Embedding a guest gold cluster in an organic host. Evaluation of the encapsulation nature in a Au18–superphane host–guest aggregate
Received
27th May 2025
, Accepted 11th September 2025
First published on 15th September 2025
Abstract
Formation of supramolecular aggregates incorporating Au18 into a suitable bioinspired polyfunctional superphane cavity provides novel functionality to the overall structure. We evaluated the favorable incorporation of the Au18 cluster into the superphane cavity. This amounted to −145.3 kcal mol−1, provided mainly by electrostatic-type interactions (54.9%). Charge transfer characteristics involving host ← guest and host → guest backdonation through S ← Au and S → Au contacts led to overall Au18 → 1 superphane charge transfer. Charge transfer consisted of a charge hopping rate (kCT) in the range of ultrafast electron transfer, calculated to be 2.2 × 1013 s−1 at 300 K. Thus, Au18 → 1 charge transfer was driven by coordinating and short contacts towards the superphane available cavity, resulting in a supramolecular structure of the donor–acceptor (D–A) system. We expect that the current approach can be useful for further rationalizing the relevant stabilizing factors to ensure the stable aggregation of metallic clusters in organic host cavities during the making of novel functional cluster-based host–guest aggregates.
Introduction
A supramolecular assembly represents a useful strategy to design and achieve multifunctional aggregate materials with applications in a wide range of fields,1–6 highlighting the use of non-covalent interactions in creating ordered architectures. Host–guest chemistry takes advantage of such interactions by including guest units into a suitable host cavity based on the mutual recognition between molecular constituents,7–17 largely exemplified by involving organic pairs.
Recently, Zhang and coworkers reported the ferritin-inspired design,18 characterizations and performance of a host–guest structure of a bare gold nanocluster (AuNC) embedded into a highly polyfunctional superphane cavity.19 It involved several coordination sites given by nitrogen, oxygen and sulfur atoms from imine, BINOL (BINOL = 1,1-binaphthyl-2,2 diol) dimethyl ether,20 and thiophene groups. The resulting superphane appeared as a prototypical organic cavity with multiple coordinating sites serving as guidance for further development of host structures prone to incorporate bare clusters, favoring a controlled synthesis, size-selectivity purification, solubility in non-polar solvents, incorporation into organic electronic devices, among other issues relevant for emergent applications of atomically precise metal nanoclusters.21–33
The AuNC–superphane host–guest pair by Zhang and colleagues was obtained by reacting the hollow superphane with AuCl3 in CH2Cl2, followed by the addition of NaBH3CN as a mild reducing agent,19 and confirmation of the inclusion of the atomically-precise Au18 bare cluster via electrospray ionization-mass spectrometry (ESI-MS). The resulting organic–inorganic host–guest structure from AuNC–superphane gives rise to a supramolecular hybrid system that can modify the inherent physical–chemical characteristics of each constituent significantly,19 as observed for different hybrid assemblies.34–41 In particular, this aggregate enhances the stability of AuNCs, providing unexpected functionality as given by a broad absorption, improving the sunlight absorption capabilities, with a promising photothermal conversion efficiency of 92.8% desired for solar-to-vapor generation.19
Interestingly, the Au18 cluster features 18 cluster electrons (18-ce) fulfilling an electronic shell order in analogy to isolated atoms42 ascribed by 1S21P61D10, in line with the superatom approach of molecular clusters providing chemical stability.43–45 Theoretical calculations on the structure of the Au18–superphane species19 have denoted a favorable match between the available host cavity and guest size, showing coordination mainly ascribed to the thiophene and BINOL sites.
Herein, we rationalized the nature of the superphane–Au18 interaction to further clarify the stabilizing factors that provide efficient cluster encapsulation into the available organic cavity. The intermolecular interaction between Au18 and the superphane host cavity was evaluated by energy decomposition analysis (EDA),46,47 electrostatic potential maps,48–50 electron density difference maps, and non-covalent index (NCI) analysis51,52 to reveal the contributing role of the different constituent sections of the organic cavity. Moreover, the charge transfer of electrons was evaluated within the Marcus theory framework to account for the reorganization energy (λ) and electronic coupling (V) involved in the processes determining the charge hopping rate in the resulting Au18–superphane aggregate. This was done to further explore the charge transfer parameters in metallic clusters embedded in suitable organic cavities.
Computational details
Calculations were done using the ADF2024 code.53 We used the triple-ξ and two polarization functions (STO-TZ2P) basis set within the generalized gradient approximation (GGA) according to the BP86 exchange–correlation functional and the empirical dispersion correction to DFT (DFT-D) given by the pairwise Grimme correction (D3)54–57 and Becke–Johnson damping functions.58,59 Dispersion-corrected DFT methods offer reliable results and improved performance in the description of supramolecular interactions at a moderate computational cost for larger systems.55,60,61 Relaxed structures were obtained through the analytical energy gradient method implemented by Versluis and Ziegler62 at the TZ2P/BP86-D3 level without any symmetry restrain. The energy convergence criterion was set to 10−5 Hartree, gradient convergence criteria to 10−4 Hartree per Å, and radial convergence criteria to 10−3 Å to achieve final optimized structures. The counterpoise correction scheme was employed to overcome basis set superposition error (BSSE) in the interaction energy analysis owing to the systematic error introduced by the use of finite basis sets, overbinding van der Waals aggregates.63–65 Solvent effects were taken into account via the conductor-like screening model (COSMO) for explicit solvation using dichloromethane (CH2Cl2) as the solvent, as implemented in the ADF code.66,67 The charge hopping rate (kCT) in the Au18–superphane aggregate was modeled using the high-temperature limit of the Marcus theory,68–72 as implemented in the ADFcode.
Molecular dynamics trajectories were obtained via eXtended tight binding (xTB) methods at the GFN0-xTB level as implemented in the standalone xtb code version.73 The temperature was set to 298.15 K for thermostatistical evaluation, whereby hydrogens were treated as deuterium atoms with an accuracy set to 2.0.
Results and discussion
The structure of Au18–superphane has been provided computationally by Zhang and coworkers.19 It denotes the suitable incorporation of the Au18 cluster into the organic host cavity. The thiophene–BINOL-based superphane host (1)19 has been inspired by the unique binding pockets from ferritin iron storage proteins,18 leading to the encapsulation and stabilization of metal ions or clusters, thereby ensuring long-term stability. The resulting host–guest structure from Au18 and 1 involves several coordinating sites retaining the gold nanoparticle (Scheme 1). This offers an interesting case of an amenable host cavity to evaluate the role of different stabilizing contributions resulting in efficient aggregation. This approach led to the characterized Au18@1 as probed by ESI-MS, UV-vis spectroscopy, CD spectroscopy, powder X-ray diffraction (PXRD), and X-ray photoelectron spectroscopy (XPS), among other techniques.19
 |
| Scheme 1 A schematic representation of the thiophene–BINOL-based superphane host (1) and its host–guest complex with Au18 (Au18@1). | |
Our calculations revealed a similar structure to that reported previously.19 We documented shorter Au18 superphane bond lengths in the range 2.406–2.517 Å involving Au–S coordinating bonds from thiophene groups, and Au–O bonds in the range 2.466–3.393 Å from methyl ether groups from methylated BINOL, ascribed mainly to the upper and central sections of the organic cavity, respectively (Fig. 1c), locating the Au18 at one side of the available cavity. Such sections featured the main coordinating sites from the host cage interacting towards the bare Au18 cluster, with complementary Au⋯H3C- and Au⋯H–thiophene contacts (Fig. 1d). For comparison, the central disposition of the Au18 cluster into 1 was evaluated, which was disfavored by 18.6 kcal mol−1.
 |
| Fig. 1 Side (a) and upper (b) views of Au18@1, denoting the location of methyl ether groups from methylated BINOL (c) and Au⋯H3C- and Au⋯H–thiophene contacts (d). The available internal cavity size is given in green from MoloVol suite74 calculations (e). | |
The available cavity size in 1 was evaluated by MoloVol suite74 employing two spherical probes to define the available cavities and the related surfaces and volumes. This evaluation led to an inner cavity in 1 of 731.84 Å3 (Fig. 1d and e), which was very suitable for incorporating the Au18 structure with a volume of 552.98 Å3.
The embedded Au18 cluster features 18 cluster electrons as provided by the respective set of 6s1 atomic shells,44,75 building up an electronic shell resembling atomic orbitals fulfilling an 1S21P61D10 order, ascribed as a superatom, denoting particular stability.76–79 The 1D10 shell contributed to the formation of frontier orbitals in the overall Au18@1 aggregate (Fig. S1, SI). This superatom cluster showed a distorted structure, which was located 34.6 kcal mol−1 above the preferred isomer (Fig. S2, SI) observed from infrared multiple photon dissociation (IR-MPD) experiments.80 Hence, the Au18 cluster could modify its structure to maximize the interaction towards the organic cavity, retaining a similar electronic structure. We wished to evaluate the interaction energy leading to the Au18–superphane host–guest aggregate. Hence, the interaction between the Au18 unit and superphane structure 1 was calculated, resulting in a sizable stabilization of −145.3 kcal mol−1 (Table 1).81–83 To bring host and guest units from their isolated relaxed structures to their structure in the resulting host–guest system, the involved geometric and electronic destabilizing deformation is accounted for by the preparation energy ΔEprep.81–84 The estimated ΔEprep amounted to 34.6 kcal mol−1 for Au18 and 41.8 kcal mol−1 for the superphane cage, leading to an overall ΔEprep of 76.4 kcal mol−1. Hence, the structure could overcome the required structural modification to give rise to the host–guest aggregate.
Table 1 Energy decomposition analysis for the Au18–1 interaction and for related models of 1. Values are in kcal mol−1. Percentage contributions for stabilizing terms are provided
|
Au18–1 |
|
Au18–(C4H4S)4 |
|
Au18–(OMe2)12 |
|
ΔEprep |
76.4 |
|
|
|
|
|
ΔEPauli |
783.8 |
|
387.2 |
|
301.46 |
|
ΔEelstat |
−510.4 |
54.9% |
−263.3 |
59.7% |
−188.8 |
52.3% |
ΔEorb |
−277.7 |
29.9% |
−153.8 |
34.9% |
−112.9 |
31.3% |
ΔEdisp |
−140.8 |
15.2% |
−24.1 |
5.5% |
−59.5 |
16.5% |
ΔEint |
−145.3 |
|
−53.9 |
|
−59.7 |
|
The characteristics of the stabilizing host–guest interaction waws further unraveled by the role of different contributing terms to the resulting interaction energy (ΔEint) evaluated via the energy decomposition analysis (EDA) given by Ziegler and Rauk,46,85,86 according to eqn (1):
|
ΔEint = ΔEPauli + ΔEelstat + ΔEorb + ΔEdisp
| (1) |
In this framework, the Pauli repulsion (ΔEPauli) results from the four electrons/two orbitals between occupied orbitals from Au18 and the superphane cavity, reflecting the steric effect associated with the interaction,87 which amounted to 783.8 kcal mol−1 (Table 1). Moreover, the stabilizing electronic part of the interaction involving electrostatic (ΔEelstat) and orbital (ΔEorb) terms accounts for the Coulomb interaction between the charge densities (ΔEelstat) and polarization and charge transfer contribution after relaxing the orbitals (ΔEorb) to the in the host–guest system.88 The dispersion interaction (ΔEdisp) was evaluated via the pairwise correction of Grimme (DFT-D3).57 The ΔEelstat and ΔEorb terms amounted to −510.4 and −277.7 kcal mol−1, respectively, complemented with the ΔEdisp term of −140.8 kcal mol−1, which could overcome the Pauli repulsion. The percentage relationship between the stabilizing terms (ΔEorb, ΔEelstat, and ΔEdisp) characterize the overall nature of the host–guest interaction, which was of mainly electrostatic character (54.9%), followed by an orbital contribution of 29.9% and, lastly, 15.2% from the dispersion interaction.
To reveal the spatial distribution of the main electrostatic interaction accounting for the ΔEelstat term, the charge reorganization at the van der Waals surface for Au18 was obtained by representing the electrostatic potential48 over an electron density surface of 0.001 electrons per Bohr3.89,90 The electrostatic potential for the isolated Au18 guest unit showed the formation of charge depletion regions as Lewis acidic sites at the low connected Au atoms (Fig. 2a) as the maxima in the surface electrostatic potential (VS,max), similar to that obtained for the Au13 cluster,50 denoted as σ-hole regions accounting for reactive sites in metallic clusters.50,91–93 Interestingly, the electrostatic potential for the overall Au18–superphane structure showed charge reorganization over the van der Waals Au18 surface at the Au–S and Au–O coordinating sites and also for Au⋯H3C- and Au⋯H–thiophene contacts (Fig. 2b). These data showed that the stabilizing ΔEelstat term was given by the contribution from different complementary sites within the superphane cavity contributing to encapsulation of the Au18 bare superatom.
 |
| Fig. 2 The electrostatic potential surface for the isolated Au18 cluster (a) and overall Au18@1 (b) aggregate drawn at a 0.001 electrons per Bohr3 electron density for the Au18 cluster (including the 1 host structure for graphical guidance). Blue denotes electro-positive sites prone to interact with Lewis bases. | |
Moreover, ΔEorb can be evaluated in terms of individual deformation density channels accounting for individual bonding contributions94,95 via the natural orbitals for chemical valence96–98 extension of EDA (EDA-NOCV).98 We documented sixteen main individual deformation density channels (Δρ1–Δρ16) (Fig. S3, SI) contributing between −20.0 to −5.2 kcal mol−1 (Table S1, SI). These data suggested host ← guest charge transfer through S ← Au contacts (Δρ1–Δρ4) (Fig. 3a), and host → guest charge transfer via S → Au contacts. These results suggested a donation and backdonation of charge leading to an overall Au18 → 1 superphane charge transfer of +0.79e as obtained from Hirshfeld charge analyses. The spatial distribution of the resulting charge transfer was denoted by the difference in electron density between the host–guest aggregate and respective fragments (Δρ(r) = ρ(r)total − (ρ(r)Au18 + ρ(r)superphane)) (Fig. 3b). These data suggested that charge accumulation remained at the host cavity near the Au18 cluster, with charge depletion at the S–Au coordinating contacts. Thus, the Au18 → 1 charge transfer was driven by coordinating and short contacts towards the superphane available cavity, resulting in a supramolecular donor–acceptor (D–A) system structure. In addition, TZ2P/PBE-D3 and TZ2P/B3LYP-D3 levels of theory were evaluated, which delivered similar features than for those calculated at the TZ2P/BP86-D3 level, supporting the calculated data at different computational levels (Table S2, SI). The obtained Au18–1 interaction energy amounted to −133.6 and −155.5 kcal mol−1 for TZ2P/PBE-D3 and TZ2P/B3LYP-D3 levels, respectively, in line with that calculated at TZ2P/BP86-D3 (−145.2 kcal mol−1). In addition, different levels of theory can be employed to gain insights into the characteristics of the Au18 inclusion into 1.99
 |
| Fig. 3 Side (a) and upper (b) views of Au18@1, denoting the location of methyl ether groups from methylated BINOL (c) and Au⋯H3C– and Au⋯H–thiophene contacts. | |
Furthermore, the contribution of London interactions as complementary weak intermolecular interactions for the overall host–guest aggregation were evaluated in the interacting structure via the independent gradient model (IGM).100–102 This approach enabled isolation of the intermolecular interaction between the host and guest via the proposed electron-density based δginter descriptor by Hénon102 and coworkers. The stabilizing and repulsive characteristics of the intermolecular interaction can be evaluated by the second eigenvalue of the electron density Hessian (λ2). This accounts for an accumulation (attractive) or depletion (repulsive) of electron density, with the sign (λ2)ρ term as a useful descriptor for stabilizing (λ2 < 0), van der Waals (λ2 ≈ 0) or repulsive-type interactions (λ2 > 0).51,52,103 IGM analysis was carried out as implemented in the Multiwfn suite.104
The resulting intermolecular interactions from the IGM analysis (Fig. 4) were mainly ascribed at S–Au contact sites, regions below the methyl ether units from BINOL motifs given as Au⋯OMe and Au⋯H3C–, and Au⋯H–thiophene contacts. These results suggested that non-covalent contributions to the Au18 encapsulation interactions were due to the multiple interacting sites from the organic cavity.
 |
| Fig. 4 The independent gradient model (IGM) isosurface (0.005 a.u.), colored in the −0.10 a.u. < sign (λ2)ρ < +0.10 a.u. range for the Au18–1 interaction in Au18@1. Blue/red accounts for stabilizing/destabilizing non-covalent interactions, whereas green isosurfaces denote London-type interactions. | |
Moreover, a simplified model involving the four closest thiophene groups retaining their structure in the overall host–guest aggregate (Au18@(SC4H4)4) was used to evaluate the contribution of units from Au–S and Au⋯H–thiophene contacts (Fig. S4, SI). This model revealed an interaction energy of −53.9 kcal mol−1, which accounted for 37.1% from the overall Au18–1 interaction energy in the Au18@1 aggregate. Similarly, for the simplified model accounting for Au⋯OMe and Au⋯H3C– interactions (Au18@(OMe2)12), the interaction energy towards the Au18 cluster amounted to −59.7 kcal mol−1, representing 41.1% from the overall interaction energy related to the encapsulation of Au18 into 1 superphane. The remaining contribution to the overall interaction energy related to formation of the Au18@1 aggregate was given by weaker contributions from several Au18⋯HC
N– contacts at one side of the cavity. Hence, efficient encapsulation of the Au18 cluster was provided by the contribution of several weak interactions driven by the contacts to the superphane cavity.
Furthermore, the resulting incorporation of Au18 can involve complex and diverse conformational landscapes.105,106 Conformational exploration was evaluated through molecular dynamics via eXtended tight binding (xTB) methods at the GFN0-xTB level within the xtb code.73 The resulting trajectory involved structures with a more disperse Au18 cluster within the 1 cage (red arrow in Fig. 5), a partially aggregated Au18 structure (blue arrow), and a completely aggregated Au18 as the most favorable structure (green arrow). Along with the trajectory steps, different Au–cage interactions evolve whereby Au18 tends to aggregate as characterized previously.19 Recently, a benchmark report on the capabilities of xTB Hamiltonians (GFN0-xTB, GFN1-xTB, and GFN2-xTB) in achieving host–guest binding features with implicit solvent models107 denoted their performance in relation to MM-based techniques, appearing to be worthy alternatives if MM-based techniques are not applicable. In our case, due to the hybrid metallic–organic nature of the Au18–superphane aggregate, only GFN0-xTB could be applied successfully, and the resulting trajectory should be regarded as a qualitative conformational sampling, whereby quantitative interaction energy analyses were reliant upon DFT calculations.
 |
| Fig. 5 Molecular dynamic trajectories for the Au18@1 structure featuring a disperse Au18 cluster isomer (red arrow), a partially aggregated Au18 structure (blue arrow), and a complete aggregated Au18 structure (green arrow). The x-axis corresponds to 200 steps, covering a total simulation time of 10 ps. | |
The initial and middle stages of the calculated molecular dynamics trajectory in Fig. 5 are shown by red and blue arrows, respectively. The calculated Au18–superphane interaction energy at these stages (Table S3) showed a repulsive interaction of +82.6 kcal mol−1 at the initial step, which stabilized to −35.2 kcal mol−1 at the middle position of the trajectory, and settled at the interaction value of −145.2 kcal mol−1 for the Au18@1 structure discussed above. The main destabilizing factor was provided by an increase in the Pauli repulsion contribution at initial and middle steps. Hence, the more compact Au18 structure (final), as discussed above, provided more favorable encapsulation reducing the steric repulsion towards the overall cavity, in contrast to when more gold atoms are spread around the cavity. Thus, the compact metal cluster distribution favored a reduction in the steric repulsion upon encapsulation, thereby leading to more stable cluster incorporation.
We wished to evaluate the intermolecular charge transfer characteristics in the formation of Au18@1. Hence, the electronic coupling between Au18 and superphane host was calculated to quantify charge transport capabilities. Calculations revealed strong electronic coupling of 52.1 meV for electron transport (Ve) and 45.8 meV for hole transport (Vh) integrals. These values are in the range of one of the most efficient p-type semiconductors: benzothienobenzothiophene (BTBT) (33 meV for hole transport).108 The charge hopping rate (kCT) in Au18@1 can be modeled using the high-temperature limit of the Marcus theory.68–72 The latter is governed by two key parameters, the reorganization energy (λ) and intermolecular effective electronic coupling for electron transport (V),68–72 where T is the temperature and kb is the Boltzmann constant, as given by eqn (2):
|
 | (2) |
The adiabatic potential energy surfaces method was used to calculate the reorganization energy (λ).109 This involved the neutral and anionic optimized geometries of the isolated guest and host at neutral and anionic states, given by GE0 and GE−, and HE0 and HE−, respectively, taking into account solvation effects via the COSMO approach amounting to 0.17 eV, which is given by eqn (3):
|
λ = λ0 + λ− = (HEanion0 − HEneutral0 + GEanion0 − GEneutral0) + (HEneutral− − HEanion− + GEneutral− − GEanion−)
| (3) |
As a result, the estimated intermolecular charge hopping rate (kCT) was 2.2 × 1013 s−1 at 300 K, in the range of ultrafast electron transfer in supramolecular aggregates (4.0 × 1011 s−1),110 and from other intermolecular interactions (5.0 × 1012 s−1).111 Thus, the Au18 → 1 charge transfer was driven by coordinating and short contacts towards the superphane available cavity, resulting in a supramolecular structure of donor(Au18)–acceptor(superphane) D–A system.
Conclusions
The favorable incorporation of the Au18 cluster into the polyfunctional superphane cavity 1 amounted to −145.3 kcal mol−1, mainly provided by electrostatic-type interactions (54.9%), leading to a stable Au18@1 aggregate. The electrostatic contribution was given by charge reorganization over Au18, which enhanced the interaction towards the 1 cavity, and accounted for the observed stability of the overall host–guest pair. Charge transfer involved host ← guest and host → guest backdonation through S ← Au and S → Au contacts, leading to an overall Au18 → 1 superphane charge transfer of +0.79e. Hypothetical models involving isolated cavity units suggested that the Au18–thiophane interaction amounted to −53.9 kcal mol−1, accounting for 37.1% of the overall stabilization, whereas the isolated methyl ether units from BINOL motifs amounted to −59.7 kcal mol−1, accounting for 41.1% of the Au18@1 aggregate stabilization. Moreover, non-covalent interactions contributed to a lesser extent to Au18@1 formation.
The resulting aggregate showed favorable charge transfer. A charge hopping rate (kCT) in the range of ultrafast electron transfer, calculated to be 2.2 × 1013 s−1 at 300 K, was documented. Thus, the Au18 → 1 charge transfer was driven by coordinating and short contacts towards the superphane available cavity, resulting in the supramolecular structure of the D–A system.
We expect that the current approach could help to characterize further the stabilizing factors leading to the aggregation of metallic clusters into organic host cavities. This strategy may aid rational design and explorative synthetic efforts, providing novel functionality for host–guest aggregates.
Conflicts of interest
There are no conflicts of interest to declare.
Data availability
The data supporting this article have been included as part of the supplementary information (SI). Supplementary information is available. See DOI: https://doi.org/10.1039/d5cp01989c.
Acknowledgements
M. P.-C. acknowledges a PhD scholarship from Vicerrectoría de Investigación y Doctorados, Universidad San Sebastián. This work was supported by ANID FONDECYT Regular 1221676. P. L. R.-K. would like to thank the CIMAT Supercomputing Laboratories of Guanajuato and Puerto Interior for their support.
References
- G. T. Williams, C. J. E. Haynes, M. Fares, C. Caltagirone, J. R. Hiscock and P. A. Gale, Advances in applied supramolecular technologies, Chem. Soc. Rev., 2021, 50, 2737–2763 RSC.
- F. Huang and E. V. Anslyn, Introduction: Supramolecular Chemistry, Chem. Rev., 2015, 115, 6999–7000 CrossRef CAS PubMed.
- P. Commins, M. B. Al-Handawi and P. Naumov, Self-healing crystals, Nat. Rev. Chem., 2025, 9, 343–355 CrossRef PubMed.
- W. He, Y. Yu, K. Iizuka, H. Takezawa and M. Fujita, Supramolecular coordination cages as crystalline sponges through a symmetry mismatch strategy, Nat. Chem., 2025, 17, 653–662 CrossRef CAS PubMed.
- H.-N. Zhang and G.-X. Jin, Synthesis of molecular Borromean links featuring trimeric metallocages, Nat. Synth., 2025, 4, 488–496 CrossRef CAS.
- C. Li, A. Iscen, H. Sai, K. Sato, N. A. Sather, S. M. Chin, Z. Álvarez, L. C. Palmer, G. C. Schatz and S. I. Stupp, Supramolecular–covalent hybrid polymers for light-activated mechanical actuation, Nat. Mater., 2020, 19, 900–909 CrossRef CAS PubMed.
- D. Zhang, A. Martinez and J.-P. Dutasta, Emergence of Hemicryptophanes: From Synthesis to Applications for Recognition, Molecular Machines, and Supramolecular Catalysis, Chem. Rev., 2017, 117, 4900–4942 CrossRef CAS PubMed.
- D. Prochowicz, A. Kornowicz and J. Lewiński, Interactions of Native Cyclodextrins with Metal Ions and Inorganic Nanoparticles: Fertile Landscape for Chemistry and Materials Science, Chem. Rev., 2017, 117, 13461–13501 CrossRef CAS PubMed.
- G. Yu, K. Jie and F. Huang, Supramolecular Amphiphiles Based on Host–Guest Molecular Recognition Motifs, Chem. Rev., 2015, 115, 7240–7303 CrossRef CAS PubMed.
- D.-H. Qu, Q.-C. Wang, Q.-W. Zhang, X. Ma and H. Tian, Photoresponsive Host–Guest Functional Systems, Chem. Rev., 2015, 115, 7543–7588 CrossRef CAS PubMed.
- H. Amouri, C. Desmarets and J. Moussa, Confined Nanospaces in Metallocages: Guest Molecules, Weakly Encapsulated Anions, and Catalyst Sequestration, Chem. Rev., 2012, 112, 2015–2041 CrossRef CAS PubMed.
- R. D. Mukhopadhyay, G. Das and A. Ajayaghosh, Stepwise control of host–guest interaction using a coordination polymer gel, Nat. Commun., 2018, 9, 1987 CrossRef PubMed.
- M. Zhang, Z. Gong, W. Yang, L. Jin, S. Liu, S. Chang and F. Liang, Regulating Host–Guest Interactions between Cucurbit[7]uril and Guests on Gold Surfaces for Rational Engineering of Gold Nanoparticles, ACS Appl. Nano Mater., 2020, 3, 4283–4291 CrossRef CAS.
- G. Kali, S. Haddadzadegan and A. Bernkop-Schnürch, Cyclodextrins and derivatives in drug delivery: New developments, relevant clinical trials, and advanced products, Carbohydr. Polym., 2024, 324, 121500 CrossRef CAS PubMed.
- H. Asghar, S. Bilal, M. H. Nawaz, G. Rasool and A. Hayat, Host–Guest Mechanism via Induced Fit Fullerene Complexation in Porphin Receptor to Probe Salivary Alpha-Amylase in Dental Caries for Clinical Applications, ACS Appl. Bio Mater., 2024, 7, 1250–1259 CrossRef CAS PubMed.
- M. Li, Y.-Q. Shi, X. Gan, L. Su, J. Liang, H. Wu, Y. You, M. Che, P. Su, T. Wu, Z. Zhang, W. Zhang, L.-Y. Yao, P. Wang and T.-Z. Xie, Coordination-Driven Tetragonal Prismatic Cage and the Investigation on Host–Guest Complexation, Inorg. Chem., 2023, 62, 4393–4398 CrossRef CAS PubMed.
- A. A. Ivanov, P. A. Abramov, M. Haouas, Y. Molard, S. Cordier, C. Falaise, E. Cadot and M. A. Shestopalov, Supramolecular Host–Guest Assemblies of [M6Cl14]2–,
M = Mo, W, Clusters with γ-Cyclodextrin for the Development of CLUSPOMs, Inorganics, 2023, 11, 77 CrossRef CAS.
- N. Song, J. Zhang, J. Zhai, J. Hong, C. Yuan and M. Liang, Ferritin: A Multifunctional Nanoplatform for Biological Detection, Imaging Diagnosis, and Drug Delivery, Acc. Chem. Res., 2021, 54, 3313–3325 CrossRef CAS PubMed.
- Y. Zhang, J. Zhou, K. Luo, W. Zhou, F. Wang, J. Li and Q. He, Ferritin-Inspired Encapsulation and Stabilization of Gold Nanoclusters for High-Performance Photothermal Conversion, Angew. Chem., Int. Ed., 2025, 64, e202500058 CrossRef CAS PubMed.
- Y. Yu, Y. Hu, C. Ning, W. Shi, A. Yang, Y. Zhao, Z. Cao, Y. Xu and P. Du, BINOL-Based Chiral Macrocycles and Cages, Angew. Chem., Int. Ed., 2024, 63, e202407034 CrossRef CAS PubMed.
- S. Maity, S. Kolay, S. Chakraborty, A. Devi, Rashi and A. Patra, A comprehensive review of atomically precise metal nanoclusters with emergent photophysical properties towards diverse applications, Chem. Soc. Rev., 2025, 54, 1785–1844 RSC.
- J. Qian, Z. Yang, J. Lyu, Q. Yao and J. Xie, Molecular Interactions in Atomically Precise Metal Nanoclusters, Precis. Chem., 2024, 2, 495–517 CrossRef CAS PubMed.
- R. J. Wilson, N. Lichtenberger, B. Weinert and S. Dehnen, Intermetalloid and Heterometallic Clusters Combining p-Block (Semi)Metals with d- or f-Block Metals, Chem. Rev., 2019, 119, 8506–8554 CrossRef CAS PubMed.
- B. Peters, N. Lichtenberger, E. Dornsiepen and S. Dehnen, Current advances in tin cluster chemistry, Chem. Sci., 2020, 11, 16–26 RSC.
- F. Li, A. Muñoz-Castro and S. C. S. C. Sevov, [Ge9{Si(SiMe3)3}3{SnPh3}]: A Tetrasubstituted and Neutral Deltahedral Nine-Atom Cluster, Angew. Chem., Int. Ed., 2012, 51, 8581–8584 CrossRef CAS PubMed.
- A. Jana, A. R. Kini and T. Pradeep, Atomically Precise Clusters: Chemical Evolution of Molecular Matter at the Nanoscale, AsiaChem Mag., 2023, 3, 56–65 Search PubMed.
- L. He, T. Dong, D. Jiang and Q. Zhang, Recent progress in atomically precise metal nanoclusters: From bio-related properties to biological applications, Coord. Chem. Rev., 2025, 535, 216633 CrossRef CAS.
- J. Zhang, Z. Li, K. Zheng and G. Li, Synthesis and characterization of size-controlled atomically precise gold clusters, Phys. Sci. Rev., 2018, 3, 20170083 Search PubMed.
- S. Biswas, S. Das and Y. Negishi, Progress and prospects in the design of functional atomically-precise Ag(I)-thiolate nanoclusters and their assembly approaches, Coord. Chem. Rev., 2023, 492, 215255 CrossRef CAS.
- R. Nakatani, S. Das and Y. Negishi, The structure and application portfolio of intricately architected silver cluster-assembled materials, Nanoscale, 2024, 16, 9642–9658 RSC.
- T. Kawawaki, T. Okada, D. Hirayama and Y. Negishi, Atomically precise metal nanoclusters as catalysts for electrocatalytic CO2 reduction, Green Chem., 2024, 26, 122–163 RSC.
- E. Khatun and T. Pradeep, New Routes for Multicomponent Atomically Precise Metal Nanoclusters, ACS Omega, 2021, 6, 1–16 CrossRef CAS PubMed.
- Y. Du, H. Sheng, D. Astruc and M. Zhu, Atomically Precise Noble Metal Nanoclusters as Efficient Catalysts: A Bridge between Structure and Properties, Chem. Rev., 2020, 120, 526–622 CrossRef CAS PubMed.
- M. A. Moussawi, N. Leclerc-Laronze, S. Floquet, P. A. Abramov, M. N. Sokolov, S. Cordier, A. Ponchel, E. Monflier, H. Bricout, D. Landy, M. Haouas, J. Marrot and E. Cadot, Polyoxometalate, Cationic Cluster, and γ-Cyclodextrin:
From Primary Interactions to Supramolecular Hybrid Materials, J. Am. Chem. Soc., 2017, 139, 12793–12803 CrossRef CAS PubMed.
- A. A. Ivanov, C. Falaise, K. Laouer, F. Hache, P. Changenet, Y. V. Mironov, D. Landy, Y. Molard, S. Cordier, M. A. Shestopalov, M. Haouas and E. Cadot, Size-Exclusion Mechanism Driving Host–Guest Interactions between Octahedral Rhenium Clusters and Cyclodextrins, Inorg. Chem., 2019, 58, 13184–13194 CrossRef CAS PubMed.
- A. A. Ivanov, C. Falaise, A. A. Shmakova, N. Leclerc, S. Cordier, Y. Molard, Y. V. Mironov, M. A. Shestopalov, P. A. Abramov, M. N. Sokolov, M. Haouas and E. Cadot, Cyclodextrin-Assisted Hierarchical Aggregation of Dawson-type Polyoxometalate in the Presence of {Re6Se8} Based Clusters, Inorg. Chem., 2020, 59, 11396–11406 CrossRef CAS PubMed.
- A. A. Ivanov, C. Falaise, P. A. Abramov, M. A. Shestopalov, K. Kirakci, K. Lang, M. A. Moussawi, M. N. Sokolov, N. G. Naumov, S. Floquet, D. Landy, M. Haouas, K. A. Brylev, Y. V. Mironov, Y. Molard, S. Cordier and E. Cadot, Host–Guest Binding Hierarchy within Redox- and Luminescence-Responsive Supramolecular Self-Assembly Based on Chalcogenide Clusters and γ-Cyclodextrin, Chem. – Eur. J., 2018, 24, 13467–13478 CrossRef CAS PubMed.
- H. Zhu, J. Li, J. Wang and E. Wang, Lighting Up the Gold Nanoclusters via Host–Guest Recognition for High-Efficiency Antibacterial Performance and Imaging, ACS Appl. Mater. Interfaces, 2019, 11, 36831–36838 CrossRef CAS PubMed.
- S. Hu, H. Zhao, P. Xie, X. Zhu, L. Liu, N. Yin, Z. Tang, K. Peng and R. Yuan, High-efficiency electrochemiluminescence of host-guest ligand-assembled gold nanoclusters preoxidated for ultrasensitive biosensing of protein, Sens. Actuators, B, 2024, 419, 136347 CrossRef CAS.
- K. I. Assaf, A. Hennig, S. Peng, D.-S. Guo, D. Gabel and W. M. Nau, Hierarchical host–guest assemblies formed on dodecaborate-coated gold nanoparticles, Chem. Commun., 2017, 53, 4616–4619 RSC.
- K. Sahoo, T. R. Gazi, S. Roy and I. Chakraborty, Nanohybrids of atomically precise metal nanoclusters, Commun. Chem., 2023, 6, 157 CrossRef CAS PubMed.
- P. Pyykkö, Understanding the eighteen-electron rule, J. Organomet. Chem., 2006, 691, 4336–4340 CrossRef.
- H. Häkkinen, in Protected Metal Clusters: From Fundamentals to Applications, ed. H. Hakkinen and T. Tsukuda, Elsevier Science, 2015, pp. 189–222 Search PubMed.
- H. Häkkinen, Atomic and electronic structure of gold clusters: understanding flakes, cages and superatoms from simple concepts, Chem. Soc. Rev., 2008, 37, 1847–1859 RSC.
- J. U. Reveles, S. N. Khanna, P. J. Roach and A. W. Castleman, Multiple valence superatoms, Proc. Natl. Acad. Sci. U. S. A., 2006, 103, 18405–18410 CrossRef CAS PubMed.
- M. von Hopffgarten and G. Frenking, Energy decomposition analysis, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 43–62 CAS.
- M. von Hopffgarten and G. Frenking, Building a bridge between coordination compounds and clusters: bonding analysis of the icosahedral molecules [M(ER)12] (M = Cr, Mo, W; E = Zn, Cd, Hg), J. Phys. Chem. A, 2011, 115, 12758–12768 CrossRef CAS PubMed.
- T. Brinck, in Theoretical and Computational Chemistry, ed. C. Párkányi, 1998, pp. 51–93 Search PubMed.
- J. Halldin Stenlid, A. J. Johansson and T. Brinck, σ-Holes and σ-lumps direct the Lewis basic and acidic interactions of noble metal nanoparticles: introducing regium bonds, Phys. Chem. Chem. Phys., 2018, 20, 2676–2692 RSC.
- J. H. Stenlid and T. Brinck, Extending the σ-Hole Concept to Metals: An Electrostatic Interpretation of the Effects of Nanostructure in Gold and Platinum Catalysis, J. Am. Chem. Soc., 2017, 139, 11012–11015 CrossRef CAS PubMed.
- G. Saleh, C. Gatti and L. Lo Presti, Non-covalent interaction via the reduced density gradient: Independent atom model vs experimental multipolar electron densities, Comput. Theor. Chem., 2012, 998, 148–163 CrossRef CAS.
- G. Saleh, C. Gatti, L. Lo Presti and J. Contreras-García, Revealing Non-covalent Interactions in Molecular Crystals through Their Experimental Electron Densities, Chem. – Eur. J., 2012, 18, 15523–15536 CrossRef CAS PubMed.
- Amsterdam Density Functional (ADF 2024) Code, Vrije Universiteit: Amsterdam, The Netherlands. Available at: https://www.scm.com.
- S. Grimme, Accurate description of van der Waals complexes by density functional theory including empirical corrections, J. Comput. Chem., 2004, 25, 1463–1473 CrossRef CAS PubMed.
- S. Grimme, Semiempirical GGA-type density functional constructed with a long-range dispersion correction, J. Comput. Chem., 2006, 27, 1787–1799 CrossRef CAS PubMed.
- S. Grimme, J. Antony, S. Ehrlich and H. Krieg, A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
- S. Grimme, Density functional theory with London dispersion corrections, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2011, 1, 211–228 CAS.
- E. R. Johnson and A. D. Becke, A post-Hartree–Fock model of intermolecular interactions, J. Chem. Phys., 2005, 123, 024101 CrossRef PubMed.
- S. Grimme, S. Ehrlich and L. Goerigk, Effect of the damping function in dispersion corrected density functional theory, J. Comput. Chem., 2011, 32, 1456–1465 CrossRef CAS PubMed.
- A. Suvitha and N. S. Venkataramanan, Trapping of organophosphorus chemical nerve agents by pillar[5]arene: A DFT, AIM, NCI and EDA analysis, J. Inclusion Phenom. Macrocyclic Chem., 2017, 87, 207–218 CrossRef CAS.
- J. C. Babón, M. A. Esteruelas, I. Fernández, A. M. López and E. Oñate, Evidence for a Bis(Elongated σ)-Dihydrideborate Coordinated to Osmium, Inorg. Chem., 2018, 57, 4482–4491 CrossRef PubMed.
- L. Versluis and T. Ziegler, The determination of molecular structures by density functional theory. The evaluation of analytical energy gradients by numerical integration, J. Chem. Phys., 1988, 88, 322–328 CrossRef CAS.
- S. Simon, M. Duran and J. J. Dannenberg, How does basis set superposition error change the potential surfaces for hydrogen-bonded dimers?, J. Chem. Phys., 1996, 105, 11024–11031 CrossRef CAS.
- F. Jensen, Basis Set Superposition Errors Are Partly Basis Set Imbalances, J. Chem. Theory Comput., 2024, 20, 767–774 CrossRef CAS PubMed.
- S. F. Boys and F. Bernardi, The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors, Mol. Phys., 1970, 19, 553–566 CrossRef CAS.
- A. Klamt and V. Jonas, Treatment of the outlying charge in continuum solvation models, J. Chem. Phys., 1996, 105, 9972 CrossRef CAS.
- G. Te Velde, F. M. Bickelhaupt, E. J. Baerends, C. Fonseca Guerra, S. J. van Gisbergen, J. G. Snijders and T. Ziegler, Chemistry with ADF, J. Comput. Chem., 2001, 22, 931–967 CrossRef CAS.
- K. Senthilkumar, F. C. Grozema, F. M. Bickelhaupt and L. D. A. Siebbeles, Charge transport in columnar stacked triphenylenes: Effects of conformational fluctuations on charge transfer integrals and site energies, J. Chem. Phys., 2003, 119, 9809–9817 CrossRef CAS.
- K. Senthilkumar, F. C. Grozema, C. F. Guerra, F. M. Bickelhaupt, F. D. Lewis, Y. A. Berlin, M. A. Ratner and L. D. A. Siebbeles, Absolute Rates of Hole Transfer in DNA, J. Am. Chem. Soc., 2005, 127, 14894–14903 CrossRef CAS PubMed.
- R. A. Marcus, Electron transfer reactions in chemistry. Theory and experiment, Rev. Mod. Phys., 1993, 65, 599–610 CrossRef CAS.
- R. A. Marcus, On the Theory of Oxidation-Reduction Reactions Involving Electron Transfer. III. Applications to Data on the Rates of Organic Redox Reactions, J. Chem. Phys., 1957, 26, 872–877 CrossRef CAS.
- N. S. Hush, Adiabatic Rate Processes at Electrodes. I. Energy-Charge Relationships, J. Chem. Phys., 1958, 28, 962–972 CrossRef CAS.
- C. Bannwarth, E. Caldeweyher, S. Ehlert, A. Hansen, P. Pracht, J. Seibert, S. Spicher and S. Grimme, Extended tight-binding quantum chemistry methods, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2021, 11, e1493 CAS.
- J. B. Maglic and R. Lavendomme, MoloVol: an easy-to-use program for analyzing cavities, volumes and surface areas of chemical structures, J. Appl. Crystallogr., 2022, 55, 1033–1044 CrossRef CAS PubMed.
- H. Häkkinen, M. Walter and H. Grönbeck, Divide and Protect: Capping Gold Nanoclusters with Molecular Gold-Thiolate Rings, J. Phys. Chem. B, 2006, 110, 9927–9931 CrossRef PubMed.
- D. E. Bergeron, A. W. Castleman, T. Morisato and S. N. Khanna, Formation of Al13I-: Evidence for the superhalogen character of Al13, Science, 2004, 304, 84–87 CrossRef CAS PubMed.
- A. C. Reber, S. N. Khanna and A. W. Castleman, Superatom Compounds, Clusters, and Assemblies: Ultra Alkali Motifs and Architectures, J. Am. Chem. Soc., 2007, 129, 10189–10194 CrossRef CAS PubMed.
- A. C. Reber and S. N. Khanna, Superatoms: Electronic and Geometric Effects on Reactivity, Acc. Chem. Res., 2017, 50, 255–263 CrossRef CAS PubMed.
- T. Tsukuda and H. Häkkinen, Protected Metal Clusters: From Fundamentals to Applications, Elsevier, 2015 Search PubMed.
- P. V. Nhat, N. T. Si, A. Fielicke, V. G. Kiselev and M. T. Nguyen, A new look at the structure of the neutral Au18 cluster: hollow versus filled golden cage, Phys. Chem. Chem. Phys., 2023, 25, 9036–9042 RSC.
- T. Ziegler and A. Rauk, A theoretical study of the ethylene-metal bond in complexes between copper(1+), silver(1+), gold(1+), platinum(0) or platinum(2+) and ethylene, based on the Hartree-Fock-Slater transition-state method, Inorg. Chem., 1979, 18, 1558–1565 CrossRef CAS.
- M. von Hopffgarten, G. Frenking, M. von Hopffgarten and G. Frenking, Energy decomposition analysis, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2012, 2, 43–62 CAS.
- L. Zhao, M. von Hopffgarten, D. M. Andrada and G. Frenking, Energy decomposition analysis, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2018, 8, e1345 Search PubMed.
- F. Kreuter and R. Tonner-Zech, Energy decomposition analysis for excited states: an extension based on TDDFT, Phys. Chem. Chem. Phys., 2025, 27, 4728–4745 RSC.
- K. Morokuma, Molecular Orbital Studies of Hydrogen Bonds. III. C
O⋯H–O Hydrogen Bond in H2CO⋯H2O and H2CO⋯2H2O, J. Chem. Phys., 1971, 55, 1236–1244 CrossRef CAS. - T. Ziegler and A. Rauk, Calculation of Bonding Energies by Hartree-Fock Slater Method. 1. Transition-State Method, Theor. Chim. Acta, 1977, 46, 1–10 CrossRef CAS.
- B. Pinter, T. Fievez, F. M. Bickelhaupt, P. Geerlings and F. De Proft, On the origin of the steric effect, Phys. Chem. Chem. Phys., 2012, 14, 9846 RSC.
- P. Su and H. Li, Energy decomposition analysis of covalent bonds and intermolecular interactions, J. Chem. Phys., 2009, 131, 014102 CrossRef PubMed.
- F. A. Bulat, A. Toro-Labbé, T. Brinck, J. S. Murray and P. Politzer, Quantitative analysis of molecular surfaces: areas, volumes, electrostatic potentials and average local ionization energies, J. Mol. Model., 2010, 16, 1679–1691 CrossRef CAS PubMed.
- J. S. Murray and P. Politzer, The electrostatic potential: an overview, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2011, 1, 153–163 CAS.
- T. Clark, σ-Holes, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2013, 3, 13–20 CAS.
- K. J. Donald, N. Pham and P. Ravichandran, Sigma Hole Potentials as Tools: Quantifying and Partitioning Substituent Effects, J. Phys. Chem. A, 2023, 127, 10147–10158 CrossRef CAS PubMed.
- J. Halldin Stenlid and F. Abild-Pedersen, Revealing Local and Directional
Aspects of Catalytic Active Sites by the Nuclear and Surface Electrostatic Potential, J. Phys. Chem. C, 2024, 128, 4544–4558 CrossRef CAS.
- A. Michalak, M. Mitoraj and T. Ziegler, Bond orbitals from chemical valence theory, J. Phys. Chem. A, 2008, 112, 1933–1939 CrossRef CAS PubMed.
- M. P. Mitoraj, A. Michalak and T. Ziegler, A combined charge and energy decomposition scheme for bond analysis, J. Chem. Theory Comput., 2009, 5, 962–975 CrossRef CAS PubMed.
- R. F. Nalewajski, J. Mrozek and A. Michalak, Two-electron valence indices from the Kohn-Sham orbitals, Int. J. Quantum Chem., 1997, 61, 589–601 CrossRef CAS.
- A. Michalak, R. L. DeKock and T. Ziegler, Bond multiplicity in transition-metal complexes: applications of two-electron valence indices, J. Phys. Chem. A, 2008, 112, 7256–7263 CrossRef CAS PubMed.
- The Chemical Bond: Fundamental Aspects of Chemical Bonding, ed. G. Frenking and S. Shaik, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, 2014 Search PubMed.
- X. Wang, Z. Wu, Z. Gong, D. Muhammad, B. Hu, P. Su and Z. Sun, All-atom Picture of Cucurbituril Detoxification Coordination, ChemRxiv, 2025, preprint DOI:10.26434/chemrxiv-2025-1x058-v2.
- T. Lu and Q. Chen, Independent gradient model based on Hirshfeld partition: A new method for visual study of interactions in chemical systems, J. Comput. Chem., 2022, 43, 539–555 CrossRef CAS PubMed.
- T. Lu and Q. Chen, Comprehensive Computational Chemistry, Elsevier, 2024, pp. 240–264 Search PubMed.
- C. Lefebvre, G. Rubez, H. Khartabil, J.-C. Boisson, J. Contreras-García and E. Hénon, Accurately extracting the signature of intermolecular interactions present in the NCI plot of the reduced density gradient versus electron density, Phys. Chem. Chem. Phys., 2017, 19, 17928–17936 RSC.
- J. Contreras-García, E. R. Johnson, S. Keinan, R. Chaudret, J.-P. P. Piquemal, D. N. Beratan and W. Yang, NCIPLOT: A Program for Plotting Noncovalent Interaction Regions, J. Chem. Theory Comput., 2011, 7, 625–632 CrossRef PubMed.
- T. Lu and F. Chen, Multiwfn: a multifunctional wavefunction analyzer, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed.
- P. Procacci, Dealing with Induced Fit, Conformational Selection, and Secondary Poses in Molecular Dynamics Simulations for Reliable Free Energy Predictions, J. Chem. Theory Comput., 2023, 19, 8942–8954 CrossRef CAS PubMed.
- Z. Sun, Z. Huai, Q. He and Z. Liu, A General Picture of Cucurbit[8]uril Host–Guest Binding, J. Chem. Inf. Model., 2021, 61, 6107–6134 CrossRef CAS PubMed.
- X. Wang, S. Li, Z. Zhang, L. Qiu and Z. Sun, Multiscale End-point Screening with Extended Tight-binding Hamiltonians, BIO Integr., 2025, 6, e982 Search PubMed.
- L. Ueberricke, J. Schwarz, F. Ghalami, M. Matthiesen, F. Rominger, S. M. Elbert, J. Zaumseil, M. Elstner and M. Mastalerz, Triptycene End-Capped Benzothienobenzothiophene and Naphthothienobenzothiophene, Chem. – Eur. J., 2020, 26, 12596–12605 CrossRef CAS PubMed.
- G. R. Hutchison, M. A. Ratner and T. J. Marks, Hopping Transport in Conductive Heterocyclic Oligomers: Reorganization Energies and Substituent Effects, J. Am. Chem. Soc., 2005, 127, 2339–2350 CrossRef CAS PubMed.
- A. Das, N. Kamatham, A. Mohan Raj, P. Sen and V. Ramamurthy, Marcus Relationship Maintained During Ultrafast Electron Transfer Across a Supramolecular Capsular Wall, J. Phys. Chem. A, 2020, 124, 5297–5305 CrossRef CAS PubMed.
- S. Mallick, L. Cao, X. Chen, J. Zhou, Y. Qin, G. Y. Wang, Y. Y. Wu, M. Meng, G. Y. Zhu, Y. N. Tan, T. Cheng and C. Y. Liu, Mediation of Electron Transfer by Quadrupolar Interactions: The Constitutional, Electronic, and Energetic
Complementarities in Supramolecular Chemistry, iScience, 2019, 22, 269–287 CrossRef CAS PubMed.
|
This journal is © the Owner Societies 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.