Next Article in Journal
Design, Synthesis, and Molecular Docking of New Hydrazide–Hydrazone Derivatives with Imidazole Scaffold as Potential Antimicrobial Agents
Previous Article in Journal
The Inert Pair Effect: An Analysis Using the Chemdex Database
Previous Article in Special Issue
Aromaticity and Antiaromaticity: How to Define Them
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Analysis of Intermolecular Interactions in Cationic π-Stacked Dimer Models of Antiaromatic Molecules

1
Department of Materials Engineering Science, Graduate School of Engineering Science, The University of Osaka, 1-3 Machikaneyama, Toyonaka 560-8531, Osaka, Japan
2
Innovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), The University of Osaka, Suita 565-0871, Osaka, Japan
3
Research Center for Solar Energy Chemistry (RCSEC), Division of Quantum Photochemical Engineering, Graduate School of Engineering Science, The University of Osaka, Toyonaka 560-8531, Osaka, Japan
4
Center for Quantum Information and Quantum Biology (QIQB), The University of Osaka, Toyonaka 560-8531, Osaka, Japan
5
Spintronics Research Network Division, Institute for Open and Transdisciplinary Research Initiatives (OTRI-Spin), The University of Osaka, Toyonaka 560-8531, Osaka, Japan
*
Author to whom correspondence should be addressed.
Chemistry 2025, 7(6), 171; https://doi.org/10.3390/chemistry7060171
Submission received: 15 September 2025 / Revised: 11 October 2025 / Accepted: 20 October 2025 / Published: 23 October 2025

Abstract

We have theoretically examined the intermolecular interactions in the cationic states of π-stacked dimers of 4nπ antiaromatic molecules. The ground state of face-to-face π-dimer models, consisting of cyclobutadienes (CBDs), was analyzed as a function of the stacking distance (d) for their monocationic and dicationic states using multi-reference second-order perturbation theory. Multi-configurational wavefunction analysis in a diabatic representation was employed to understand the electronic structures of the dimer models in terms of the monomer electron configurations. It is found that the monocationic dimer exhibits a local minimum at about d = 2.4 Å in the ground state, where each monomer is represented by a superposition between neutral triplet and cationic doublet electron configurations. Crossing of the ground and excited states occurs through changing d, which is due to the small energy gap between the highest occupied and lowest unoccupied molecular orbitals of antiaromatic molecules.

1. Introduction

Understanding the nature of intermolecular interactions between (anti)aromatic molecular systems has garnered attention for advancing fundamental chemistry and biology, as well as for creating functional molecular materials that exhibit molecular recognition capabilities [1,2,3]. In the field of controlling stimuli-responsive behaviors, tuning the strength and manner of interactions between atomic or molecular entities that change with electron oxidations or reductions is particularly intriguing [4,5,6,7,8,9,10,11,12,13].
It is known that two closed-shell neutral helium atoms do not prefer to form a covalently bonded diatomic molecule, whereas its monocationic state, He2+, can exhibit a bound state with a finite interatomic distance [14,15,16,17]. From a molecular orbital (MO) perspective, this situation can be understood by evaluating the formal bond-order (BO), which is defined as (nbnab)/2, where nb and nab are the occupation numbers of bonding and antibonding MOs, respectively, constructed by the linear combinations of the 1s atomic orbitals (AOs). The formal BO is 0 for He2 since nb = nab = 2, whereas the value is 0.5 for He2+ where nb = 2 and nab = 1 in the ground electron configuration. In the dicationic He22+, electrostatic repulsion between the He+ ions becomes considerable although the formal BO is 1 (nb = 2 and nab = 0).
On the other hand, valence-bond (VB) theory and related methods have provided an intuitive understanding of chemical bonds [18,19,20,21,22]. Pauling investigated the potential energy curves of He2m (m = +1 and +2), based on the resonance theory [15]. A local minimum of the potential energy curve was found not only in He+ but also in He22+, although the potential curve around the minimum is shallow and is overall repulsive in the latter case. Pauling also demonstrated how the resonance of several valence electron configurations contributes to energy stabilization at the local minimum. In the VB theory, although the BO may not be defined as straightforwardly as the MO theory, it still offers insight into the electron configuration of each structural element (namely, an atomic/molecular unit) that is sometimes difficult to understand with canonical MOs delocalized over the entire system. A comprehensive understanding of chemical bonding can be achieved by translating between the MO- and VB-based interpretations of the wavefunction for the system of interest.
In a series of theoretical studies [23,24,25], we have investigated the mechanism of decreasing antiaromatic characters in closely stacked face-to-face dimers of 4nπ molecules [26,27], a phenomenon usually referred to as stacked-ring aromaticity. In addition to the explanations by several pioneering theoretical and experimental studies [26,27,28,29,30,31,32,33,34,35,36,37,38], we proposed an interpretation using a VB-based multi-configurational wavefunction model, that is, stacked-ring aromatic systems can be viewed as assemblies of T1-like Baird-aromatic units interacting with each other [23]. Specifically, an electron promotion from the HOMO to the LUMO, which results in a high-spin T1 configuration, is induced in each monomer by intermolecular interactions. A structural change from the distorted lower symmetry to the higher symmetry of the 4nπ-electron system is also triggered, enabling the HOMO and LUMO of the monomer to become degenerate and mix. Consequently, the 2 × 2 unpaired electrons of T1-like monomers contribute to form multi-center covalent-like double bonds delocalized over several atomic sites, although their energy gain is not expected to be the main factor in dimer formation. Perhaps these interpretations are similar to “promotion and hybridization,” which explain covalent bond formation in CH4 [excitation from the 3P (1s22s22p2) to the 5S (1s22s12p3) states and subsequent s-p atomic orbital mixing]. Recently, we found that a similar change in electronic structure in a 4nπ molecule can also be caused by molecule–surface interactions [39].
From the MO perspective, there is a stacking distance (dc) at which the order of energy levels of the intermolecular bonding and antibonding MOs reverses [27,28,32,40]. As a result, the formal BOs of the neutral and cationic states can differ between the regions d > dc and d < dc (see Figure 1). For example, the formal BOs of the neutral, monocationic, and dicationic states are 0, 0.5, and 1 at d > dc, which is similar to the helium dimer cases, whereas they are 2, 1.5, and 1 at d < dc, with multi-center bonding characters. Such variations in the BO of the dimers of antiaromatic molecules can be used to control molecular aggregation and physicochemical properties [41].
Developing a VB-based model of electronic structures and intermolecular interactions in the cationic states of dimers of antiaromatic molecules is expected to improve the understanding of how aromaticity of the system changes through redox reactions. Therefore, we have attempted to apply VB-based multi-configurational wavefunction models to analyze electronic structures and intermolecular interactions in face-to-face dimer models composed of cyclobutadienes (CBDs). We compared our results with those of the helium dimer cations developed by Pauling and discussed their similarities and differences, because the small HOMO-LUMO gaps of antiaromatic molecules cause mixing of several excitation configurations into the ground state that cannot be observed in the helium cases.

2. Computational Details

2.1. Calculated Models

In this study, we have examined the electronic structures of π-stack dimer models consisting of CBDs. CBD has often been used as a model molecule to extract essential information about antiaromatic and stacked-ring aromatic characters [26,29,30,33,34,36]. First, we optimized the geometry of CBD monomer 1 in the neutral singlet ground state at the strongly contracted n-electron valence state second-order perturbation theory [42,43,44], SC-NEVPT2-CASSCF(4e, 4o)/cc-pVTZ level, under the constraints of D2h and D4h symmetries, which is the same level as that used in the previous study [23]. The single point calculation results of 1 in the neutral singlet, neutral triplet, and cationic doublet states of the D2h and D4h structures are listed in Table S1. Then, we constructed face-to-face π-stack dimer models with various stacking distances d, consisting of CBDs with D2h or D4h symmetry. We call these models 12(D2h) and 12(D4h), respectively (see Figure 2). In this paper, we did not account for structural relaxation effects caused by intermolecular interactions or changes in charge state, meaning each monomer geometry was fixed, and no further geometry optimization was performed to focus only on the electronic structure changes. This is because we aim to focus only on the modulation of π-electronic structures of each monomer in the interacting π-dimers.

2.2. Calculation and Analysis Methods

Based on the geometries of the π-stack dimer models with various d, we conducted single-point calculations in the cationic states, 12(D2h)m and 12(D4h)m (m = +1, +2). The results of the neutral dimer can be found in our previous paper [23]. We performed the state-averaged (SA-)CASSCF calculation, considering the full-π-valence [(8 − m)e, 8o] active space. The ten lowest doublet (m = +1) and singlet (m = +2) states were included in the state-averaging process. Using the reference SA10-CASSCF wavefunctions, single-point calculations at the quasi-degenerate (QD-)SC-NEVPT2 [45] level were conducted to determine the potential energy surface (PES) along the stacking direction for each system. The cc-pVTZ basis set was used throughout the calculation. All the multi-reference calculations were carried out using the Orca 5.0.4 program package [46].
To understand the d-dependent electronic structures of the π-stack dimer models, we performed a diabatization of the multi-configurational wavefunctions that we previously developed for analyzing the neutral systems [23]. The active (pseudo-)canonical molecular orbitals (MOs) obtained from the SA10-CASSCF calculation were rotated to localize the active π-orbitals on each monomer (see Figure 3a). Details of the orbital rotation procedure are available in the Supporting Information and in the previous paper. Then, the QD-NEVPT2 calculation was carried out using the CASCI wavefunctions, constructed from the localized active orbitals, as the reference wavefunctions. This transformation enables us to analyze wavefunctions based on the electron configurations of monomers. It was reported that the S0-T1 gap of the neutral CBD monomer is somewhat underestimated by the NEVPT2 [47]. In the previous paper, we analyzed how this underestimation influences the contributions of T1-like monomer configurations [23].
Figure 3b shows examples of electron configurations in the monocationic doublet dimer models, 12(D2h)+ and 12(D4h)+. Although we have considered the full-π-valence CAS(7e, 8o) active space during multi-reference calculations, here we highlight the key electron configurations arising from the CAS(3e, 4o) subspace, which includes the HOMO and LUMO of each monomer. Here, g and d denote monomer configurations where the HOMO or LUMO is doubly occupied, i.e., the ground and doubly excited configurations, respectively. t refers to the single excitation from the HOMO to the LUMO with the triplet multiplicity. c and c* (a and a*) represent the monocationic (monoanionic) states of the monomer, where the HOMO or LUMO is singly (doubly) occupied, respectively. c2 indicates the dicationic configuration of the monomer.
The electron configurations of the dimer (diabatic states), such as gc, shown in Figure 3b, are constructed as pairs of monomer electron configurations, considering the total spin symmetry. For example, gc (gc*) represents the diabatic state characterized by a pair of monomer configurations, g and c (c*), where one of the monomers takes the neutral ground configuration and the other takes the monocationic configuration. dc (dc*) and tc (tc*) represent the neutral–monocation pair states, where one of the monomers has the doubly excited singlet and triplet configurations, respectively. ac2 and a*c2 represent the dimer configurations composed of monoanionic and dicationic monomer pairs. Each electron configuration of the dimer has a corresponding counterpart, such as gc and cg. Then, the weight of each diabatic electron configuration in the adiabatic state wavefunctions was evaluated.

3. Results and Discussion

3.1. Monocationic π-Dimer Models

Figure 4 shows the calculated results of the state energies and wavefunctions of the monocationic dimer models, 12(D2h)+ and 12(D4h)+. Note that, regarding the total contribution of the linear combination (resonance) between gc and cg, it is represented as “gc” in Figure 4b for simplicity (the same applies to other configurations). First, we examine the results of 12(D2h)+ (the left panel of Figure 4). At a sufficiently large d, it is natural to assume that the gc and cg are the lowest-energy configurations among all. Indeed, the lowest 1B2g state, as described by this contribution at any value of d, is the ground state of 12(D2h)+ for d > 2.7 Å, showing a local minimum at d ~ 3.0 Å. The energy stabilization of the 2B2g state results from the resonance between gc and cg (gc + cg) which is characterized by the transfer integral between the HOMOs of monomers, tH-H. This situation is similar to the He2+ case, where the bound state results from the resonance between two configurations corresponding to gc and cg, associated with the energy splitting between the (gc + cg) and (gccg) states [15]. A similar energy splitting is observed in the present case between the 2B2g and 2B3u states (see Figure S3).
The 2B2u state is an excited state with a large d, mainly described by the tc, where one monomer is in the triplet state and the other is in the cationic ground state. The weight of gc* in the 2B2u state increases as d decreases, associated with a rapid energy stabilization. This energy stabilization would result from the resonance between the tc and c*g (ct and gc*), which is also characterized by tH-H (see Figure 3b). For d < 3.4 Å, where the electron clouds of monomers overlap, the weight of ac2 started to increase slightly. Then, the 1B2u state exhibits a local minimum at d ~ 2.4 Å.
In the case of 12(D4h)+ (the right panel of Figure 4), the doubly degenerate 2B2u and 2B3u states are the ground state, slightly lower than the doubly degenerate 2B2g and 2B3g states (we treated the system as if it had D2h symmetry during the calculation to simplify comparison with 12(D2h)+). Then, the ground state takes a local minimum at d ~2.4 Å. The minimum point of the 2B2u/2B3u states in 12(D4h)+ is similar to that of the 2B2u state in 12(D2h)+. Because the HOMO and LUMO of each monomer are degenerated in the D4h monomer, the ground state of the neutral monomer is described by the linear combination of the configurations g and d with equal weighting, and the energies of the dimer configurations, gc, gc*, dc, and dc*, are identical. Therefore, these configurations describe the 1B2g/1B3g states. The tc and tc* configurations mix in the 2B2g/2B3g states to some extent since the energies of tc and tc* are slightly higher than those of gc, gc*, dc, and dc*. In contrast, the tc and tc* are dominant in the 2B2u/2B3u states. The gc* and dc can mix in the 2B2u/2B3u states because the energies of these configurations are close to those of tc and tc*. In addition, the gc* (dc) can be generated from the ct (c*t) by one-electron transfer (see Figure 3b). A rapid energy stabilization of the 2B2u/2B3u states results from the resonance between tc and c*g (ct and gc*), characterized by tH-H, and that between tc* and cd (c*t and dc), characterized by tL-L (transfer integral between the LUMOs of monomers). Although the number of electron configurations contributing to the 2B2u/2B3u states increases compared to 12(D2h)+, the degree of energy stabilization remains similar: The total energy difference between d = 5.0 Å and d = 2.4 Å (minimum) is −48.92 kcal/mol and −49.72 kcal/mol for 12(D2h)+ and 12(D4h)+, respectively.
The local minimum of the 2B2u/2B3u states in 12(D4h)+ is lower than that of the 2B2u state in 12(D2h)+, indicating that the D4h structure is energetically favored. From these results, in the ground state of the monocationic dimer at sufficiently small d, each monomer is expected to have both neutral triplet and cationic doublet characters as the superposition.

3.2. Dicationic π-Dimer Models

Figure 5 shows the calculated results for the dicationic dimer models, 12(D2h)2+ and 12(D4h)2+. In both cases, the key electron configuration is cc, where each monomer is monocationic (see Figure 5b), and Coulomb repulsion between the cationic monomers is assumed to describe the PESs of these models. Figure 5c shows the ground state energy of the dimer relative to that at d = 100 Å, i.e., ΔE(d) = E(d) − E(d = 100 Å). Note that the reversed order of ΔE(d) between 12(D2h)2+ and 12(D4h)2+ is due to the difference in the monomer energy (see Table S1). We also plot the potential energy (V) between point charges with q = +e (e: elementary charge). From the comparison, the PESs of 12(D2h)2+ and 12(D4h)2+ are modeled well by Coulomb repulsion between the positive charges at large d. However, there is a shallow minimum around d = 2.7 Å. From the wavefunction analysis results (Figure 5b), the contribution of charge-transfer (CT) configurations, that is, neutral–dication pairs, increases at d ≤ 3.4 Å, where the overlap of electron clouds enables electron transfer between the monomers. The electronic coupling between the cc and CT forms is characterized by tH-H (and tL-L as well in the case of 12(D4h)2+), and the resonance between these configurations causes the energy stabilization at small d. A similar result was obtained for the He22+ case by Pauling [15]. However, the PES around the minimum is shallow and is essentially repulsive overall. Therefore, a stable dimer is expected to be difficult to form in the dicationic state.

4. Conclusions

Using the high-precision QD-NEVPT2 calculation combined with a VB-based multi-configurational wavefunction expression, we have theoretically analyzed the electronic structures of the cationic states of face-to-face π-stacked dimer models consisting of CBDs as a function of d. In the monocationic state, two types of states with different characters appear in the ground state depending on the stacking distance d. This situation is similar to the neutral state case demonstrated in our previous study. Within the scope of the considered model, the lowest-energy state is provided by a local minimum around approximately 2.4 Å in the D4h dimer, where the ground state is mainly expressed by the tc and ct configurations. That is, each monomer is represented by a superposition between the neutral triplet and the cationic doublet states.
The ground state of the dicationic state is dominated by the cc configuration at d > 3.4 Å, where each monomer is in the monocationic state, and intermolecular Coulombic repulsion causes a repulsive trend of the PES. On the other hand, in the region where the monomers stack closely together, electron transfer between the monomers becomes possible, leading to an increase in the CT configuration, which stabilizes the energy of the dimer to some extent. The PES near the local minimum is shallow, yet it still exhibits an essentially repulsive PES. While the results for the CBD dimer models have several similarities with those for the helium dimer cases, the state crossing in the monocationic state, attributed to the CBD’s narrow HOMO-LUMO gap, which is comparable to the magnitudes of tH-H and tL-L around dc, is considered unique to stacked systems of antiaromatic molecules.
The insights into electronic configurations obtained in this study are crucial for selecting the appropriate level of approximation required for structural optimization of cationic dimers and magnetic field response calculations. Moving forward, analyzing not only the cationic state but also the photo-accessible electronic excited states is considered a key theme for future work, aiming to pioneer functional switching based on the stacked system of antiaromatic molecules.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry7060171/s1, Figure S1: The active canonical molecular orbitals (CMOs) for the CBD dimer models obtained at the SA10-CASSCF[(8 − m)e, 8o] calculations (a), and the active localized molecular orbitals (LMOs) obtained by localizing the active CMOs using the Pipek–Mezey method (b); Figure S2: Orbital energies of the active canonical orbitals constructed from the HOMOs and LUMOs of monomers calculated at the CASSCF(7e, 8o)/cc-pVTZ for 12(D2h)+; Figure S3: Calculation results of the energy (a) and wavefunction (b) of the 2B3u state of 12(D2h)+ expressed mainly with the anti-symmetric linear combination, gccg; Figure S4: Energies of the diabatic and adiabatic states of 12(D2h)+; Table S1: Total energy (Hartree) of the cyclobutadiene monomer 1 in the neutral singlet, neutral triplet, and cationic doublet states, calculated at the SC-NEVPT2-CASSCF/cc-pVTZ level; Table S2: Cartesian coordinates (Å) of the cyclobutadiene monomer 1.

Author Contributions

Conceptualization, methodology, formal analysis, investigation, K.N., K.O., R.S. and R.K.; writing—original draft preparation, K.N., K.O. and R.K.; writing—review and editing, K.T., R.K. and Y.K.; visualization, K.N. and R.K.; supervision, R.K. and Y.K.; project administration, R.K. and Y.K.; funding acquisition, K.T., R.K. and Y.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by JSPS KAKENHI, grant numbers JP21K04995, JP22H04974, JP22H02050, JP21H05489, JP24H00459, and JP25K08589.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials, further inquiries can be directed to the corresponding authors.

Acknowledgments

Theoretical calculations were partly performed using the Research Center for Computational Science, Okazaki, Japan (Projects: 25-IMS-C004).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Salonen, L.M.; Ellermann, M.; Diederich, F. Aromatic Rings in Chemical and Biological Recognition: Energetics and Structures. Angew. Chem. Int. Ed. 2011, 50, 4808–4842. [Google Scholar] [CrossRef]
  2. Tsuzuki, S. Interactions with Aromatic Rings. In Intermolecular Forces and Clusters I; Wales, D.J., Ed.; Springer: Berlin/Heidelberg, Germany, 2005; pp. 149–193. [Google Scholar]
  3. Ninković, D.B.; Blagojević Filipović, J.P.; Hall, M.B.; Brothers, E.N.; Zarić, S.D. What Is Special about Aromatic–Aromatic Interactions? Significant Attraction at Large Horizontal Displacement. ACS Cent. Sci. 2020, 6, 420–425. [Google Scholar] [CrossRef]
  4. Bhattacharjee, R.; Lischka, H.; Kertesz, M. Pancake Bonding in the Stabilization of Cationic Acene Dimers. ACS Mater. Au 2025, 5, 365–376. [Google Scholar] [CrossRef]
  5. Matsumoto, M.; Inokuchi, Y.; Ohashi, K.; Nishi, N. Charge Delocalization in Benzene−Naphthalene Hetero-Dimer Cation. J. Phys. Chem. A 1997, 101, 4574–4578. [Google Scholar] [CrossRef]
  6. Martinez, C.R.; Iverson, B.L. Rethinking the Term “Pi-Stacking”. Chem. Sci. 2012, 3, 2191–2201. [Google Scholar] [CrossRef]
  7. Badger, B.; Brocklehurst, B. Formation of Dimer Cations of Aromatic Hydrocarbons. Nature 1968, 219, 263. [Google Scholar] [CrossRef]
  8. Meot-Ner, M. Dimer Cations of Polycyclic Aromatics. Experimental Bonding Energies and Resonance Stabilization. J. Phys. Chem. 1980, 84, 2724–2728. [Google Scholar] [CrossRef]
  9. Hübler, P.; Heinze, J. Kinetic Studies on Reversible Dimerization of Thianthrene and 2,3,7,8-Tetramethoxythianthrene Radical Cations. Berichte Bunsenges. Für Phys. Chem. 1998, 102, 1506–1509. [Google Scholar] [CrossRef]
  10. Small, D.; Zaitsev, V.; Jung, Y.; Rosokha, S.V.; Head-Gordon, M.; Kochi, J.K. Intermolecular π-to-π Bonding between Stacked Aromatic Dyads. Experimental and Theoretical Binding Energies and Near-IR Optical Transitions for Phenalenyl Radical/Radical versus Radical/Cation Dimerizations. J. Am. Chem. Soc. 2004, 126, 13850–13858. [Google Scholar] [CrossRef]
  11. Rhee, Y.M.; Lee, T.J.; Gudipati, M.S.; Allamandola, L.J.; Head-Gordon, M. Charged Polycyclic Aromatic Hydrocarbon Clusters and the Galactic Extended Red Emission. Proc. Natl. Acad. Sci. USA 2007, 104, 5274–5278. [Google Scholar] [CrossRef] [PubMed]
  12. Kochi, J.K.; Rathore, R.; Maguères, P.L. Stable Dimeric Aromatic Cation−Radicals. Structural and Spectral Characterization of Through-Space Charge Delocalization. J. Org. Chem. 2000, 65, 6826–6836. [Google Scholar] [CrossRef]
  13. Attah, I.K.; Platt, S.P.; Meot-Ner (Mautner), M.; El-Shall, M.S.; Peverati, R.; Head-Gordon, M. What Is the Structure of the Naphthalene–Benzene Heterodimer Radical Cation? Binding Energy, Charge Delocalization, and Unexpected Charge-Transfer Interaction in Stacked Dimer and Trimer Radical Cations. J. Phys. Chem. Lett. 2015, 6, 1111–1118. [Google Scholar] [CrossRef] [PubMed]
  14. Slater, J.C. The Normal State of Helium. Phys. Rev. 1928, 32, 349–360. [Google Scholar] [CrossRef]
  15. Pauling, L. The Normal State of the Helium Molecule-Ions He2+ and He2++. J. Chem. Phys. 1933, 1, 56–59. [Google Scholar] [CrossRef]
  16. Guilhaus, M.; Brenton, A.G.; Beynon, J.H.; Rabrenović, M.; von Ragué Schleyer, P. He22+, the Experimental Detection of a Remarkable Molecule. J. Chem. Soc. Chem. Commun. 1985, 4, 210–211. [Google Scholar] [CrossRef]
  17. Callear, A.B.; Hedges, R.E.M. Metastability of Rotationally Hot Dihelium at 77° K. Nature 1967, 215, 1267–1268. [Google Scholar] [CrossRef]
  18. Pauling, L. The Nature of the Chemical Bond. Application of Results Obtained From the Quantum Mechanics and From a Theory of Paramagnetic Susceptibility to the Structure of Molecules. J. Am. Chem. Soc. 1931, 53, 1367–1400. [Google Scholar] [CrossRef]
  19. Harris, M.L. Chemical Reductionism Revisited: Lewis, Pauling and the Physico-Chemical Nature of the Chemical Bond. Stud. Hist. Philos. Sci. Part A 2008, 39, 78–90. [Google Scholar] [CrossRef]
  20. Wu, W.; Su, P.; Shaik, S.; Hiberty, P.C. Classical Valence Bond Approach by Modern Methods. Chem. Rev. 2011, 111, 7557–7593. [Google Scholar] [CrossRef]
  21. Shaik, S.; Hiberty, P.C. Valence Bond Theory, Its History, Fundamentals, and Applications: A Primer. In Reviews in Computational Chemistry; John Wiley & Sons, Inc.: Hoboken, NY, USA, 2004; pp. 1–100. [Google Scholar]
  22. McWeeny, R. The Valence Bond Theory of Molecular Structure—I. Orbital Theories and the Valence-Bond Method. Proc. Roy. Soc. Lond. A Math. Phys. Sci. 1997, 223, 63–79. [Google Scholar]
  23. Sugimori, R.; Okada, K.; Kishi, R.; Kitagawa, Y. Stacked-Ring Aromaticity from the Viewpoint of the Effective Number of π-Electrons. Chem. Sci. 2025, 16, 1707–1715. [Google Scholar] [CrossRef]
  24. Sugimori, R.; Kishi, R.; Shinokubo, H.; Kitagawa, Y. Diradical Character View of Energy Stabilization in π-Stacked Dimers of Antiaromatic Molecules. Chem. Lett. 2024, 53, upae224. [Google Scholar] [CrossRef]
  25. Sugimori, R.; Ikeuchi, M.; Kishi, R.; Kitagawa, Y. Theoretical Study on Aromatic Characters in π-Stacked Multimers Composed of Antiaromatic Molecules. Bull. Chem. Soc. Jpn. 2024, 97, uoae133. [Google Scholar] [CrossRef]
  26. Corminboeuf, C.; von Ragué Schleyer, P.; Warner, P. Are Antiaromatic Rings Stacked Face-to-Face Aromatic? Org. Lett. 2007, 9, 3263–3266. [Google Scholar] [CrossRef]
  27. Bean, D.E.; Fowler, P.W. Stacked-Ring Aromaticity: An Orbital Model. Org. Lett. 2008, 10, 5573–5576. [Google Scholar] [CrossRef]
  28. Wang, Q.; Sundholm, D.; Gauss, J.; Nottoli, T.; Lipparini, F.; Kino, S.; Ukai, S.; Fukui, N.; Shinokubo, H. Changing Aromatic Properties through Stacking: The Face-to-Face Dimer of Ni(Ii) Bis(Pentafluorophenyl)Norcorrole. Phys. Chem. Chem. Phys. 2024, 26, 14777–14786. [Google Scholar] [CrossRef]
  29. Alonso, M.; Poater, J.; Solà, M. Aromaticity Changes along the Reaction Coordinate Connecting the Cyclobutadiene Dimer to Cubane and the Benzene Dimer to Hexaprismane. Struct. Chem. 2007, 18, 773–783. [Google Scholar] [CrossRef]
  30. Okazawa, K.; Tsuji, Y.; Yoshizawa, K. Frontier Orbital Views of Stacked Aromaticity. J. Phys. Chem. A 2023, 127, 4780–4786. [Google Scholar] [CrossRef] [PubMed]
  31. Orozco-Ic, M.; Restrepo, A.; Muñoz-Castro, A.; Merino, G. Molecular Helmholtz Coils. J. Chem. Phys. 2019, 151, 014102. [Google Scholar] [CrossRef] [PubMed]
  32. Kino, S.; Ukai, S.; Fukui, N.; Haruki, R.; Kumai, R.; Wang, Q.; Horike, S.; Phung, Q.M.; Sundholm, D.; Shinokubo, H. Close Stacking of Antiaromatic Ni(II) Norcorrole Originating from a Four-Electron Multicentered Bonding Interaction. J. Am. Chem. Soc. 2024, 146, 9311–9317. [Google Scholar] [CrossRef]
  33. Tsuji, Y.; Okazawa, K.; Yoshizawa, K. Hückel Molecular Orbital Analysis for Stability and Instability of Stacked Aromatic and Stacked Antiaromatic Systems. J. Org. Chem. 2023, 88, 14887–14898. [Google Scholar] [CrossRef]
  34. Orozco-Ic, M.; Sundholm, D. On the Antiaromatic–Aromatic–Antiaromatic Transition of the Stacked Cyclobutadiene Dimer. Phys. Chem. Chem. Phys. 2023, 25, 12777–12782. [Google Scholar] [CrossRef]
  35. Nozawa, R.; Tanaka, H.; Cha, W.-Y.; Hong, Y.; Hisaki, I.; Shimizu, S.; Shin, J.-Y.; Kowalczyk, T.; Irle, S.; Kim, D.; et al. Stacked Antiaromatic Porphyrins. Nat. Commun. 2016, 7, 13620. [Google Scholar] [CrossRef]
  36. Aihara, J. Origin of Stacked-Ring Aromaticity. J. Phys. Chem. A 2009, 113, 7945–7952. [Google Scholar] [CrossRef]
  37. Kawashima, H.; Ukai, S.; Nozawa, R.; Fukui, N.; Fitzsimmons, G.; Kowalczyk, T.; Fliegl, H.; Shinokubo, H. Determinant Factors of Three-Dimensional Aromaticity in Antiaromatic Cyclophanes. J. Am. Chem. Soc. 2021, 143, 10676–10685. [Google Scholar] [CrossRef] [PubMed]
  38. Nozawa, R.; Kim, J.; Oh, J.; Lamping, A.; Wang, Y.; Shimizu, S.; Hisaki, I.; Kowalczyk, T.; Fliegl, H.; Kim, D.; et al. Three-Dimensional Aromaticity in an Antiaromatic Cyclophane. Nat. Commun. 2019, 10, 3576. [Google Scholar] [CrossRef]
  39. Tada, K.; Sugimori, R.; Kishi, R.; Kitagawa, Y. Unveiled Open-Shell Nature of an Antiaromatic Molecule by Surface–Molecule Interaction: A Concept Suggestion by Spin-Projected Density Functional Theory. Chem. Commun. 2025, 61, 11758–11761. [Google Scholar] [CrossRef]
  40. Fujiyoshi, J.-Y.; Tonami, T.; Yamane, M.; Okada, K.; Kishi, R.; Muhammad, S.; Al-Sehemi, A.G.; Nozawa, R.; Shinokubo, H.; Nakano, M. Theoretical Study on Open-Shell Singlet Character and Second Hyperpolarizabilities in Cofacial π-Stacked Dimers Composed of Weak Open-Shell Antiaromatic Porphyrins. ChemPhysChem 2018, 19, 2863–2871. [Google Scholar] [CrossRef] [PubMed]
  41. Joseph, J.; Berville, M.; Wytko, J.; Weiss, J.; Jacquot de Rouville, H.-P. π-Mers and π-Dimers: Two Radical Supramolecular Interactions—A Tutorial Review. Chem. Eur. J. 2025, 31, e202403115. [Google Scholar] [CrossRef]
  42. Angeli, C.; Cimiraglia, R.; Evangelisti, S.; Leininger, T.; Malrieu, J.-P. Introduction of N-Electron Valence States for Multireference Perturbation Theory. J. Chem. Phys. 2001, 114, 10252–10264. [Google Scholar] [CrossRef]
  43. Angeli, C.; Cimiraglia, R.; Malrieu, J.-P. N-Electron Valence State Perturbation Theory: A Fast Implementation of the Strongly Contracted Variant. Chem. Phys. Lett. 2001, 350, 297–305. [Google Scholar] [CrossRef]
  44. Angeli, C.; Cimiraglia, R.; Malrieu, J.-P. N-Electron Valence State Perturbation Theory: A Spinless Formulation and an Efficient Implementation of the Strongly Contracted and of the Partially Contracted Variants. J. Chem. Phys. 2002, 117, 9138–9153. [Google Scholar] [CrossRef]
  45. Angeli, C.; Borini, S.; Cestari, M.; Cimiraglia, R. A Quasidegenerate Formulation of the Second Order N-Electron Valence State Perturbation Theory Approach. J. Chem. Phys. 2004, 121, 4043–4049. [Google Scholar] [CrossRef] [PubMed]
  46. Neese, F. Software Update: The ORCA Program System—Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606. [Google Scholar] [CrossRef]
  47. Monino, E.; Boggio-Pasqua, M.; Scemama, A.; Jacquemin, D.; Loos, P.-F. Reference Energies for Cyclobutadiene: Automerization and Excited States. J. Phys. Chem. A 2022, 126, 4664–4679. [Google Scholar] [CrossRef]
Figure 1. MO and VB-based interpretations of intermolecular interactions between 4nπ antiaromatic molecules in the neutral and cationic states. The energy splittings between the bonding and anti-bonding orbitals consisting of the HOMOs (~2|tH-H|) or the LUMOs (~2|tL-L|) are indicated. The light- and dark-gray arrows in each electron configuration indicate the electrons that are removed in the cationic states.
Figure 1. MO and VB-based interpretations of intermolecular interactions between 4nπ antiaromatic molecules in the neutral and cationic states. The energy splittings between the bonding and anti-bonding orbitals consisting of the HOMOs (~2|tH-H|) or the LUMOs (~2|tL-L|) are indicated. The light- and dark-gray arrows in each electron configuration indicate the electrons that are removed in the cationic states.
Chemistry 07 00171 g001
Figure 2. Structures of π-dimer models, 12(D2h) and 12(D4h).
Figure 2. Structures of π-dimer models, 12(D2h) and 12(D4h).
Chemistry 07 00171 g002
Figure 3. Localized active MOs constructed from the HOMOs and LUMOs of monomers (a) and key electron configurations within the CAS(3e, 4o) subspace (b). g, d, and t denote the neutral ground singlet, doubly excited singlet, and singly excited triplet configurations, c and c* (a and a*) represent the ground and excited monocationic (monoanionic) configurations, and c2 indicates the dicationic configuration of the monomer part, respectively. The electronic structure corresponding to each monomer configuration is schematically shown below the configuration diagram, where the red asterisks (*) and (**) indicate the singly and doubly excited states. The diabatic electron configurations of the dimer are represented as pairs of the configurations of monomers A and B. Red- and blue-colored arrows indicate the possible one-electron transfer coupling between the HOMOs (tH-H) or between the LUMOs (tL-L).
Figure 3. Localized active MOs constructed from the HOMOs and LUMOs of monomers (a) and key electron configurations within the CAS(3e, 4o) subspace (b). g, d, and t denote the neutral ground singlet, doubly excited singlet, and singly excited triplet configurations, c and c* (a and a*) represent the ground and excited monocationic (monoanionic) configurations, and c2 indicates the dicationic configuration of the monomer part, respectively. The electronic structure corresponding to each monomer configuration is schematically shown below the configuration diagram, where the red asterisks (*) and (**) indicate the singly and doubly excited states. The diabatic electron configurations of the dimer are represented as pairs of the configurations of monomers A and B. Red- and blue-colored arrows indicate the possible one-electron transfer coupling between the HOMOs (tH-H) or between the LUMOs (tL-L).
Chemistry 07 00171 g003
Figure 4. Calculated energies (a) and electron configurations (b) for 12(D2h)+ and 12(D4h)+ models. The total contribution of the linear combination (resonance) between gc and cg, it is represented as “gc” in (b) for simplicity (the same applies to other configurations).
Figure 4. Calculated energies (a) and electron configurations (b) for 12(D2h)+ and 12(D4h)+ models. The total contribution of the linear combination (resonance) between gc and cg, it is represented as “gc” in (b) for simplicity (the same applies to other configurations).
Chemistry 07 00171 g004
Figure 5. Calculated energies (a), electron configurations (b), and the ΔE(d) = E(d) − E(d = 100 Å) (c) for 12(D2h)2+ and 12(D4h)2+ models. The dotted line in (c) represents the potential energy, V, calculated from the point-charge model with q = +e.
Figure 5. Calculated energies (a), electron configurations (b), and the ΔE(d) = E(d) − E(d = 100 Å) (c) for 12(D2h)2+ and 12(D4h)2+ models. The dotted line in (c) represents the potential energy, V, calculated from the point-charge model with q = +e.
Chemistry 07 00171 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nishino, K.; Okada, K.; Sugimori, R.; Tada, K.; Kishi, R.; Kitagawa, Y. Theoretical Analysis of Intermolecular Interactions in Cationic π-Stacked Dimer Models of Antiaromatic Molecules. Chemistry 2025, 7, 171. https://doi.org/10.3390/chemistry7060171

AMA Style

Nishino K, Okada K, Sugimori R, Tada K, Kishi R, Kitagawa Y. Theoretical Analysis of Intermolecular Interactions in Cationic π-Stacked Dimer Models of Antiaromatic Molecules. Chemistry. 2025; 7(6):171. https://doi.org/10.3390/chemistry7060171

Chicago/Turabian Style

Nishino, Kosei, Kenji Okada, Ryota Sugimori, Kohei Tada, Ryohei Kishi, and Yasutaka Kitagawa. 2025. "Theoretical Analysis of Intermolecular Interactions in Cationic π-Stacked Dimer Models of Antiaromatic Molecules" Chemistry 7, no. 6: 171. https://doi.org/10.3390/chemistry7060171

APA Style

Nishino, K., Okada, K., Sugimori, R., Tada, K., Kishi, R., & Kitagawa, Y. (2025). Theoretical Analysis of Intermolecular Interactions in Cationic π-Stacked Dimer Models of Antiaromatic Molecules. Chemistry, 7(6), 171. https://doi.org/10.3390/chemistry7060171

Article Metrics

Back to TopTop