Formation of Coinage-Metal···Fullerene Adducts. Evaluation of the

Oct 11, 2018 - Evaluation of the Interaction Nature between Triangular Coinage Metal ... The recent formation of [M3(3,5-(CF3)2Pz)3]-C60 cocrystals in...
0 downloads 0 Views 711KB Size
Subscriber access provided by UNIV OF CAMBRIDGE

C: Physical Processes in Nanomaterials and Nanostructures

Formation of Coinage-Metal···Fullerene Adducts. Evaluation of the Interaction Nature between Triangular Coinage Metal Complexes (M=Cu, Ag and Au) and C through Relativistic DFT Calculations 3

60

Carolina Olea Ulloa, Miguel Ponce-Vargas, and Alvaro Muñoz-Castro J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b08417 • Publication Date (Web): 11 Oct 2018 Downloaded from http://pubs.acs.org on October 12, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Formation of Coinage-Metal∙∙∙Fullerene Adducts. Evaluation of the Interaction Nature between Triangular Coinage Metal Complexes (M3=Cu, Ag and Au) and C60 through Relativistic DFT Calculations

Carolina Olea Ulloa,a Miguel Ponce-Vargas,b Alvaro Muñoz-Castroa,* aLaboratorio

de Química Inorgánica y Materiales Moleculares, Universidad Autonoma de Chile,

Llano Subercaceaux 2801, San Miguel, Santiago, Chile. bInstitut

de Chimie Moléculaire de Reims,UMR CNRS 7312, University of Reims Champagne-

Ardenne, Moulin de la Housse, Reims 51687, France Abstract The recent formation of [M3(3,5-(CF3)2Pz)3]-C60 cocrystals in a 4:1 ratio, have shown the coinage metal complexes ability to bind fullerene, acting as Buckycatchers. Here, we clarify the nature of such interaction accounting for the stabilization of the whole assembly via two models of different 4:1 and 1:1 ratios, within the framework of relativistic dispersion corrected DFT. Our results exhibit a strong van der Waals character in the interaction, supported by the electrostatic character of the acidic-M3 ring provided by coinage metals. This feature held constant throughout the coinage metal group, providing further guidance for explorative synthesis efforts seeking to increase the strength and versatility of the binding capabilities of coinage metal complexes, which is useful for predicting the rise of noncovalent interactions towards less symmetric fullerenes and endohedral metallofullerenes. In addition, the observed van der Waals character is retained in the hypothetical systems involving representative fullerene fragments, based on coranulenne (C20H10) and sumanene (C21H12). These results highlight the versatility of trinuclear complexes which can adopt a convex distortion highly suitable for interacting with curved π-surfaces. Moreover, the nature and strength of the interaction do not significantly vary with the number of [M3] complexes 1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

in the C60 fullerene case, suggesting that the number of coinage-metal units involved in the adduct formation is related to the π-surface area available in the fullerene structure, and the stoichiometry employed in the co-crystallization. Hence, we envisage the exploration of novel supramolecular arrays for the formation of structures featuring pre-organized domains involving fullerenes and other appealing π-systems. Keywords: coinage metal, Fullerenes, corannulene, non-covalent *[email protected]

2 ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Introduction The discovery of buckminsterfullerene (C60),1–5 opened a new era of carbon nanotechnology because of the unique electronic and physicochemical properties of such carbon clusters.6–10 Several approaches have been employed to achieve useful quantities of C60, ranging from rational synthesis11 to hydrocarbon combustion,12 besides the usual carbon soot method.13 This allows for the fabrication of further devices based on C60,14–16 highlighting self-assembly via non-covalent interactions, introducing the required flexibility to facilitate construction of molecular arrays towards controlled larger assemblies.17–20 To date, octaethyl-metalloporphyrins (MOEP) have been widely exploited in the successful formation of self-assembly arrays20–23 as efficient co-crystallization agents, enabling structural characterization of several fullerenes and endohedral fullerenes.24–26 Such MOEP-fullerene selforganized systems offer a plausible “bottom-up” approach for molecule-scale devices involving fullerene derivatives. In addition, different co-crystallization agents have been studied owing to their relevance in controlling the shape and characteristics of the self-assembly array. In this concern, the intermolecular fullerene-substrate interaction nature is relevant in determining different supramolecular nanostructures, which can be controlled via modification of fullerenes and/or variation of different chemical groups attached to substrates. Recently, Dias and coworkers report a new member to the widely employed organic porphyrin derivatives,27,28,29,30 introducing trinuclear coinage-metal complexes resulting in [M3(3,5(CF3)2Pz)3]-C60 4:1 cocrystals, with M = Cu, Ag, Au.31 A tetrahedral disposition of coinage-metal complexes was discovered around fullerene, which is persistent throughout the group, providing further tailoring of new templates in self-assembly. They can be considered promising building blocks given their inherent supramolecular assembly capabilities.32–37 [M3(3,5-(CF3)2Pz)3] (M= Cu, Ag, Au) species, can be viewed as interesting substrates, as they can establish acid-M3∙∙∙π-surface interactions, providing an alternative to the usual π−π interactions observed in ball-and-socket arrangements, in addition to more flexibility of the ball relative to the socket, based on the enhanced acidity of the central M3 ring.36,38,39 Hence, such species can provide promising building blocks for the formation of extended networks featuring pre-organized domains towards functional molecules such as fullerenes and other interesting π-systems.28 Thus, planar coinage metal rings appear as a novel alternative to the well employed aromatic hydrocarbons in the construction of π−π

3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

intermolecular, presenting interesting challenges facing the design of more versatile assemblies for fullerene recognition and isolation. Herein we report our findings using dispersion-corrected density functional calculations, taking into account relativistic effects, in order to better understand the intermolecular interaction between [M3(3,5-(CF3)2Pz)3] and C60 in two ratios given by 4:1 and 1:1, with M=Cu(I), Ag(I), Au(I). In addition, the interaction with corannulene (C20H10) and sumanene (C21H12), as prototypical fullerene fragments accounting for surfaces with both five- and six-membered central rings, respectively, is given. The nature of the stabilizing fullerene-substrate interaction was explored by means of energy decomposition analysis (EDA),40,41 electrostatic potential maps,42–44 and a noncovalent index analysis (NCI).45,46

4 ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Computational Details Geometry optimizations and subsequent calculations were performed using the ADF2016 code.47,48 Triple-ξ and two polarization functions (STO-TZ2P) basis sets were employed within the generalized gradient approximation (GGA) according to the BP86 exchange-correlation functional, showing improved performance on long-range interactions in large structures.49–51 The pair-wise Grimme correction (D3)49,52–54 and Becke-Johnson damping functions55,56 were taken into account for the empirical dispersion correction to DFT (DFT-D). The molecular structures were optimized through the analytical energy gradient method implemented by Versluis and Ziegler57 at the TZ2P/BP86-D level without any symmetry restraint. An energy convergence criterion of 10-5 Hartree, gradient convergence criteria of 10-4 Hartree/Å and radial convergence criteria of 10-3 Å were employed for the evaluation of the final relaxed structures, which allows to account for energy differences of 10-3 kcal∙mol-1 of accuracy. The analysis of non-covalent interactions (NCI)57,58 was carried out using the NCIPLOT program developed by Weitao Yang et al.58 based on the analysis of electronic density descriptors. This allows a better understanding of the spatial distribution of non-covalent interactions, on the basis of the reduced density gradient (s(ρ)) in low-density regions (ρ(r) < 0.06 a.u.), equation 1, where the second eigenvalue of the electronic density Hessian (λ2) is calculated to determine the type of interaction. Consequently, stabilizing forces such as hydrogen bonds are characterized by λ2 < 0, steric repulsion by λ2 > 0, and London interactions by λ2 ≈ 0. The product ρ(r)*sign(λ2) is employed as an indicator of the nature of the interactions as denoted by the respective color scale. (1) 𝑠=

1

|∇𝜌| 1

4

2(3𝜋 ) 𝜌3 2 3

The molecular electrostatic potential (V(r), equation 2) is a well-established and useful tool in the quantitative analysis of intermolecular stabilizing or destabilizing interactions,42,59–61 which can be calculated in order to account for charge distribution in the whole molecular structure, denoting σholes among other types of electrostatic interactions. (2)

5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Where V(r) is the potential created by the nuclear and electronic density of the charge density, in atomic units ZA is the charge on nucleus located at RA, given by the atomic number of atom A, and ρ(r) is the electronic density. The sign and strength of V(r) is given by the predominance of nuclei (positive) or electronic (negative) effects on different molecular regions. The surface was set to 0.001 a.u. (electrons/Bohr3), as a contour of the electronic density accounting for van der Waals surface of the molecule.59,62

6 ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Results and Discussion The representative structures for 4:1 and 1:1 [M3(3,5-(CF3)2Pz)3]-C60 species are displayed in Figure 1 ([M3]-C60, in short). The resulting distance between the coinage-metal complexes and the surface of C60 (Table 1), is similar to that reported in related 4:1 cocrystals by Dias and coworkers,31 with values of 2.804, 2.860 and 2.894 Å for M = Cu, Ag and Au, respectively (Exp: 2.771, 2.919 and 2.824 Å, respectively). In the 1:1 model, such distances are slightly elongated amounting to 2.818, 2.049 and 3.010 Å, respectively. In order to maximize the intermolecular interaction with the curved π-surface of C60, in the 4:1 ratio, the planar [M3(3,5-(CF3)2Pz)3] complexes are allowed to curve leading to a bowl-shape deviation with a bowl-deep of 0.531, 0.550, and 0.551 Å, for M = Cu, Ag and Au, species respectively (Exp: 0.574, 0.658, 0.567 Å). This planarity deviation is similar in the 1:1 model, amounting to a bowl-depth of 0.531, 0.571 and 0.543 Å, respectively. Different orientations of C60 were evaluated via the 1:1 model, in order to account for the twelve orientations obtained in the X-ray structures.31 An interesting variation of the rotation barrier throughout the series is found, with values of 1.19 kcal∙mol-1 for Cu, 0.39 kcal∙mol-1 for Ag, and 0.53 kcal∙mol-1 for Au species. The M3 located over the 6-membered rings (6MR) from C60, results as the slightly more preferred disposition for the Cu counterpart, where for Ag, a M3-5MR disposition is given. Moreover, Au counterpart is found to prefer both dispositions with a negligible difference (0.05 kcal∙mol-1). Thus, the lighter [M3] counterpart offers better capabilities to hinder the molecular tumbling in the cocrystal aggregates. The resulting structures exhibit a geometrical distortion of coinage-metal complexes in relation to the isolated counterpart, mostly given by the loss of planarity in order to match the geodesic πsurface of fullerene, at [M3]-C60 species. This geometrical distortion is associated with an energetic destabilization accounted for by the deformation energy (ΔEdef) which ranges from 1.5 to 2.0 kcal∙mol-1 for each [M3(3,5-(CF3)2Pz)3] unit at the 4:1 structure, and 1.6 to 2.3 kcal∙mol-1 in the 1:1 species. Hence, such trinuclear complexes provide a versatile moiety which is able to exhibit a concave distortion in order to interact with curved surfaces, which can be further tuned according to the functionalization introduced via the pyrazolate rings. The formation of [M3]-C60 species in solution has been demonstrated via NMR spectroscopy, showing a slight deshielding effect over C60

13C-NMR

signals (142.9 ppm for Cu, and 143.0 for

Ag, vs 142.68 ppm for free C60), suggesting a weak interaction within the adduct.31 Such an 7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

observation indicates the possible role of different weak stabilizing contributions to the overall adduct formation. In this concern, we gain further insights into the nature of the [M3]-C60 interaction, by evaluating the energy of the interaction (ΔEint) in terms of different chemically meaningful quantities within the Energy Decomposition Analysis (EDA) provided by Ziegler and Rauk.40,63,64 In this framework, ΔEint can be decomposed in several terms according to: ΔEint = ΔEPauli + ΔEelstat + ΔEorb + ΔEdisp the ∆EPauli term is related to the destabilizing character owing to the four-electron/two-orbital repulsion between occupied orbitals from the different fragments. The stabilizing ΔEelstat and ΔEorb account for the favorable electrostatic and covalent character of the interaction. Moreover, the dispersion interaction (ΔEdisp) was evaluated via the pairwise correction of Grimme54 (DFT-D3), denoting a stabilizing character. To overcome basis set superposition error (BSSE), the counterpoise method was employed denoting values lower than 1.0 kcal∙mol-1. The relation between the stabilizing terms accounts for the character of the overall interaction, which is conveniently described in terms of relative percentage. In this framework, the coinage-metal complex∙∙∙fullerene interaction energy (ΔEint) in 4:1 [M3]4C60 species is estimated to be about -104.47, for Cu, -98.73 for Ag, and -99.97 kcal∙mol-1 for Au (Table 2), revealing a substantial stabilization in the adduct formation involving the four triangular complexes. Interestingly, the evaluation of the 1:1 model ([M3]-C60) exhibits values given by 25.71, -22.49, and -24.08 kcal∙mol-1, respectively, amounting to about one quarter of that obtained for [M3]4-C60 species, showing that the contribution to ΔEint per [M3] unit remains almost independent of the number of adduct molecules. Thus, this suggests that the number of coinagemetal units involved in the adduct formation is mainly given by the availability of π-surface within the fullerene structure, and the stoichiometry employed in the co-crystallization. According to the relative contribution terms given by ΔEorb, ΔEelstat and ΔEdisp, the character of the stabilizing interaction can be accounted for. For both 4:1 and 1:1 species, ΔEdisp term appears to be the main contribution to the adduct formation, which accounts for more than the 55% of the stabilizing energy, suggesting a van der Waals nature for the [M3]-C60 interaction.31 In addition, the electrostatic character demonstrated by ΔEelstat amounts for more than 24% of the stabilizing contribution, with a covalent character accounting for more than 15%. 8 ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

A comparison between different coinage-metals from the group reveals that the [Ag3]-C60 species exhibit a slightly less favorable situation, owing to the decrease in stabilizing terms of the longrange interaction which results in the lower interaction energy within the -98.73 - -104.47 kcal∙mol-1 range. Interestingly, in the studied species, the ΔEorb and ΔEelstat terms remain similar (±3 kcal∙mol-1), suggesting that in addition to the d10-M/π-C60 interaction, a complementary πPz/π-C60 interaction is also relevant in the adduct formation. The role of [M3] complexes in the spatial charge distribution over the C60 surface was evaluated by using the molecular electrostatic potential (MEP)42 surfaces (Figure 2) drawn within the C60 electron density. This provides further understanding in the electrostatic character of the interaction, accounting for the ΔEelstat term. By comparing the respective potential energy surface for the free C60, characteristic features resulting from the [M3]-C60 interaction are shown, denoting a positive region at the [M3]-C60 interface, and a charge accumulation (negative region) at the complementary sites. For [Cu3]-C60 species, the positive region is located mainly below the Cu(I) centers, which is more extended in Ag(I) and Au(I) species. Interestingly, the charge redistribution upon formation of [M3]-C60 species in all the series is mainly located below the M3-ring, suggesting that the electrostatic character of the interaction accounted for by ΔEelstat mainly originates in the all-metal ring fragment. In this sense, related heterometallic M3 rings

65,66

can be

interesting counterparts for the current adduct formation. Moreover, the major contributor, i.e. ΔEdisp term, amounts to -124.08, -117.07 and -122.76 kcal∙mol-1 throughout the series in [M3]4-C60 species, accounting for the stabilizing character of the interaction. In the 1:1 model, the related values are about one-fourth of that obtained for [M3]4-C60 species, supporting the theory that the contribution from the ΔEdisp term remains almost independent of the number of involved [M3] units, as discussed above for the [M3]-C60 interaction energy. With that in mind, the noncovalent interaction index (NCI)45,46,67 was employed for the studied series in order to assess the spatial distribution of the van der Waals nature accounted for by the ΔEdisp term. The NCI index offers a suitable description of noncovalent interactions based on the reduced density gradient (s(ρ)) which exhibits small values in regions where non-covalent interactions are located, where its stabilizing or destabilizing nature can be addressed by the second eigenvalue of the electron density Hessian (λ2). This accounts for the accumulation (attractive) or depletion (repulsive) of density in the plane perpendicular to the interaction. Thus, the product between ρ(r) and the sign of λ2 has been proposed as a useful descriptor denoting: 9 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

stabilizing (λ2 < 0), London (λ2 ≈ 0) or repulsive-type interactions (λ2 > 0)45,46,67, where, ρ*sign(λ2) ranges from negative to positive values unraveling the nature of the non-covalent interaction. Figure 3 accounts for the regions where non-covalent interactions are relevant, where the ρ*sign(λ2) product ranges from -0.30 to +0.30 a.u. This allows us to unravel the regions involving van der Waals forces between [M3] and C60 in both 4:1 and 1:1 stoichiometries, as regions with ρ*sign (λ2) close to zero. Such regions are strongly related to the M3N6 ring core, thus they involve the coinage-metal ring and the coordinating nitrogen atoms from the trifluoromethyl-pyrazole ligand (3,5-(CF3)2Pz), accounting for both non-covalent π-Pz∙∙∙π-C60 and d10-M∙∙∙π-C60 interactions, where the former comprises a larger region. Actually, the ratio between van der Waals interactions contributed by π-Pz∙∙∙π-C60 and d10-M∙∙∙π-C60 interactions is about 6:3, and thus the organic ring is more relevant in the formation of such interactions, and in turn to the ΔEdisp term. In addition, weak -F∙∙∙C60 interactions are observed. Our results show a similar scenario in both 4:1 and 1:1 models, which is in accordance with the similar ΔEdisp term per [M3] unit. Thus, the relevant van der Waals interactions are contributed mainly by the organic ring, and the electrostatic character displayed by the acidic coinage-metal M3 ring,36,38,39 which ensures the stabilizing character of the adduct formation. This means that further derivatization of the pyrazole ring with aromatic rings can provide larger regions of van der Waals character for adduct formation towards less symmetric fullerene and endohedral metallofullerene species. Further comparison to octaethyl-metalloporphyrins, which are usually employed as efficient cocrystallization agents, shown by the [ZnOEP]-C60 adduct,20–23 unravels a similar stabilizing scenario with and ΔEint of -41.07 kcal∙mol-1, involving ΔEdisp term as the more relevant quantity (55.7%), followed by ΔEelstat (26.4%) and ΔEorb (18.0%), with the repulsive ΔEPauli term amounting to 56.12 kcal∙mol-1. In comparison between [ZnOEP] and [M3], the former exhibits 1:1 or 2:1 adducts with fullerene,23,26 whereas the coinage-metal complexes, adducts of 4:1, have been characterized by Dias, allowing for a larger amount of energy in the interaction for the adduct formation. Thus, [M3(3,5-(CF3)2Pz)3] complexes can be viewed as more versatile counterparts to [MOEP] porphyrins, for further exploration. Furthermore, the interaction of [M3(3,5-(CF3)2Pz)3] complexes to fullerene fragments is evaluated by considering bowl-shaped polyaromatic hydrocarbons resembling both 5MR and 6MR faces,68,69 given by coranulenne (C20H10) and sumanene (C21H12) (Figure 4). Two possible conformations 10 ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

were evaluated in the [M3]-C20H10 and [M3]-C21H12 species, accounting for both convex- and concave-curved π-surface (3a/4a and 3b/4b, respectively in Figure 4), with the former disposition being related to C60. Our results indicate that both conformations are similarly favored with energy differences between 0.12 and 1.69 kcal∙mol-1. Hence, the formation of extended columnar aggregates as observed for a related [Hg3] complex and corannulene derivatives,68,69 showing a favorable interaction in the crystal between [Hg3] towards both convex- and concave-faces of bowl-shaped polyaromatic hydrocarbons, is suggested as a possibility for the studied [M3] complexes. The resulting structures exhibit slightly larger [M3]-π-surface distances than the one observed for [M3]-C60 species. The energy obtained from the interaction (ΔEInt) reveals a more stabilizing situation in comparison to the aforementioned [M3]-C60 species (Table 3), owing to the enhancement of favorable terms. The deformation energy (ΔEdef) increases, which can be , mainly associated to the concave-π-face interaction, owing to the larger distortion, as accounted for in the bowl-depth value. The relative contribution per each stabilizing term remains similar throughout the coinage metal series, where for the convex interaction of the coranulenne (C20H10) adducts (3a), the ΔEElstat term represents 36.0% of the stabilizing quantities, with ΔEDisp decreasing to 47.1%. For the remaining species, i.e. concave-[M3]-C20H10 (3b) and convex/concave-[M3]-C21H12 species (4a/4b), the contribution to ΔEInt is similar to that obtained for [M3]-C60 species. The regions involving van der Waals forces between [M3] and the respective fullerene fragment are illustrated in Figure 5. As occurs with [M3]-C60, such regions are associated mainly with the M3N6 ring and the polyaromatic hydrocarbons (PAH), where the π-Pz∙∙∙π-PAH interaction shows a larger region in contrast to d10-M/π-PAH interactions. In addition, weak -F∙∙∙PAH and -F∙∙∙H interactions are also observed. Hence, the interactions in fragments of fullerenes are of similar nature to those previously mentioned for [M3]-C60 species, where the relevant van der Waals interactions receive a strong contribution from the aromatic ring of the organic ligand belonging to the coinage-metal complex, and the electrostatic character given by the acidic coinage-metal M3 ring. Lastly, the role of relativistic effects was evaluated by comparing calculation at the non-relativistic level of the [M3]4-C60 system (Table S1), which reveals a sizable decrease in the strength of the [M3]4/C60 long-range interaction, as denoted by -78.88 kcal∙mol-1 for M=Cu; -75.03 kcal∙mol-1 for M=Ag; and -78.80 kcal∙mol-1 for M=Au, showing a relevant decrease in the ΔEdisp term for all the

11 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

series, suggesting that relativistic effects are predominant in the stabilization of both lighter and heavier members of the coinage-metal complexes.

12 ACS Paragon Plus Environment

Page 12 of 28

Page 13 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Conclusions The formation of [M3(3,5-(CF3)2Pz)3]-C60 species (M=Cu, Ag and Au), unraveled the versatility of coinage-metal complexes to serve as Buckycatchers. The Energy Decomposition Analysis reveals an interesting interplay between the van der Waals forces provided by the aromatic-ring belonging to the coinage-metal complex, and the electrostatic character given by acidic coinage-metal M3 ring, which ensures the stabilizing character of the adduct formation of similar character along the coinage metal group. Such results provide guidelines for future explorative synthesis efforts seeking to increase the coinage-metal complex∙∙∙fullerene interactions by decorating with aromatic rings the pyrazole ligands, leading to larger van der Waals regions, which could be useful for obtaining novel assemblies involving less symmetric fullerenes and endohedral metallofullerene species. The resulting van der Waals character is retained in the hypothetical systems involving representative fullerene fragments, given by coranulenne (C20H10) and sumanene (C21H12), which act as a source of extended columnar aggregates, owing to favorable interactions between both convex- and concave-faces of such bowl-shaped polyaromatic hydrocarbons, expanding the hostguest chemistry of the studied [M3] complexes. Thus, such trinuclear complexes provide a versatile system which is able to exhibit a convex distortion in order to interact to curved π-surfaces. Interestingly, the nature and strength of the interaction remain almost independent of the number of involved [M3] complexes within C60 fullerene, suggesting that the number of coinage-metal units involved in the adduct formation is mainly provided by the availability of π-surface within the fullerene structure, and the stoichiometry employed in the co-crystallization. Thus, novel cocrystals can be further explored by including coinage metal complexes as co-crystallizing agents, which can be useful for for the formation of extended networks featuring pre-organized domains, involving fullerenes and other interesting π-systems.

13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Supporting Information. . Schematic representation of the bowl-depth in the coinage-metal complex, defined as the distance between two different centroids. Surfaces for the non-covalent index analysis for [M3]4-C60 and [M3]-C60. Surfaces for the non-covalent index analysis for [M3]4-C20H10 and [M3]-C21H12, for M=Cu, Ag, and Au. Energy decomposition analysis for the [M3]4-C60 interaction at the relativistic and non-relativistic level of theory.

AUTHOR INFORMATION

Corresponding Author *Alvaro Muñoz-Castro

Laboratorio de Química Inorgánica y Materiales Moleculares, Universidad Autonoma de Chile, Llano Subercaceaux 2801, San Miguel, Santiago, Chile.

Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.

14 ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Funding Sources The authors thank the financial support from FONDECYT 1140359 and 1180683. Notes Any additional relevant notes should be placed here.

15 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 28

References 1

H. W. Kroto, J. R. Heath, S. C. O’Brien, R. F. Curl and R. E. Smalley, C60: Buckminsterfullerene, Nature, 1985, 318, 162–163.

2

R. Taylor, J. P. Hare, A. K. Abdul-Sada and H. W. Kroto, Isolation, Separation and Characterisation of the Fullerenes C60 and C70: The Third form of Carbon, J. Chem. Soc. Chem. Commun., 1990, 1423.

3

H. W. Kroto, C60: Buckminsterfullerene, The Celestial Sphere that fell to Earth, Angew. Chemie Int. Ed. English, 1992, 31, 111–129.

4

W. Krätschmer, K. Fostiropoulos and D. R. Huffman, The Infrared and Ultraviolet Absorption Spectra of Laboratory-produced Carbon Dust: Evidence for the Presence of the C60 Molecule, Chem. Phys. Lett., 1990, 170, 167–170.

5

H. W. Kroto, A. W. Allaf and S. P. Balm, C60: Buckminsterfullerene, Chem. Rev., 1991, 91, 1213–1235.

6

K. M. Kadish and R. S. Ruoff, Fullerenes: Chemistry, Physics, and Technology, Wiley-Interscience, New York, 2000.

7

A. Hirsch and M. Brettreich, Fullerenes: Chemistry and Reactions, Wiley-VCH, Weinheim, 2005.

8

E. Sheka, Fullerenes: Nanochemistry, Nanophotonics, CRC Press, 2011.

9

A. Rodríguez-Fortea, S. Irle and J. M. Poblet, Fullerenes: Formation, Stability, and Reactivity, Wiley Interdiscip. Rev. Comput. Mol. Sci., 2011, 1, 350–367.

10

P. Jena and A. W. Castleman, Clusters: A Bridge across the Disciplines of Physics and Chemistry, Proc. Natl. Acad. Sci., 2006, 103, 10560–10569.

11

L. T. Scott, A Rational Chemical Synthesis of C60, Science, 2002, 295, 1500–1503.

12

S. Mitra, C. J. Pope, K. K. Gleason, Y. Makarovsky, A. L. Lafleur and J. B. Howard, Synthesis of Fullerenes (C60 AND C70) by Combustion of Hydrocarbons in A Flat Flame Burner, MRS Proc., 1992, 270, 149.

13

M. Mojica, J. A. Alonso and F. Méndez, Synthesis of Fullerenes, J. Phys. Org. Chem., 2013, 26, 526–539.

14

M. Hufnagel, M.-A. Muth, J. C. Brendel and M. Thelakkat, Fullerene-Grafted Copolymers Exhibiting High Electron Mobility without Nanocrystal Formation, Macromolecules, 2014, 47, 2324–2332.

15

C. J. Hawker, A Simple and Versatile Method for the Synthesis of C60 Copolymers, Macromolecules, 1994, 27, 4836–4837.

16

G. Yilmaz, B. Iskin, F. Yilmaz and Y. Yagci, Visible Light-Induced Cationic Polymerization Using Fullerenes, ACS Macro Lett., 2012, 1, 1212–1215.

17

M. Schmittel, B. He and P. Mal, Supramolecular Multicomponent Self-Assembly of Shape-Adaptive Nanoprisms: Wrapping up C60 with Three Porphyrin Units, Org. Lett., 2008, 10, 2513–2516.

Nanomagnetism,

16 ACS Paragon Plus Environment

Nanomedicine,

Page 17 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

18

T. Dinh and Y.-S. Shon, Direct Assembly of Photoresponsive C60−Gold Nanoparticle Hybrid Films, ACS Appl. Mater. Interfaces, 2009, 1, 2699–2702.

19

T. Nakamura, H. Ube, R. Miyake and M. Shionoya, A C60 -Templated Tetrameric Porphyrin Barrel Complex via Zinc-Mediated Self-Assembly Utilizing Labile Capping Ligands, J. Am. Chem. Soc., 2013, 135, 18790–18793.

20

F. Sedona, M. Di Marino, A. Basagni, L. Colazzo and M. Sambi, Structurally Tunable Self-Assembled 2D Cocrystals of C60 and Porphyrins on the Ag (110) Surface, J. Phys. Chem. C, 2014, 118, 1587–1593.

21

D. Sun, F. S. Tham, C. A. Reed, L. Chaker, M. Burgess and P. D. W. Boyd, Porphyrin−Fullerene Host−Guest Chemistry, J. Am. Chem. Soc., 2000, 122, 10704– 10705.

22

R. E. Smalley, Self-assembly of the Fullerenes, Acc. Chem. Res., 1992, 25, 98–105.

23

D. V. Konarev, S. S. Khasanov and R. N. Lyubovskaya, Fullerene Complexes with Coordination Assemblies of Metalloporphyrins and Metal Phthalocyanines, Coord. Chem. Rev., 2014, 262, 16–36.

24

S. Stevenson, G. Rice, T. Glass, K. Harich, F. Cromer, M. R. Jordan, J. Craft, E. Hadju, R. Bible, M. M. Olmstead, K. Maitra, A. J. Fisher, A. L. Balch and H. C. Dorn, Small-bandgap Endohedral Metallofullerenes in High Yield and Purity, Nature, 1999, 401, 55–57.

25

Y. Sado, S. Aoyagi, R. Kitaura, Y. Miyata, E. Nishibori, H. Sawa, K. Sugimoto and H. Shinohara, Structure of Tm@C82 (I) Metallofullerene by Single-Crystal X-ray Diffraction Using the 1:2 Co-Crystal with Octaethylporphyrin Nickel (Ni(OEP)), J. Phys. Chem. C, 2013, 117, 6437–6442.

26

M. M. Olmstead, D. A. Costa, K. Maitra, B. C. Noll, S. L. Phillips, P. M. Van Calcar and A. L. Balch, Interaction of Curved and Flat Molecular Surfaces. The Structures of Crystalline Compounds Composed of Fullerene (C60, C60O, C70, and C120O) and Metal Octaethylporphyrin Units, J. Am. Chem. Soc., 1999, 121, 7090–7097.

27

D. Sun, F. S. Tham, C. A. Reed, L. Chaker and P. D. W. Boyd, Supramolecular Fullerene-Porphyrin Chemistry. Fullerene Complexation by Metalated “Jaws Porphyrin” Hosts, J. Am. Chem. Soc., 2002, 124, 6604–6612.

28

D. Bonifazi, A. Kiebele, M. Stöhr, F. Cheng, T. Jung, F. Diederich and H. Spillmann, Supramolecular Nanostructuring of Silver Surfaces via Self-Assembly of [60]Fullerene and Porphyrin Modules, Adv. Funct. Mater., 2007, 17, 1051–1062.

29

C. Li, X. J. Gao, H. Yao, H. Huang, R. Cui, X. Guo, L. Zhang, B. Liu, B. Xu, W. Shi, J. Dong, L. Feng, X. Gao and B. Sun, Highly delocalized endohedral metal in Gd@C2v(9)-C82 metallofullerenes co-crystallized with α-S8, Nano Res., 2018, 11, 2277–2284.

30

P. Chakraborty, A. Nag, G. Paramasivam, G. Natarajan and T. Pradeep, FullereneFunctionalized Monolayer-Protected Silver Clusters: [Ag29(BDT)12(C60)n]3– (n = 1– 9), ACS Nano, 2018, acsnano.7b07759.

31

N. B. Jayaratna, M. M. Olmstead, B. I. Kharisov and H. V. R. Dias, Coinage Metal Pyrazolates [(3,5-(CF3)2Pz)M]3 (M = Au, Ag, Cu) as Buckycatchers, Inorg. Chem., 17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

2016, 55, 8277–8280. 32

G. Yang and R. G. Raptis, Supramolecular Assembly of Trimeric Gold(I) Pyrazolates through Aurophilic Attractions, Inorg. Chem., 2003, 42, 261–263.

33

R. N. McDougald, B. Chilukuri, H. Jia, M. R. Perez, H. Rabaâ, X. Wang, V. N. Nesterov, T. R. Cundari, B. E. Gnade and M. A. Omary, Molecular and Electronic Structure of Cyclic Trinuclear Gold(I) Carbeniate Complexes: Insights for Structure/Luminescence/Conductivity Relationships, Inorg. Chem., 2014, 53, 7485– 7499.

34

M. Veronelli, N. Kindermann, S. Dechert, S. Meyer and F. Meyer, Crowning of Coinage Metal Pyrazolates: Double-Decker Homo- and Heteronuclear Complexes with Synergic Emissive Properties, Inorg. Chem., 2014, 53, 2333–2341.

35

J.-P. Zhang, Y.-B. Zhang, J.-B. Lin and X.-M. Chen, Metal Azolate Frameworks: From Crystal Engineering to Functional Materials, Chem. Rev., 2012, 112, 1001– 1033.

36

L. Hou, W.-J. Shi, Y.-Y. Wang, H.-H. Wang, L. Cui, P.-X. Chen and Q.-Z. Shi, Trinuclear-based Copper(I) Pyrazolate Polymers: Effect of Trimer πAcid···Halide/Pseudohalide Interactions on the Supramolecular Structure and Phosphorescence, Inorg. Chem., 2011, 50, 261–270.

37

E. Zangrando, M. Casanova and E. Alessio, Trinuclear Metallacycles: Metallatriangles and Much More, Chem. Rev., 2008, 108, 4979–5013.

38

M. A. Omary, M. A. Rawashdeh-Omary, M. W. A. Gonser, O. Elbjeirami, T. Grimes, T. R. Cundari, H. V. K. Diyabalanage, C. S. P. Gamage and H. V. R. Dias, Metal Effect on the Supramolecular Structure, Photophysics, and Acid−Base Character of Trinuclear Pyrazolato Coinage Metal Complexes †, Inorg. Chem., 2005, 44, 8200–8210.

39

A. Burini, J. P. Fackler, R. Galassi, T. A. Grant, M. A. Omary, M. A. RawashdehOmary, B. R. Pietroni and R. J. Staples, Supramolecular Chain Assemblies Formed by Interaction of a π Molecular Acid Complex of Mercury with π-Base Trinuclear Gold Complexes, J. Am. Chem. Soc., 2000, 122, 11264–11265.

40

M. von Hopffgarten and G. Frenking, Energy Decomposition Analysis, Wiley Interdiscip. Rev. Comput. Mol. Sci., 2012, 2, 43–62.

41

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, 12758–68.

42

T. Brinck, in Theoretical and Computational Chemistry, ed. C. Párkányi, 1998, pp. 51–93.

43

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.

44

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. 18 ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

45

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.

46

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.

47

G. Te Velde and E. J. Baerends, Precise Density-Functional Method for Periodic Structures, Phys. Rev. B, 1991, 44, 7888–7903.

48

C. Fonseca Guerra, J. G. Snijders, G. te Velde and E. J. Baerends, Towards an orderN DFT method, Theor. Chem. Accounts Theory, Comput. Model. (Theoretica Chim. Acta), 1998, 99, 391–403.

49

S. Grimme, Semiempirical GGA-type Density Functional constructed with a Longrange Dispersion Correction, J. Comput. Chem., 2006, 27, 1787–1799.

50

A. Suvitha and N. S. Venkataramanan, Trapping of Organophosphorus Chemical Nerve Agents by Pillar[5]arene: A DFT, AIM, NCI and EDA analysis, J. Incl. Phenom. Macrocycl. Chem., 2017, 87, 207–218.

51

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.

52

S. Grimme, Accurate Description of van der Waals Complexes by Density Functional Theory including Empirical Corrections, J. Comput. Chem., 2004, 25, 1463–1473.

53

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.

54

S. Grimme, Density Functional Theory with London Dispersion Corrections, Wiley Interdiscip. Rev. Comput. Mol. Sci., 2011, 1, 211–228.

55

E. R. Johnson and A. D. Becke, A Post-Hartree–Fock Model of Intermolecular Interactions, J. Chem. Phys., 2005, 123, 024101.

56

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.

57

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.

58

J. Contreras-García, E. R. Johnson, S. Keinan, R. Chaudret, J. P. Piquemal, D. N. Beratan and W. Yang, NCIPLOT: A Program for Plotting Noncovalent Interaction Regions, J. Chem. Theory Comput., 2011, 7, 625–632.

59

J. S. Murray and P. Politzer, The Electrostatic Potential: An Overview, Wiley Interdiscip. Rev. Comput. Mol. Sci., 2011, 1, 153–163.

60

E. Bartashevich and V. Tsirelson, A Comparative View on the Potential acting on an Electron in a Molecule and the Electrostatic Potential through the Typical Halogen 19 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Bonds, J. Comput. Chem., 2018, 39, 573–580. 61

J. S. Murray and P. Politzer, Molecular Electrostatic Potentials and Noncovalent Interactions, Wiley Interdiscip. Rev. Comput. Mol. Sci., 2017, 7, e1326.

62

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.

63

K. Morokuma, Molecular Orbital Studies of Hydrogen Bonds. III. C=O···H–O Hydrogen Bond in H2 CO···H2O and H2CO···2H2O, J. Chem. Phys., 1971, 55, 1236–1244.

64

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.

65

A. A. Mohamed, R. Galassi, F. Papa, A. Burini and J. P. Fackler, Gold(I) and Silver(I) Mixed-Metal Trinuclear Complexes: Dimeric Products from the Reaction of Gold(I) Carbeniates or Benzylimidazolates with Silver(I) 3,5-Diphenylpyrazolate, Inorg. Chem., 2006, 45, 7770–7776.

66

R. Galassi, M. M. Ghimire, B. M. Otten, S. Ricci, R. N. McDougald, R. M. Almotawa, D. Alhmoud, J. F. Ivy, A.-M. M. Rawashdeh, V. N. Nesterov, E. W. Reinheimer, L. M. Daniels, A. Burini and M. A. Omary, Cupriphication of Gold to sensitize d10–d10 Metal–Metal Bonds and Near-unity Phosphorescence Quantum Yields, Proc. Natl. Acad. Sci., 2017, 201700890.

67

J. Contreras-García, E. R. Johnson, S. Keinan, R. Chaudret, J.-P. Piquemal, D. N. Beratan and W. Yang, NCIPLOT: A Program for Plotting Noncovalent Interaction Regions, J. Chem. Theory Comput., 2011, 7, 625–632.

68

A. S. Filatov, A. K. Greene, E. A. Jackson, L. T. Scott and M. A. Petrukhina, Molecular Curvature Tradeoffs: Bending a Planar Trimercury Unit over Bowlshaped Polyaromatic Hydrocarbons, J. Organomet. Chem., 2011, 696, 2877–2881.

69

A. S. Filatov, E. A. Jackson, L. T. Scott and M. A. Petrukhina, Foregoing Rigidity to Achieve Greater Intimacy, Angew. Chemie Int. Ed., 2009, 48, 8473–8476.

20 ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 1. Representation of the structures for 4:1 and 1:1 ratios between coinage metal complex and fullerene, given by [M3]4-C60 (1) and [M3]-C60 (2), for M=Cu, Ag, and Au.

21 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. Molecular electrostatic potential surface at C60, and the respective the [M3]4-C60 (1), above, and [M3]-C60 (2), below, depicting the variation upon adduct formation.

22 ACS Paragon Plus Environment

Page 22 of 28

Page 23 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 3. Surfaces for the non-covalent index analysis for [Au3]4-C60 (1), above, and [Au3]C60 (2), below, as representative examples. For a complete picture involving M=Cu, Ag and Au complexes, see supporting information.

Figure 4. Representation of the structures for coranulenne (C20H10) and sumanene (C21H12), above, and the respective coinage metal complex and in both conformation given by 23 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

convex- and concave--surface, below, for [M3]4-C20H10 (3a, 3b) and [M3]-C21H12 (4a, 4b), for M=Cu, Ag, and Au.

Figure 5. Surfaces for the non-covalent index analysis for [M3]4-C20H10 (3a, 3b) and [M3]C21H12 (4a, 4b), for M=Cu, Ag, and Au. For a complete picture involving M=Cu, Ag and Au complexes, see supporting information.

24 ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Tables Table 1. Selected structural parameters for [M3]4-C60 (1) and [M3]-C60 (2) species, denoting the [M3]-π-surface distance and [M3] bowl-depth distortion. Values in Angstrom (Å). In addition, parameters for [M3]-C20H10 and [M3]-C21H12 structures are also given, for both conformations. Cu Ag Au [M3]4-C60 (1) M-π-surface Calc 2.804 2.860 2.920 Exp 2.771 2.919 2.824 aBowl-depth Calc 0.531 0.603 0.551 Exp 0.574 0.658 0.567 [M3]-C60 (2) M-π-surface aBowl-depth

2.818 0.531

3.049 0.571

[M3]-C20H10 (3a) M-π-surface 2.777 3.083 aBowl-depth 0.174 0.245 [M3]-C20H10 (3b) M-π-surface 3.758 4.193 aBowl-depth 0.464 0.370 [M3]-C21H12 (4a) M-π-surface 2.911 3.176 aBowl-depth 0.143 0.623 [M3]-C21H12 (4b) M-π-surface 3.825 3.967 aBowl-depth 0.661 0.628 aBowl-depth obtained as the difference between supporting information).

3.010 0.543

3.093 0.237 4.205 0.171 3.174 0.409 3.949 0.412 M3-centroid and 4-Pz-centroid (see

Table 2. Energy decomposition analysis for the [M3]4-C60 and [M3]-C60 interaction in 1 and 2, respectively. Values in kcal∙mol-1.

[M3]4-C60 (1) ΔEdef ΔEPauli

Cu

Ag

Au

6.42

6.10

7.96

123.04

117.96

128.39

25 ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ΔEElstat ΔEOrb ΔEDisp ΔEInt [M3]-C60 (2) ΔEdef ΔEPauli ΔEElstat ΔEOrb ΔEDisp ΔEInt

-64.41 -39.02 -124.08 -104.47

28.3% 17.2% 54.5%

-62.55 -37.07 -117.07 -98.73

28.9% 17.1% 54.0%

Page 26 of 28

-66.56 -39.04 -122.76 -99.97

1.69

1.64

2.26

27.84 -14.38 -9.23 -29.94 -25.71

16.58 -8.08 -6.02 -24.97 -22.49

23.35 -11.71 -7.54 -28.18 -24.08

26.9% 17.2% 55.9%

20.7% 15.4% 63.9%

29.1% 17.1% 53.8%

24.7% 15.9% 59.4%

Table 3. Energy decomposition analysis for the [M3]4-C20H10 and [M3]-C21H12 interaction in 3 and 4, respectively, in both concave- (3a/4a) and convex-π-surface (3b/4b) conformations. Values in kcal∙mol-1. Cu [M3]-C20H10 (3a) ΔEdef ΔEPauli ΔEElstat ΔEOrb ΔEDisp ΔEInt

[M3]-C20H10 (3b) ΔEdef ΔEPauli ΔEElstat ΔEOrb ΔEDisp ΔEInt

Ag

Au

1.49

1.87

1.94

53.2 -31.9 -15.1 -38.8 -32.62

47.1 -27.8 -12.7 -38.8 -32.17

55.5 -31.1 -14.7 -40.9 -31.18

37.2% 17.6% 45.3%

35.1% 16.0% 48.9%

Cu

Ag

Au

2.96

2.86

2.20

39.9 -22.4 -10.2 -40.8 -33.37

40.9 -23.2 -10.1 -40.9 -33.32

41.6 -21.8 -9.9 -42.9 32.984523

30.5% 13.9% 55.6%

31.3% 13.6% 55.1%

26 ACS Paragon Plus Environment

35.9% 16.9% 47.2%

29.2% 13.2% 57.6%

Page 27 of 28 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

9

[M3]-C21H12 (4a) ΔEdef ΔEPauli ΔEElstat ΔEOrb ΔEDisp ΔEInt

[M3]-C21H12 (4b) ΔEdef ΔEPauli ΔEElstat ΔEOrb ΔEDisp ΔEInt

Cu

Ag

Au

1.12

3.99

3.01

52.1 -29.9 -15.0 -38.7 -31.45

40.9 -23.2 -12.1 -39.5 -33.82

40.8 -19.0 -10.7 -42.5 -31.37

35.8% 17.9% 46.3%

31.0% 16.1% 52.8%

Cu

Ag

Au

4.81

4.11

2.81

40.1 -20.8 -11.3 -40.7 -32.68

40.2 -22.9 -12.0 -39.2 -33.93

39.3 -18.3 -10.4 -41.9 -31.29

28.5% 15.6% 55.9%

31.0% 16.2% 52.9%

27 ACS Paragon Plus Environment

26.3% 14.8% 58.8%

25.9% 14.7% 59.4%

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

TOC Graphic Title: Formation of Coinage-Metal∙∙∙Fullerene Adducts. Evaluation of the Interaction Nature between Triangular Coinage Metal Complexes (M3=Cu, Ag and Au) and C60 through relativistic DFT calculations Carolina Olea Ulloa, Miguel Ponce-Vargas, Alvaro Muñoz-Castro

28 ACS Paragon Plus Environment

Page 28 of 28