, M.-H.; Liang, W.-Y.; Chen, C.; Zarate, X.; Orozco-Ic, M.; Cui, Z.-H.; Merino, G. Structure and bonding of molecular stirrers with formula B7M2 and B8M2 (M = Zn, Cd, Hg). Phys. Chem. Chem. Phys. 2020, 22, 12312–12320. [Google Scholar]
  23. Wen, L.; Zhou, D.; Yang, L.-M.; Li, G.; Ganz, E. Atomistic structures, stabilities, electronic properties, and chemical bonding of boron–aluminum mixed clusters B3Aln0/−/+ (n = 2–6). J. Cluster Sci. 2021, 32, 1261–1276. [Google Scholar]
  24. Ortiz-Chi, F.; Merino, G. A Hierarchical Algorithm for Molecular Similarity (H-FORMS); GLOMOS: Mérida, Mexico, 2020. [Google Scholar]
  25. Grande-Aztatzi, R.; Martínez-Alanis, P.R.; Cabellos, J.L.; Osorio, E.; Martínez, A.; Merino, G. Structural evolution of small gold clusters doped by one and two boron atoms. J. Comput. Chem. 2014, 35, 2288–2296. [Google Scholar] [CrossRef]
  26. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar]
  27. Dunning, T.H., Jr.; Hay, P.J. Modern Theoretical Chemistry; Schaefer, H.F., III, Ed.; Plenum: New York, NY, USA, 1977; Volume 3, pp. 1–28. [Google Scholar]
  28. Weigend, F.; Ahlrichs, R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar]
  29. Purvis, G.D., III; Bartlett, R.J. A full coupled-cluster singles and doubles model: The inclusion of disconnected triples. J. Chem. Phys. 1982, 76, 1910–1918. [Google Scholar]
  30. Glendening, E.D.; Landis, C.R.; Weinhold, F. NBO 6.0: Natural bond orbital analysis program. J. Comput. Chem. 2013, 34, 1429–1437. [Google Scholar] [CrossRef]
  31. Wiberg, K.B. Application of the pople-santry-segal CNDO method to the cyclopropylcarbinyl and cyclobutyl cation and to bicyclobutane. Tetrahedron 1968, 24, 1083–1096. [Google Scholar]
  32. Millam, J.M.; Bakken, V.; Chen, W.; Hase, W.L.; Schlegel, H.B. Ab initio classical trajectories on the Born-Oppenheimer surface: Hessian-based integrators using fifth-order polynomial and rational function fits. J. Chem. Phys. 1999, 111, 3800–3805. [Google Scholar] [CrossRef]
  33. Jusélius, J.; Sundholm, D.; Gauss, J. Calculation of current densities using gauge-including atomic orbitals. J. Chem. Phys. 2004, 121, 3952–3963. [Google Scholar] [CrossRef] [PubMed]
  34. Fliegl, H.; Taubert, S.; Lehtonen, O.; Sundholm, D. The gauge including magnetically induced current method. Phys. Chem. Chem. Phys. 2011, 13, 20500–20518. [Google Scholar] [CrossRef] [PubMed]
  35. Sundholm, D.; Fliegl, H.; Berger, R.J.F. Calculations of magnetically induced current densities: Theory and applications. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2016, 6, 639–678. [Google Scholar] [CrossRef]
  36. Sundholm, D.; Dimitrova, M.; Berger, R.J.F. Current density and molecular magnetic properties. Chem. Commun. 2021, 57, 12362–12378. [Google Scholar] [CrossRef] [PubMed]
  37. Merino, G.; Heine, T.; Seifert, G. The induced magnetic field in cyclic molecules. Chem.-Eur. J. 2004, 10, 4367–4371. [Google Scholar] [CrossRef] [PubMed]
  38. Heine, T.; Islas, R.; Merino, G. σ and π contributions to the induced magnetic field: Indicators for the mobility of electrons in molecules. J. Comput. Chem. 2007, 28, 302–309. [Google Scholar] [CrossRef] [PubMed]
  39. Islas, R.; Heine, T.; Merino, G. The induced magnetic field. Acc. Chem. Res. 2012, 45, 215–228. [Google Scholar] [CrossRef]
  40. Orozco-Ic, M.; Cabellos, J.L.; Merino, G. Aromagnetic; Cinvestav-Mérida: Mérida, Mexico, 2016. [Google Scholar]
  41. Becke, A.D. A new mixing of Hartree-Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377. [Google Scholar] [CrossRef]
  42. Lehtola, S.; Dimitrova, M.; Fliegl, H.; Sundholm, D. Benchmarking magnetizabilities with recent density functionals. J. Chem. Theory Comput. 2021, 17, 1457–1468. [Google Scholar] [CrossRef]
  43. Ditchfield, R. Self-consistent perturbation theory of diamagnetism. Mol. Phys. 1974, 27, 789–807. [Google Scholar] [CrossRef]
  44. Wolinski, K.; Hinton, J.F.; Pulay, P. Efficient implementation of the gauge-independent atomic orbital method for NMR chemical shift calculations. J. Am. Chem. Soc. 1990, 112, 8251–8260. [Google Scholar] [CrossRef]
  45. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 16, Revision C.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  46. Zubarev, D.Y.; Boldyrev, A.I. Developing paradigms of chemical bonding: Adaptive natural density partitioning. Phys. Chem. Chem. Phys. 2008, 10, 5207–5217. [Google Scholar] [CrossRef] [PubMed]
  47. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  48. Cervantes-Navarro, F.; Martínez-Guajardo, G.; Osorio, E.; Moreno, D.; Tiznado, W.; Islas, R.; Donald, K.J.; Merino, G. Stop rotating! One substitution halts the B19 motor. Chem. Commun. 2014, 50, 10680–10682. [Google Scholar] [CrossRef]
  49. Berger, R.J.F.; Dimitrova, M.; Nasibullin, R.T.; Valiev, R.R.; Sundholm, D. Integration of global ring currents using the Ampère-Maxwell law. Phys. Chem. Chem. Phys. 2022, 24, 624–628. [Google Scholar] [CrossRef]
Figure 1. (a) The Wankel rotor B19 and (b) Three-dimensional molecular Reuleaux triangle B3Al4+.
Figure 1. (a) The Wankel rotor B19 and (b) Three-dimensional molecular Reuleaux triangle B3Al4+.
Molecules 27 07407 g001
Figure 2. PBE0/def2-TZVP structures of the three lowest-lying energy isomers of B3Al4+.
Figure 2. PBE0/def2-TZVP structures of the three lowest-lying energy isomers of B3Al4+.
Molecules 27 07407 g002
Figure 3. Structural evolution of three-dimensional molecular Reuleaux triangle B3Al4+ during the dynamic rotation.
Figure 3. Structural evolution of three-dimensional molecular Reuleaux triangle B3Al4+ during the dynamic rotation.
Molecules 27 07407 g003
Figure 4. The adaptive natural density partitioning (AdNDP) bonding pattern for B3Al4+ (GM). ON is the occupation number.
Figure 4. The adaptive natural density partitioning (AdNDP) bonding pattern for B3Al4+ (GM). ON is the occupation number.
Molecules 27 07407 g004
Figure 5. (a) Bindz isolines plotted in the plane of the Al4 framework (left), the B triangle (middle), and in a transverse plane (right) to B3Al4+. (b) Jind vector maps plotted near B3Al4+. The arrows indicate the direction of the current density. The |Jind| scale is in atomic units (1 au = 100.63 nA/T/Å2). The external magnetic field is oriented parallel to the z-axis, perpendicular to the Al4 plane.
Figure 5. (a) Bindz isolines plotted in the plane of the Al4 framework (left), the B triangle (middle), and in a transverse plane (right) to B3Al4+. (b) Jind vector maps plotted near B3Al4+. The arrows indicate the direction of the current density. The |Jind| scale is in atomic units (1 au = 100.63 nA/T/Å2). The external magnetic field is oriented parallel to the z-axis, perpendicular to the Al4 plane.
Molecules 27 07407 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bai, L.-X.; Orozco-Ic, M.; Zarate, X.; Sundholm, D.; Pan, S.; Guo, J.-C.; Merino, G. B3Al4+: A Three-Dimensional Molecular Reuleaux Triangle. Molecules 2022, 27, 7407. https://doi.org/10.3390/molecules27217407

AMA Style

Bai L-X, Orozco-Ic M, Zarate X, Sundholm D, Pan S, Guo J-C, Merino G. B3Al4+: A Three-Dimensional Molecular Reuleaux Triangle. Molecules. 2022; 27(21):7407. https://doi.org/10.3390/molecules27217407

Chicago/Turabian Style

Bai, Li-Xia, Mesías Orozco-Ic, Ximena Zarate, Dage Sundholm, Sudip Pan, Jin-Chang Guo, and Gabriel Merino. 2022. "B3Al4+: A Three-Dimensional Molecular Reuleaux Triangle" Molecules 27, no. 21: 7407. https://doi.org/10.3390/molecules27217407

Article Metrics

Back to TopTop