Next Article in Journal
Triboelectric and Piezoelectric Nanogenerators for Self-Powered Healthcare Monitoring Devices: Operating Principles, Challenges, and Perspectives
Next Article in Special Issue
Size Effect of Electrical and Optical Properties in Cr2+:ZnSe Nanowires
Previous Article in Journal
Classical Analog and Hybrid Metamaterials of Tunable Multiple-Band Electromagnetic Induced Transparency
Previous Article in Special Issue
First-Principles Molecular Dynamics Simulations on Water–Solid Interface Behavior of H2O-Based Atomic Layer Deposition of Zirconium Dioxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Distributed Polarizability Model for Covalently Bonded Fullerene Nanoaggregates: Origins of Polarizability Exaltation

by
Denis Sh. Sabirov
* and
Alina A. Tukhbatullina
Laboratory of Mathematical Chemistry, Institute of Petrochemistry and Catalysis UFRC RAS, 450075 Ufa, Russia
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(24), 4404; https://doi.org/10.3390/nano12244404
Submission received: 24 November 2022 / Revised: 6 December 2022 / Accepted: 7 December 2022 / Published: 9 December 2022
(This article belongs to the Special Issue First-Principles Investigations of Low-Dimensional Nanomaterials)

Abstract

:
Polarizability exaltation is typical for (C60)n nanostructures. It relates to the ratio between the mean polarizabilities of (C60)n and C60: the first one is higher than the n-fold mean polarizability of the original fullerene. This phenomenon is used in the design of novel fullerene compounds and the understanding of its properties but still has no chemical rationalization. In the present work, we studied the distributed polarizability of (C60)2 and isomeric (C60)3 nanoaggregates with the density functional theory method. We found that polarizability exaltation increases with the size of the nanostructure and originates from the response of the sp2-hybridized carbon atoms to the external electric field. The highest contributions to the dipole polarizability of (C60)2 and (C60)3 come from the most remote atoms of the marginal fullerene cores. The sp3-hybridized carbon atoms of cyclobutane bridges negligibly contribute to the molecular property. A similar major contribution to the molecular polarizability from the marginal atoms is observed for related carbon nanostructures isomeric to (C60)2 (tubular fullerene and nanopeanut). Additionally, we discuss the analogy between the polarizability exaltation of covalently bonded (C60)n and the increase in the polarizability found in experiments on fullerene nanoclusters/films as compared with the isolated molecules.

1. Introduction

Fullerene multicage compounds are an interesting class of carbon-rich compounds, whose molecules contain two and more fullerene cores [1,2,3]. The cores may be directly connected via cyclobutane motifs resulting from the chemical reaction of [2+2] addition between the double bonds of the fullerene [4,5]. In the case of the C60 fullerene, this reaction occurs via the 6.6 bonds (this designation means that the bond is shared by two hexagons of the C60 skeleton). Thus, there is only one synthesized isomer of the [2+2] dimer (C60)2. The next addition of the fullerene cores is possible in different positions because in the (C60)2 dimer, the 6.6 bonds become inequivalent due to symmetry loss as compared with original C60. Such regioisomeric trimers were synthesized and studied experimentally and theoretically [6,7]. As found, molecular and physicochemical properties strongly depend on the addition pattern in fullerene trimers (C60)3.
Covalently bonded C60 fullerene [2+2] aggregates are produced using the mechanochemical reaction induced by potassium/sodium cyanide or solid aromatic amines under the “high-speed vibration milling” conditions [2,4,6,8]. These ways lead to a fullerene dimer and trimers, which are large molecules visualized with scanning tunneling microscopy [6]. Indeed, these molecules are unusual due to their shape and size. They are rigid structures with no possible conformational transitions with distinct symmetries and topological properties [9]. This makes them potentially applicable in carbon nanotechnologies (e.g., for qubits [10,11] and molecular materials with nonlinear optic properties [12]). From a fundamental point of view, fullerene oligomers are simultaneously molecules and nano-objects. Unfortunately, this interesting feature creates some obstacles for chemical studies of these molecules, even for chemical identification. Fullerene oligomers larger than (C60)3 were detected only with mass spectrometry techniques that provide only the contents of the substances—not chemical structures [13,14].
As covalently bonded fullerene aggregates are promising for applications, there is a necessity to know their behavior in electric fields [15]. Previously, we systematically studied the dipole polarizability (hereinafter, polarizability) of various dimers, trimers, and higher oligomers of C60, C70, and other fullerenes. This quantity corresponds to the molecule’s response to the external electric field; it is applied in physical chemistry to assess reactivity, intermolecular interactions, optical properties, collisions, and so on [16]. To calculate the polarizability, we used density functional theory methods. The calculated values were interesting not only themselves—we found the superadditivity of the mean polarizabilities of (C60)n called exaltation of polarizability [3,5,7]. We observed the positive deviation of the calculated values α[(C60)n] from nα(C60). We were not stopped on the explanation of this phenomenon and focused on demonstrating that polarizability exaltation is typical for all classes of fullerene multicage compounds regardless of the chemical nature of the bridges between the fullerene units. After scrutinizing our works, Swart and van Duijnen [17] provided an explanation for the exaltation based on an analogy with the classical electric theory. They considered a fullerene dimer as a system of interacting dielectric spheres. This allowed for distinguishing parallel ( α ) and perpendicular components ( α ) of mean polarizability (relative to the applied electric field) of such system, and these contributions could be estimated within the classic theory if key input parameters are extracted from the relevant quantum chemical calculations. In general, ( α ) and ( α ) differ so that the exaltation arises. The model explains the polarizability exaltation for (C60)2 but works with assumptions in the case of higher oligomers (as the number of cages is greater, and we have shown that only the most remote cages define the exaltation value). Thus, there is still no chemical explanation for the polarizability exaltation of fullerene oligomers.
Recently, the distributed polarizability models have been applied in fullerene chemistry. This approach allows for assessing the contributions of the atoms to the molecular polarizability using the calculated atomic charges with and without an applied electric field [18]. In general, it is applicable to various chemical structures, from small molecules to proteins and nano-objects. As for fullerenes, this approach was applied to their endohedral complexes [19] and exohedral derivatives [20]. We believe that it could be useful for rationalizing the peculiarities of the dipole polarizability of fullerene oligomers.
In the present work, we study the dipole polarizability of the fullerene dimer (C60)2 and trimers (C60)3 within the distributed polarizability model. Additionally, we discuss the relevance of our computed polarizability exaltation of (C60)n to experimental data and theoretical results of other groups.

2. Computational Details

The TPSS/TZVP density functional theory method was used for calculations (program Gaussian 09 Rev. D3.01 [21]). Our choice is based on our previous studies of fullerene and its derivatives whereby this method demonstrated high reliability in describing mean polarizabilities of the molecules. The structures of (C60)n and C60 itself as the reference compound were completely optimized. The confirmation that the obtained structures correspond to the minima of the potential energy surfaces was performed based on their hessian, which had no imaginary frequencies. The calculations of the dipole polarizability were performed according to the distributed polarizability technique [18]. In general, this methodology uses a finite-field approach, and its details are below. The optimized structures and output data for distributed polarizability calculations can be found in Supplementary Materials (Tables S1–S10).
The mean polarizabilities of the molecules are analytically calculated. To assess the atomic contributions to the molecular polarizability, the atomic partitions into the electronic density are calculated according to the Hirshfeld scheme [22]. The Hirshfeld atomic charges are computed in four modes for each structure: with fields Fγ having one nonzero component along axis γ (γ = x, y, or z) and one calculation without the field. The γ component of the distributed dipole moment for atom i for each field Fγ is calculated as
μ γ i = q i R γ i + δ μ γ i
where qi and Rγi are the Hirshfeld atomic charges and Cartesian coordinates of the atoms. The δμγi values are taken as is from the Gaussian output file in a.u. The γγ component of the contribution of the i-th atom to the mean polarizability of the molecule follows from the definition of polarizability (below, μind is the induced dipole moment):
μ i n d = α F
and equals
α γ γ i = μ γ i F γ μ γ i 0 F γ
where Fγ = −0.001 a.u. (the recommended value [19]). The isotropic value corresponding to atom i equals:
α i = 1 3 α x x i + α y y i + α z z i
The summation of atomic values αi for all atoms of the nanostructure gives its molecular polarizability:
α C 60 n = i = 1 60 n α i
The polarizability exaltation of (C60)n is calculated as the difference between the actual molecular polarizability (obtained according to Equation (5)) and the corresponding additive value. The last one is based on the proposition that the polarizability of (C60)n results from n molecular polarizabilities of original C60 (this reference value equals 80.3 Å3 as calculated with the TPSS/TZVP method):
Δ α C 60 n = α C 60 n n α C 60

3. Results

The general formula of the fullerene oligomers studied in the present work is shown in Figure 1. In this study, we focus on the experimentally obtained (C60)2 and (C60)3 structures. For designations of trimers, we use rational nomenclature designations to specify the addition patterns (Figure 1). Note that in the (C60)2 molecule, there are nine nonequivalent 6.6 bonds, which are able to attach chemical groups (cis-1, cis-2, cis-3, ee, ef, trans-4, trans-3, trans-2, and trans-1). As C60 is a bulky molecular fragment, only six positions are open for attaching the next C60 core to (C60)2. These are cis-2, ee, ef, trans-4, trans-3, trans-2, and trans-1 [23]. Herewith, the addition to the equatorial bonds ee and ef leads to the same structure designated as e-(C60)3. In the case of the cis-2 addition, the marginal C60 cores are arranged too close so that they are bridged (also via a cis-2 pattern). This trimer is called cyclic, and structurally, it stays apart from the other (C60)3, which have the central and marginal fullerene cores. Figure 1 shows only the bonds taking part in the addition reactions resulting in the fullerene trimers. All of the discussed structures were synthesized and separated as individual substances [2,4,6,8] except for trans-1-(C60)3. The latter structure is least stable among (C60)3. However, there is an experimental detection of the trans-1 addition pattern in (C60)n polymers obtained inside the carbon nanotubes [24].
The optimized geometries of the fullerene compounds are collected in Supplementary Materials. The TPSS/TZVP-calculated values on (C60)2 and (C60)3 polarizability are shown in Table 1. According to the additive scheme (Equation (6)), the polarizability exaltation is typical for all (C60)3. It means that the mean polarizabilities of the (C60)n molecules are higher than the n-fold mean polarizability of the original fullerene (n = 2 or 3). These computational results agree with the previous estimates with the PBE/3ζ method [5,7].
Table 1 also contains the results of the distributed polarizability calculations, that is, contributions to the molecular mean polarizability from the central and marginal cores. These values have been obtained quantum-chemically as the sums of the atomic contributions to α[(C60)n]. The atomic contributions themselves are collected in Supplementary Materials and visualized as the polarizability maps in Section 4.2. In this section, the ball-and-cylinder models of the nanostructures under study could be found.

4. Discussion

4.1. Relevance to the Experimental Data on the C60 Polarizability

We previously reviewed experimental and theoretical works on fullerene polarizability (see, e.g., our work [7] and some key experimental works in this field [25,26,27,28,29,30]). Based on that analysis, the mean polarizability of C60 is approximately 80 Å3, so our computed value of 80.3 Å3 perfectly agrees. As for the (C60)2 and (C60)3 molecules, there are no experimental data on their dipole polarizabilities. We consider that the absence of the experiments is due to the facile dissociation of such fullerene compounds under optical treatments required for typical polarizability measurements. Nevertheless, we would like to discuss the relevance of the computed polarizability exaltation to the experimental α(C60) values deduced from the experiments on C60 in different states.
Munn and Petelenz [31] found that the measured α(C60) values increase from the isolated molecules to solid: ~76.5 Å3 for the molecules in the gas phase < ~79 Å3 for fullerene clusters containing several C60 molecules < ~85 Å3 for the fullerene films. Usually, such a mismatch between the polarizability of the isolated molecules α and the effective molecular polarizability αeff deduced from solid-state measurements is due to (i) the changes in molecular structure under the transit to the condensed state, (ii) compression of electronic clouds of the molecules, (iii) inducing heterogeneous electric fields in the formed crystals, and so on [32]. All mentioned points are not applicable to fullerene: as is known, (i) C60 keeps the initial molecular geometry in a crystal (fullerite), (ii) compressing electronic clouds must decrease the resulting polarizability, and (iii) the emergence of heterogeneous electric fields in fullerite has low probability due to the high symmetry of the molecule. The authors [31] explained the mentioned increase with the generation of the instantaneous charges on the fullerene cages in neighboring crystal lattice nodes and provided a relation for assessing the polarizability exaltation effect. Their model indicates the possibility of 30% of polarizability gain in crystals as compared with the isolated C60 molecules. Note that the instantaneous charges invoked for explaining aggregate-state-dependent polarizability exaltation of C60 are not observables.
In our calculations (Table 1), the mean polarizabilities of C60 moiety deduced from the calculated (C60)n are ~9…15% higher than the computed α(C60) (the range of percentages corresponds to (C60)2trans-1-(C60)3). This increase corresponds to the observed one for fullerene samples. However, we compare similar but not the same cases. In this work, we deal with the chemically bonded C60 cores. Thus, we hypothesize that the polarizability exaltation of (C60)2 and higher chemically bonded fullerene oligomers is analogous to increasing the mean polarizability of C60 from gas to crystalline state. We assume that the nature of both polarizability exaltations relates to interacting π-electron systems of the C60 cores regardless of whether they are bonded with chemical bonds or intermolecular forces. This hypothesis will be tested with relevant computational techniques in our further works.
Previously, we studied the influence of molecular topology and molecular geometry on the mean polarizability of isomeric (C60)6 [33]. As found, the geometric factor has a major effect on α: the resulting α[(C60)6] values increase with the maximal remoteness of the C60 cores in the nanostructure. The DFT computations of the present study fit into this regularity: we demonstrate it with the polarizability exaltation values (Figure 2). However, the molecular topology also has an effect. This follows from the mean polarizability of the cyclic trimer, which falls out the regularity. Topologically [9], this trimer stays apart from other (C60)3 due to its closed structure.
Note that Δα values relate to a rough estimation of the polarizability effects of (C60)n. They are based on the assumption that all fullerene cores equally contribute to the molecular property, but this holds true only for cyclic (C60)3. To look deeper into the matter of the fullerene nanostructures, we apply the distributed polarizability model below.

4.2. Distributed Polarizability Model of (C60)3: The Responses of the Cores Differ

In the distributed polarizability computations, we obtained the atomic contributions to the molecular mean polarizabilities of the (C60)3 trimers. These data allow for assessing the contributions of the fullerene cores to the molecular property (α). We found that, in general, the contribution of the central core αcentr (here, we do not treat (C60)2 and cyclic (C60)3, that is, the molecules, whereby the cores are identical) is pronouncedly lower as compared with the marginal ones (Table 1). Only in the most compact nanoaggregates cyclic and e-(C60)3, αcentr > α(C60). Other structures demonstrate diminishing contributions αcentr relative to the mean polarizability of the original fullerene, α(C60). In the case of the marginal fullerene cores, αmarg > α(C60) for all studied nanostructures (including (C60)2, which could be considered to be constructed with two marginal cores), and this enlargement varies in the range of 6.97…24.73 Å3.
It is obvious that the mentioned effects in (C60)3 depend on the shape of the nanostructures. A decrease in the polarizability contributions of the central cores and an increase in the polarizability contributions of the marginal ones become greater with the distance between the last ones (Figure 3). Accordingly, the marginal cores play a crucial role in the polarizability exaltation effects as their contributions dominate.
There is an analogy between the fullerene derivatives that belong to different classes. A similar situation is observed in the case of fullerene bisadducts C60X2 (X = O and CH2) [20]: the contribution of the C60 moiety in these compounds goes down when increasing the distance between X. In our case, the peripheral fullerene cores play the role of addends X (i.e., X = C60).

4.3. Distributed Polarizability Maps of (C60)3

At the last stage of detailing the polarizability of fullerene covalently bonded nanoaggregates, we analyze the atomic contributions themselves. The corresponding numerical values are presented in Supplementary Materials. Here, it is convenient to visualize them as the colored maps (Figure 4).
The maps were drawn from the following considerations. To reveal regularities general for all (C60)3, we selected in each nanostructure the atoms with the highest contributions and took them as 100%. Other atoms, which have lower contributions, were discriminated depending on the percentage relative to the highest contribution in the structure (Figure 4). We must work with percentages because maximal contributions are not the same in different (C60)3.
According to the performed computations, the most polarizable atoms in (C60)3 are located at the margins of the marginal fullerene cores (Figure 4, red atoms). It means that the atomic charges are induced at these atoms when the nanostructure interacts with the electric field. When moving to the central cores of (C60)3, the atomic contributions to the molecular property become lower. The fact that the maximal charges should arise on the most distant atoms allows for rationalizing the size dependence of the polarizability exaltation of (C60)3. Indeed, any dipole moment (both permanent and induced as in Equation (2)) depends on the distance L between the centers of the opposite-sign charges in the molecule:
μ i n d = Δ q L
where Δq is the charge transfer value [34]. Thus, according to the distributed polarizability calculations, the charges emerge on the most distant atoms of (C60)3. This is reflected with the polarizability exaltation effect increasing with the linear size of the nanostructure (and the remoteness of the marginal fullerene cores plays a parameter of the size).
Further, we pay attention to the atoms with the smallest positive atomic contributions to the molecular polarizability of (C60)3 (Figure 4, beige atoms). These atoms are mainly sp3-hybridized atoms of the cyclobutane moieties that connect fullerene units in one nanostructure. We were surprised to find in the studied nanostructures the atoms with negative atomic contributions. Indeed, dipole polarizability is always a positive value. However, this molecular property may be “nonuniformly” distributed over the atoms. In the classic theory of chemical structure, polarizability is associated with the volume of the molecule’s electronic cloud (in favor of it, α has the dimension of volume in centimeter–gram–second system of units). We believe that the negative atomic contributions relate to the shift of the electronic density to the other moieties of the molecule. Anyway, the atoms with nonzero contributions are sp2-hybridized, and their contributions relate to the response of the rich π-electron systems of the fullerene cores. Thus, atoms of these types are responsible for the polarizability exaltation of (C60)n. In contrast, the site of the connection of the fullerene units has no π-electrons, and therefore, their contributions are negligible. A similar situation is observed for other chemical objects. For example, the dipole polarizability of nanodiamonds [35] consisting of only sp3-hybridized carbon atoms linearly increases with the size of the particle; that is, there is no polarizability exaltation in sp3-carbon nanostructures.

4.4. Prospective: From Fundamental Properties to Nanoapplications

The polarizability exaltation of the fullerene dimer and oligomers has been known for almost 10 years and studied with quantum-chemical and classical approaches [5,7,17]. The present computational study allows for deeper understanding this phenomenon in fundamental and applied aspects.
For this purpose, we compare three isomeric all-carbon nano-objects: [2 + 2] dimer, tubular fullerene, and nanopeanut (Figure 5). The two last structures represent convenient model objects for structural studies [9,36,37,38,39]. Their distributed polarizability maps have been obtained similarly to the title compounds. Interestingly, the marginal atoms of the two novel all-carbon objects also provide the major contributions to the molecular polarizability. Herewith, all three structures differ: the structure of the fullerene dimer has been discussed above; the C120 fullerene consists of the hexagons and isolated pentagons since a nanopeanut’s structures include heptagons in the fusion motif. In all cases, marginal carbon atoms make decisive contributions to the molecular polarizability, and it seems to be a common property of carbon nanostructures. Central atoms of these nanostructures have smaller contributions.
This theoretical study is a part of our ongoing research program on the polarizability of fullerene compounds. Besides the fundamental interest in the relations between their chemical structure and polarizability parameters [40], it has the applied aspects. For example, dipole polarizability plays a crucial role in assessing the collisions of fullerene nanoclusters in the gas phase [41] and their ordering in films [42,43]. As recently found, fullerene dimers demonstrate catalytic properties due to the polarizability exaltation [44,45]. As for nanodevices, there are examples of thermally driven fullerene-based nanocars [46], but their analogues, which could be manipulated with electric fields [47], are a matter of the future. Here, dipole polarizability and its exaltation could provide novel insights into the design of the fullerene-based nanostructures with a high response to external electric fields. Dipole polarizability estimates have already been used in the synthetic chemistry of fullerene-based materials from C60 and C70 under the action of various fields [48,49,50,51].

5. Conclusions

The polarizability exaltation of (C60)n is a phenomenon that consists of the relation α[(C60)n] > α(C60). Our quantum-chemical computations within the distributed polarizability model allows for assessing the atomic contributions to the molecular polarizability of (C60)n. Using (C60)3 as the simplest example, we have shown that the polarizability exaltation of covalently bonded C60 nanoaggregates originates from the response of the sp2-hybridized carbon atoms to the external electric field. Atoms of this type are the centers, where facilely polarizable π-electrons are localized. The biggest contributions to the dipole polarizability of (C60)3 come from the most remote atoms of the marginal fullerene cores. This explains why the polarizability effect increases with the remoteness of the marginal fullerene cores in (C60)3.
We point the analogy between the polarizability exaltation of covalently bonded (C60)n and the increase in the polarizability found in experimental studies of fullerene nanoclusters and solids as compared with the isolated molecules. Indeed, both carbon nanomaterials contain rich π-electron systems. The difference is in the presence or absence of the covalent bonds between the fullerene cores in the mentioned structures. As our computations reveal that the π-electron systems play a decisive role for the polarizability exaltation, we assume the common nature of these two phenomena.
The major contributions of marginal atoms to the molecular polarizability is also typical for other carbon nanostructures (tubular fullerene and nanopeanut). It seems that oblong carbon nanostructures share this feature. We believe that it could be useful in material science and nanotechnology.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12244404/s1, Tables S1–S10: Cartesian coordinates of the carbon nanostructures; Tables S11 and S12: Atomic contributions to the mean polarizability of C60, (C60)2, (C60)3, and related nanostructures C120.

Author Contributions

Conceptualization, D.S.S.; methodology, D.S.S. and A.A.T.; validation, D.S.S. and A.A.T.; formal analysis, D.S.S. and A.A.T.; investigation, D.S.S. and A.A.T.; writing—original draft preparation, D.S.S. and A.A.T.; visualization, A.A.T.; project administration, D.S.S.; funding acquisition, D.S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the President’s Grants Council (grant number MD-874.2021.1.3, Russia).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The work was performed under the theme “new approaches and algorithms in computer modeling of the structure, physicochemical properties, and complex chemical reactions of organic and organoelement compounds” (FMRS-2022-0078, Institute of Petrochemistry and Catalysis UFRC RAS).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Segura, J.L.; Martin, N. [60] Fullerene dimers. Chem. Soc. Rev. 2000, 29, 13–25. [Google Scholar] [CrossRef]
  2. Komatsu, K.; Fujiwara, K.; Tanaka, T.; Murata, Y. The fullerene dimer C120 and related carbon allotropes. Carbon 2000, 38, 1529–1534. [Google Scholar] [CrossRef]
  3. Tukhbatullina, A.A.; Shepelevich, I.S.; Sabirov, D.S. Exaltation of polarizability as a common property of fullerene dimers with diverse intercage bridges. Fuller. Nanotub. Carbon Nanostruct. 2018, 26, 661–666. [Google Scholar] [CrossRef]
  4. Komatsu, K.; Wang, G.-W.; Murata, Y.; Tanaka, T.; Fujiwara, K.; Yamamoto, K.; Saunders, M. Mechanochemical synthesis and characterization of the fullerene dimer C120. J. Org. Chem. 1998, 63, 9358–9366. [Google Scholar] [CrossRef]
  5. Sabirov, D.S.; Terentyev, A.O.; Bulgakov, R.G. Polarizability of fullerene [2+2]-dimers: A DFT study. Phys. Chem. Chem. Phys. 2014, 16, 14594–14600. [Google Scholar] [CrossRef]
  6. Kunitake, M.; Uemura, S.; Ito, O.; Fujiwara, K.; Murata, Y.; Komatsu, K. Structural analysis of C60 trimers by direct observation with scanning tunneling microscopy. Angew. Chem. Int. Ed. 2002, 41, 969–972. [Google Scholar] [CrossRef]
  7. Sabirov, D.S. Polarizability of C60 fullerene dimer and oligomers: The unexpected enhancement and its use for rational design of fullerene-based nanostructures with adjustable properties. RSC Adv. 2013, 3, 19430–19439. [Google Scholar] [CrossRef]
  8. Komatsu, K.; Fujiwara, K.; Murata, Y. The mechanochemical synthesis and properties of the fullerene trimer C180. Chem. Lett. 2000, 29, 1016–1017. [Google Scholar] [CrossRef]
  9. Sabirov, D.S.; Ori, O.; Tukhbatullina, A.A.; Shepelevich, I.S. Covalently bonded fullerene nano-aggregates (C60)n: Digitalizing their energy–topology–symmetry. Symmetry 2021, 13, 1899. [Google Scholar] [CrossRef]
  10. Goedde, B.; Waiblinger, M.; Jakes, P.; Weiden, N.; Dinse, K.-P.; Weidinger, A. ‘Nitrogen doped’ C60 dimers (N@C60–C60). Chem. Phys. Lett. 2001, 334, 12–17. [Google Scholar] [CrossRef]
  11. Plant, S.R.; Jervic, M.; Morton, J.J.L.; Ardavan, A.; Khlobystov, A.N.; Briggs, G.A.D.; Porfyrakis, K. A two-step approach to the synthesis of N@C60 fullerene dimers for molecular qubits. Chem. Sci. 2013, 4, 2971–2975. [Google Scholar] [CrossRef]
  12. Ma, F.; Li, Z.-R.; Zhou, Z.-J.; Wu, D.; Li, Y.; Wang, Y.-F.; Li, Z.-S. Modulated nonlinear optical responses and charge transfer transition in endohedral fullerene dimers Na@C60C60@F with n-fold covalent bond (n = 1, 2, 5, and 6) and long range ion bond. J. Phys. Chem. C 2010, 114, 11242–11247. [Google Scholar] [CrossRef]
  13. Ohtsuki, T.; Masumoto, K.; Tanaka, T.; Komatsu, K. Formation of dimer, trimer, and tetramer of C60 and C70 by γ-ray, charged-particle irradiation, and their HPLC separation. Chem. Phys. Lett. 1999, 300, 661–666. [Google Scholar] [CrossRef]
  14. Kurokawa, S.; Yamamoto, D.; Hirashige, K.; Sakai, A. Possible formation of one-dimensional chains of C20 fullerenes observed by scanning tunneling microscopy. Appl. Phys. Express 2016, 9, 045102. [Google Scholar] [CrossRef]
  15. Zhechkov, L.; Heine, T.; Seifert, G. D5h C50 fullerene:  A building block for oligomers and solids? J. Phys. Chem. A 2004, 108, 11733–11739. [Google Scholar] [CrossRef]
  16. Bonin, K.D.; Kresin, V.V. Electric-Dipole Polarizabilities of Atoms, Molecules, and Clusters; World Scientific: Singapore; River Edge, NJ, USA, 1997. [Google Scholar]
  17. Swart, M.; van Duijnen, P.T. Rapid determination of polarizability exaltation in fullerene-based nanostructures. J. Mater. Chem. 2015, 3, 23–25. [Google Scholar] [CrossRef]
  18. Jackson, K.; Yang, M.; Jellinek, J. Site-specific analysis of dielectric properties of finite systems. J. Phys. Chem. C 2007, 111, 17952–17960. [Google Scholar] [CrossRef]
  19. Marenich, A.V.; Cramer, C.J.; Truhlar, D.G. Reduced and quenched polarizabilities of interior atoms in molecules. Chem. Sci. 2013, 4, 2349–2356. [Google Scholar] [CrossRef]
  20. Tukhbatullina, A.A.; Khamitov, E.M.; Sabirov, D.S. Distributed polarizability of fullerene [2+1]-adducts C60X (n = 1 and 2) with symmetric addends X = CH2 and O: A fresh view on the effect of positional isomerism. Comput. Theor. Chem. 2019, 1149, 31–36. [Google Scholar] [CrossRef]
  21. Frisch, M.J.; Trucks, G.; Schlegel, H.B.; Scuseria, G.E. Gaussian 03, Revision D3.01; Gaussian Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  22. Hirshfeld, F.L. Bonded-atom fragments for describing molecular charge densities. Theor. Chim. Acta 1977, 44, 129–138. [Google Scholar] [CrossRef]
  23. Sabirov, D.S.; Terentyev, A.O.; Cataldo, F. Bisadducts of the C60 and C70 fullerenes with anthracene: Isomerism and DFT estimation of stability and polarizability. Comput. Theor. Chem. 2016, 1081, 44–48. [Google Scholar] [CrossRef]
  24. Sun, D.; Reed, C.A. Crystal eng ineering a linear polymer of C60 fullerene via supramolecular pre-organization. Chem. Commun. 2000, 2391. [Google Scholar] [CrossRef] [Green Version]
  25. Antoine, R.; Dugourd, P.; Rayane, D.; Benichou, E.; Broyer, M.; Chandezon, F.; Guet, C. Direct measurement of the electric polarizability of isolated C60 molecules. J. Chem. Phys. 1999, 110, 9771–9772. [Google Scholar] [CrossRef]
  26. Sohmen, E.; Fink, J.; Krätschmer, W. Electron energy-loss spectroscopy studies on C60 and C70 fullerite. Z. Phys. B Condens. Matter. 1992, 86, 87–92. [Google Scholar] [CrossRef]
  27. Ballard, A.; Bonin, K.; Louderback, J. Absolute measurement of the optical polarizability of C60. J. Chem. Phys. 2000, 113, 5732–5735. [Google Scholar] [CrossRef] [Green Version]
  28. Hackermüller, L.; Hornberger, K.; Gerlich, S.; Gring, M.; Ulbricht, H.; Arndt, M. Optical polarizabilities of large molecules measured in near-field interferometry. Appl. Phys. 2007, 89, 469–473. [Google Scholar] [CrossRef] [Green Version]
  29. Berninger, M.; Stefanov, A.; Deachapunya, S.; Arndt, M. Polarizability measurements of a molecule via a near-field matter-wave interferometer. Phys. Rev. A 2007, 76, 013607. [Google Scholar] [CrossRef] [Green Version]
  30. Hornberger, K.; Gerlich, S.; Ulbricht, H.; Hackermüller, L.; Nimmrichter, S.; Goldt, I.V.; Boltalina, O.; Arndt, M. Theory and experimental verification of Kapitza–Dirac–Talbot–Lau interferometry. New J. Phys. 2009, 11, 043032. [Google Scholar] [CrossRef] [Green Version]
  31. Munn, R.W.; Petelenz, P. Mechanism for effective polarizability enhancement in molecular crystals: C60. Chem. Phys. Lett. 2004, 392, 7–10. [Google Scholar] [CrossRef]
  32. Munn, R.; Malagoli, M.; Panhuis, M.I.H. Environmental effects on molecular response in materials for non-linear optics. Synth. Met. 2000, 109, 29–32. [Google Scholar] [CrossRef]
  33. Pankratyev, E.Y.; Tukhbatullina, A.A.; Sabirov, D.S. Dipole polarizability, structure, and stability of [2+2]-linked fullerene nanostructures (C60)n (n ≤ 7). Phys. E 2017, 86, 237–242. [Google Scholar] [CrossRef]
  34. Broyer, M.; Antoine, R.; Benichou, E.; Compagnon, I.; Dugourd, P.; Rayane, D. Structure of nano-objects through polarizability and dipole measurements. Comptes Rendus Phys. 2002, 3, 301–317. [Google Scholar] [CrossRef]
  35. Sabirov, D.S.; Ōsawa, E. Dipole polarizability of nanodiamonds and related structures. Diamond Relat. Mater. 2015, 55, 64–69. [Google Scholar] [CrossRef]
  36. Ueno, H.; Ōsawa, S.; Ōsawa, E.; Takeuchi, K. Stone-Wales rearrangement pathways from the hinge-opened [2 + 2] C60 dimer to Ipr C120 fullerenes. Vibrational analysis of intermediates. Fuller. Sci. Technol. 1998, 6, 319–338. [Google Scholar] [CrossRef]
  37. Takashima, A.; Nishii, T.; Onoe, J. Formation process and electron-beam incident energy dependence of one-dimensional uneven peanut-shaped C60 polymer studied using in situ high-resolution infrared spectroscopy and density-functional calculations. J. Phys. D Appl. Phys. 2012, 45, 485302. [Google Scholar] [CrossRef]
  38. Wang, G.; Li, Y.; Huang, Y. Structures and electronic properties of peanut-shaped dimers and carbon nanotubes. J. Phys. Chem. 2005, 109, 10957–10961. [Google Scholar] [CrossRef]
  39. Noda, Y.; Ono, S.; Ohno, K. Geometry dependence of electronic and energetic properties of one-dimensional peanut-shaped fullerene polymers. J. Phys. Chem. A 2015, 119, 3048–3055. [Google Scholar] [CrossRef]
  40. Musil, F.; Grisafi, A.; Bartók, A.P.; Ortner, C.; Csányi, G.; Ceriotti, M. Physics-inspired structural representations for molecules and materials. Chem. Rev. 2021, 121, 9759–9815. [Google Scholar] [CrossRef]
  41. Huber, S.E.; Gatchell, M.; Zettergren, H.; Mauracher, A. A precedent of van der Waals interactions outmatching Coulomb explosion. Carbon 2016, 109, 843–850. [Google Scholar] [CrossRef] [Green Version]
  42. Mezour, M.A.; Voznyy, O.; Sargent, E.H.; Lennox, R.B.; Perepichka, D.F. Controlling C60 Organization through Dipole-Induced Band Alignment at Self-Assembled Monolayer Interfaces. Chem. Mater. 2016, 28, 8322–8329. [Google Scholar] [CrossRef]
  43. Hong, I.-H.; Gao, C.-J. Large area self-ordered parallel C60 molecular nanowire arrays on Si(110) surfaces. Carbon 2016, 107, 925–932. [Google Scholar] [CrossRef]
  44. López-Andarias, J.; Bauzá, A.; Sakai, N.; Frontera, A.; Matile, S. Remote control of anion-π catalysis on fullerene-centered catalytic triads. Angew. Chem. 2018, 130, 11049–11053. [Google Scholar] [CrossRef] [Green Version]
  45. Bornhof, A.; Vázquez-Nakagawa, M.; Rodríguez-Pérez, L.; Angeles Herranz, M.; Sakai, N.; Martín, N.; Matile, S.; López-Andarias, J. Anion−π catalysis on carbon nanotubes. Angew. Chem. Int. Ed. 2019, 58, 16097–16100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Shirai, Y.; Osgood, A.J.; Zhao, Y.; Kelly, K.F.; Tour, J.M. Directional control in thermally driven single-molecule nanocars. Nano Lett. 2005, 5, 2330–2334. [Google Scholar] [CrossRef]
  47. Akimov, A.V.; Kolomeisky, A.B. Unidirectional rolling motion of nanocars induced by electric field. J. Phys. Chem. 2012, 116, 22595–22601. [Google Scholar] [CrossRef]
  48. Dorner-Reisel, A.; Ritter, U.; Moje, J.; Freiberger, E.; Scharff, P. Effect of fullerene C60 thermal and tribomechanical loading on Raman signals. Diamond Relat. Mater. 2022, 126, 109036. [Google Scholar] [CrossRef]
  49. Liu, D.; Lungerich, D.; Nakamuro, T.; Harano, K.; Nakamura, E. Excited state modulation of C70 dimerization in a carbon nanotube under a variable electron acceleration voltage. Micron 2022, 160, 103316. [Google Scholar] [CrossRef] [PubMed]
  50. Weippert, J.; Hohmann, L.; Strelnikov, D.; Weis, P.; Pop, M.L.; Böttcher, A.; Kappes, M.M. Desorption of fullerene dimers upon heating non-IPR fullerene films on HOPG. J. Phys. Chem. 2019, 123, 5721–5730. [Google Scholar] [CrossRef]
  51. BinSabt, M.H.; Al-Matar, H.M.; Balch, A.L.; Shalaby, M.A. Synthesis and electrochemistry of novel dumbbell-shaped bis-pyrazolino [60] fullerene derivatives formed using microwave radiation. ACS Omega 2021, 6, 20321–20330. [Google Scholar] [CrossRef]
Figure 1. The formula of a [2+2] fullerene dimer (C60)2. The 6.6 bonds, which are able to be connected to the other C60 core, are shown with arrows. Their designation corresponds to the conventional nomenclature of the C60 bisadducts [23].
Figure 1. The formula of a [2+2] fullerene dimer (C60)2. The 6.6 bonds, which are able to be connected to the other C60 core, are shown with arrows. Their designation corresponds to the conventional nomenclature of the C60 bisadducts [23].
Nanomaterials 12 04404 g001
Figure 2. Dependence of the polarizability exaltation on the distance between the marginal fullerene cores in (C60)3. The data associated with the plot could be found in Table 1.
Figure 2. Dependence of the polarizability exaltation on the distance between the marginal fullerene cores in (C60)3. The data associated with the plot could be found in Table 1.
Nanomaterials 12 04404 g002
Figure 3. The contributions from central and marginal fullerene cores (αcentr and αmarg, respectively) to the mean polarizabilities of the (C60)3 nanostructures as a function of the marginal cores’ remoteness. The point corresponding to the cyclic fullerene trimer is separately selected as all its fullerene units are equivalent. The data associated with the plot can be found in Table 1.
Figure 3. The contributions from central and marginal fullerene cores (αcentr and αmarg, respectively) to the mean polarizabilities of the (C60)3 nanostructures as a function of the marginal cores’ remoteness. The point corresponding to the cyclic fullerene trimer is separately selected as all its fullerene units are equivalent. The data associated with the plot can be found in Table 1.
Nanomaterials 12 04404 g003
Figure 4. The distributed polarizability maps of the (C60)3 trimers: (a) cyclic, (b) e, (c) trans-4, (d) trans-3, (e) trans-2, and (f) trans-1. Color code: blue—negative atomic contributions; other colors—positive atomic contributions: beige (0%–4%), green (5%–68%), pink (69%–96%), and red (96%–100%) correspond to the percentage of the contribution relative to the maximal contribution to the dipole polarizability of the nanostructure.
Figure 4. The distributed polarizability maps of the (C60)3 trimers: (a) cyclic, (b) e, (c) trans-4, (d) trans-3, (e) trans-2, and (f) trans-1. Color code: blue—negative atomic contributions; other colors—positive atomic contributions: beige (0%–4%), green (5%–68%), pink (69%–96%), and red (96%–100%) correspond to the percentage of the contribution relative to the maximal contribution to the dipole polarizability of the nanostructure.
Nanomaterials 12 04404 g004aNanomaterials 12 04404 g004b
Figure 5. The distributed polarizability maps of the (C60)2 dimer (a) and its isomers, tubular fullerene (b), and nanopeanut (c). The color code is the same with Figure 4.
Figure 5. The distributed polarizability maps of the (C60)2 dimer (a) and its isomers, tubular fullerene (b), and nanopeanut (c). The color code is the same with Figure 4.
Nanomaterials 12 04404 g005aNanomaterials 12 04404 g005b
Table 1. The TPSS/TZVP-calculated polarizability and structural parameters of (C60)2 and (C60)3.
Table 1. The TPSS/TZVP-calculated polarizability and structural parameters of (C60)2 and (C60)3.
MoleculeMean Polarizability, α (Å3)Polarizability Exaltation, Δα (Å3) 1Central Core Contribution (Å3)Marginal Core Contribution (Å3)Remoteness of Marginal Cores, L (Å) 2
(C60)2174.5413.94n/a87.279.710
Cyclic (C60)3272.9032.0090.97 390.97 39.655
e-(C60)3267.8326.9383.2692.2812.833
trans-4-(C60)3270.7729.8778.1396.3214.714
trans-3-(C60)3273.4932.6075.5298.9915.804
trans-2-(C60)3276.3835.4870.56102.9117.319
trans-1-(C60)3277.2936.3967.22105.0318.210
1 Calculated with Equation (6), taking the TPSS/TZVP-calculated mean polarizability of the C60 fullerene equal to 80.3 Å3. 2 Remoteness means the distance between the mass centers of the fullerene cores. Note that all distances between the fullerene cages are the same in cyclic (C60)3. 3 No marginal and central cores could be selected in cyclic (C60)3 due to its closed structure. All fullerene cores contribute equally to the molecular property.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sabirov, D.S.; Tukhbatullina, A.A. Distributed Polarizability Model for Covalently Bonded Fullerene Nanoaggregates: Origins of Polarizability Exaltation. Nanomaterials 2022, 12, 4404. https://doi.org/10.3390/nano12244404

AMA Style

Sabirov DS, Tukhbatullina AA. Distributed Polarizability Model for Covalently Bonded Fullerene Nanoaggregates: Origins of Polarizability Exaltation. Nanomaterials. 2022; 12(24):4404. https://doi.org/10.3390/nano12244404

Chicago/Turabian Style

Sabirov, Denis Sh., and Alina A. Tukhbatullina. 2022. "Distributed Polarizability Model for Covalently Bonded Fullerene Nanoaggregates: Origins of Polarizability Exaltation" Nanomaterials 12, no. 24: 4404. https://doi.org/10.3390/nano12244404

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop