Li
Wang
*,
Yan-Li
Liu
,
Sheng-Hui
Chen
,
De
He
,
Quan-Jiang
Li
and
Mei-Shan
Wang
*
School of Physics and Optoelectronics Engineering, Ludong University, Yantai, 264025, Shandong, China. E-mail: wangl@ldu.edu.cn; mswang1971@163.com
First published on 25th November 2020
In the formation of noncovalent complexes, the stacking arrangements of corannulene and fullerene are diverse, most of which are combinations of multiple corannulenes and fullerene. Here, a composition ratio of 2:1 was selected for the complex between corannulene and fullerene (C60 and C70) to investigate the effects of different superposition modes, including concave–convex and convex–convex interactions, on the stability and third-order nonlinear optical (NLO) properties of the composite materials. It was found that the concave–convex interaction was stronger and it was reported to stabilize the charge-transfer (CT) complex more effectively than the convex–convex interaction. The dispersion range of the concave–convex interaction was larger than that of the convex–convex interaction, which is consistent with the interaction energy results. The packing design with the double convex–convex interactions exhibited the largest linear optical response and third-order NLO response, which showed that the convex–convex interaction was more likely to be excited and cause intermolecular CT as compared to the concave–convex interaction. This work confirmed that the packing arrangement significantly affected the NLO response and will advance the development of NLO crystals.
Along with the discovery of fullerenes, such as spherical carbon molecules C60 and C70 with convex surfaces, it was determined that shape complementarity effectively promotes the π–π intermolecular interactions between curved–curved surfaces.7 Numerous studies have examined the effective binding of fullerenes and other functional groups through non-covalent interactions and observed how host–guest supramolecular systems with specific structures are formed, and then realized their functionalization by regulating the π–π interactions between various functional groups. The development of novel host structures that are combined with the guest fullerene is very meaningful work in host–guest chemistry of fullerenes.8–11 A series of main compounds has been developed, including two different types of compounds. One type consists of rich π-electronic compounds containing the derivatives of ferrocene,12 porphyrin,13 and the corannulene system.14 Another type consists of large ring compounds including the derivatives of cyclodextrin,15,16 calixarene,17 and oligomers.18
Bowl-shaped corannulene (C20H10) naturally offers satisfactory geometric compatibility with the convex surface of fullerenes, which can enhance intermolecular electrostatic interactions for concave–convex and convex–convex interactions.19 Due to its curvature, this compound possesses large dipole moments that provide enhanced π–π intermolecular interactions.20 It has been shown that π–π interactions between pristine C20H10 and C60 can be promoted by the adsorption of the former on a Cu(110)-surface. The apparent height of C60 with respect to C20H10 is thus expected to be approximately 3.4 Å.19
In our previous study, we calculated the complex composed of C20H10 and fullerenes of different sizes,21 which predicted that the complexation of the concave face of C20H10 toward fullerenes would be energetically similar to the π⋯π interactions between homodimers of corannulene.22 However, multiple corannulenes and fullerene are often combined, and therefore, the stacking modes of corannulenes and fullerene are diverse. In this work, composites of corannulenes and fullerene compounded at a ratio of 2:1 were explored to determine the influence of various stacking modes on the stability and nonlinear optical properties of composites. The investigation on the effect of different arrangements of fullerene-corannulene complexes on their NLO properties not only provides additional data for the further development of organic and polymer optoelectronic materials through adjustable molecular stacking and non-covalent interactions, but it will also facilitate the further development of more interesting properties.
The binding interaction energy (Ebind) of the complex was calculated at the B97D/6-31+G(d) level using the counterpoise (CP) procedure, which can correct the basis set superposition error (BSSE).26,27Ebind can be expressed as:
Ebind = E(AB)AB − E(A)AB − E(B)B | (1) |
In order to reveal the averaged weak interaction regions and distinguish three different types of noncovalent interactions, (i) π–π interaction, (ii) hydrogen-binding, and (iii) steric hindrance, a pair of very important functions, reduced density gradient (RDG) and sin(λ2)*ρ were used:28
(2) |
It is well known that the NLO response is caused by the polarization of the material under the action of a strong electric field. All hyperpolarizabilities in this work are reported in the “T” convention.29
(3) |
In this expression, E0 denotes the molecular energy without the applied electric field, and i, j, and k denote different components along the x, y, and z directions, respectively. Fi denotes the Cartesian component of the applied electric field along the i direction. μi denotes the permanent dipole moment along the i direction; αij, βijk, and γijkl are the polarizability, first, and second hyperpolarizability tensors, respectively.
The isotropic average polarizability (α) and second hyperpolarizability (γ) can be expressed as the differential form of dipole moment with respect to the electric field intensity F.
(4) |
(5) |
The isotropic average polarizability (α) is calculated as:
(6) |
The orientationally averaged second hyperpolarizability (γ) has been calculated using the following expression:29
(7) |
The γ value was calculated by means of the analytic and finite field hybrid method. This is a method to obtain hyperpolarizability that automatically performs one or more finite-difference methods on the basis of low-order analytic derivatives to obtain high-order derivatives. Here, it is based on third order analytical derivatives, and then automatically conducts one higher order finite difference to obtain the required fourth order derivative. Compared with the finite field method, this method is more efficient and decreases computing costs. The DFT method has been widely and effectively used for the calculation of (super) polarizability and is considered as an undeniable alternative to the demanding ab initio method. Herein, functionals CAM-B3LYP and BHandHLYP were employed in these calculations. The CAM-B3LYP functional, which adds a long-range-corrected version of B3LYP using the Coulomb-attenuating method, includes 19% and 65% of the short- and long-range HF exchange.30 The BHandHLYP functional is obtained by including 50% of the exact exchange in the rgw BLYP functional.31
Time-dependent (TD)-DFT has currently become a general tool to enable understanding and prediction of the behavior of the electron transition property. The keyword SCRF = (PCM, solvent = o-dichlorobenzene) was added to the command line of TD calculation to simulate the absorption spectrum under the solvent. The PCM method in Gaussian 09 includes an external iteration procedure whereby the program computes the energy in solution by ensuring that the solvent reaction field is self-consistent with the solute electrostatic potential (the latter being generated from the computed electron density with the specified model chemistry).
Fig. 1 (a) Three five-membered rings in different positions are indicated by a, b, and c. (b) Isomers of the complex (2C20H10/C60) obtained at the B3LYP-D3/6-31G(d) level. |
To determine the stability of the isomers, the single point energy calculation for three types of isomers was studied (Table S1, ESI†). The stabilization energy difference between position a and position b ranges from 0.000012 to 0.021607 atomic units (a.u.), and a lower stabilization energy is associated with position a. Therefore, the isomers of position a for complex 1–5 are more stable. However, such a small difference of 0.000012 a.u. for complex 6 indicated that they are of similar energetic ordering. To further determine the stability of isomers, the energy differences (Egap) between the HOMO and the LUMO were taken into account (Table S1, ESI†). It is well known that the smaller the Egap, the more dynamic and unstable the system. From Table S1 (ESI†), the Egap values of position b are all smaller than that of position a, and therefore, the isomers of position a are more stable. In addition, the structural models are derived from the crystal structure of the experimental results.32 The molecular structure reported in the experiment is the isomer of position a. Therefore, we chose structures of position a for further studies herein.
Based on the discussion above, complexes 1–6 based on 3 possible concave/convex arrangements for each complex were further examined (Fig. 2). The relative stabilization energies of these isomers are also shown, and they reveal that complexes 1 and 4 have the lowest relative energies of −22.33 and −22.17 kcal mol−1. This indicates that the stacking mode of two concave surfaces stacked with the convex surface of fullerene is the most stable isomer. On the contrary, structures 3 and 6 with the stacking mode of two convex surfaces stacked with the convex surface of fullerene are the most active with highest stabilization energy. This is due to the concave–convex charge transfer (CT) interaction that was reported to stabilize the CT complex more effectively than the convex–convex interaction because of the large electrostatic interaction that plays a pivotal role in determining the energetically accessible stacking motifs.22
Geometrical complementarity between the concave/convex aromatic face of C20H10 and convex face of fullerene was established under the distance between the bottom five-membered ring of C20H10 and C60 of 3.22–3.47 Å. The distances between the concave face of C20H10 and convex face of C60 and C70 are 3.47 and 3.46 Å, respectively, which is similar to the previous result of 3.5 Å between C20H10 and C60.21 C20H10 has five six-membered rings around a central five-membered ring, as segments of C60 and C70, thus indicating similar distances. C20H10 is involved in convex–convex interactions with the C60-surface, with the shortest distance being 3.22 Å, which is similar to the reported distance of 3.23 Å for the convex–convex interaction between C28H14 and C60. The bottom of the carbon cage is parallel to the central five-element torus of C20H10, and thus achieves the maximum concave–convex interaction.
The energy levels and molecular orbital diagrams of the HOMO and LUMO are shown in Fig. 3, and for these complexes, the HOMO and LUMO are mainly on fullerene. According to molecular orbital theory, there is no orbital overlap between C20H10 and fullerene, which also reveals that the weak interaction between C60 and C20H10 is formed by concave–convex interactions between C20H10 and fullerenes. The energy differences (Egap) between the HOMO and the LUMO of C60 and C70 obtained by B3LYP/6-31G(d) and complexes 1–6 are embedded in Fig. 2. Compared to pure C60 and C70, complexes 1–6 have smaller Egap values, indicating that the weak interaction causes a slight lowering of the energy gaps of C60 and C70. The Egap value of C60 was evaluated as 2.89 eV along with −6.03 eV of HOMO and −3.14 eV of LUMO. The Egap values of 1–3 were evaluated as 2.68 eV > 2.67 eV > 2.64 eV, respectively, which indicated that the weak interaction causes a slight lowering of the Egap values of C60 ranging from 0.21 to 0.25 eV. Furthermore, it was found that the convex–convex interaction was more conducive for reducing the energy gap as compared to the concave–convex interaction. Specifically, the weak interaction raised the HOMO energy levels by 0.31 eV (1), 0.43 eV (2), and 0.53 eV (3) and raised the LUMO energy levels by 0.1 eV (1), 0.21 eV (2), and 0.28 eV (3). Likewise, the Egap values of 4, 5, and 6 were also evaluated as 2.71 eV > 2.69 eV > 2.67 eV, which were lower than that of pure C70 (2.73 eV) by 0.02 eV, 0.04 eV, and 0.06 eV. It was also found that the convex–convex interaction was more conducive to reducing the energy gap as compared to the concave–convex interaction. Specifically, the weak interaction raised the HOMO energy levels by 0.19 eV (4), 0.29 eV (5), and 0.39 eV (6) and raised the LUMO energy levels by 0.17 eV (4), 0.25 eV (5), and 0.33 eV (6).
Fig. 3 The frontier molecular orbital diagrams and HOMO and LUMO energy levels of the studied complexes obtained at the B3LYP-D3/6-31G(d) level. |
Complex | E up | E down | E int |
---|---|---|---|
1 | −17.26 | −17.26 | −34.52 |
2 | −16.75 | −9.34 | −26.21 |
3 | −9.16 | −9.16 | −18.72 |
4 | −17.36 | −17.36 | −34.76 |
5 | −17.29 | −9.46 | −26.87 |
6 | −9.24 | −9.24 | −18.90 |
Reduced density gradient (RDG) is a fundamental dimensionless quantity in DFT used to describe the deviation from a homogeneous electron distribution. The noncovalent interactions including the concave–convex and convex–convex interaction are visualized by the figures showing electron density and low-gradient isosurfaces (RDG) in this section. RDG indicates the type of interaction, and its strength can be derived from the density on the noncovalent interaction surface because noncovalent interactions are highly nonlocal and manifest in real space as low-gradient isosurfaces with low densities. The low-density, low-gradient spike with low electron density (−0.05 < sin(λ2)ρ < 0.05) in Fig. 4 (black figure) indicates the presence of weak noncovalent interactions, which provide a satisfactory balance between the weakly repulsive and attractive interactions, resulting in greater thermal stability of the complex.
To distinguish between the three different types of non-covalent interactions, including strong attraction, π–π interaction, and steric hindrance, RDG (s = 0.5 a.u.) vs. sin(λ2)ρ for the complexes was plotted. sin(λ2)ρ is a definition of a real space function, namely, the product of the sine of λ2 and ρ. Different colors were used to represent the values of this function according to the color bar below, and they were mapped to RDG isosurfaces to designate where weak interactions occur, and also to easily capture the interaction type. The noncovalent interaction regions of all complexes are filled with green or light brown, which confirms an overwhelming preponderance of the π–π interaction. In Fig. 4, the dispersion range of the concave–convex interaction was evaluated as compared to that of the convex–convex interaction, and it is consistent with the result of the Eint value. It was found that the area of RDG can represent the interaction energy, which is positively correlated with the dispersion range to a large extent.
To investigate the electronic transition, electron density difference maps (EDDMs) corresponding to the crucial electronic transitions are shown in Fig. 5. The electron density difference can be exactly evaluated as:
Δρ(r) = ρele(r) − ρhole(r) | (8) |
Complex | State | λ | ΔEge | f os | f os 2/ΔEge5 | Excitation (% composition) |
---|---|---|---|---|---|---|
a Assignment: H = HOMO, L = LUMO, H−1 = HOMO−1, L+1 = LUMO+1, etc. | ||||||
1 | S34 | 322.1 | 3.85 | 0.004 | 1.89 × 10−8 | Ha−10 → L+2 (29%), H-9 → L+1 (29%) H-5 → L (11%) |
2 | S32 | 323.7 | 3.83 | 0.336 | 1.37 × 10−4 | H−6 → L + 1 (13%), H−6 → L+2 (18%) H−5 → L+1 (21%), H−5 → L+2 (12%) |
3 | S26 | 333.2 | 3.72 | 0.727 | 7.42 × 10−4 | H−6 → L+1 (39%), H−5 → L (42%) |
4 | S2 | 472.6 | 2.62 | 0.078 | 4.93 × 10−5 | H−1 → L (26%), H → L+1 (54%) |
S30 | 362.8 | 3.42 | 0.218 | 1.02 × 10−4 | H−3 → L (29%), H → L+3 (56%) | |
5 | S2 | 472.8 | 2.62 | 0.086 | 5.99 × 10−5 | H−1 → L+2 (20%), H → L+1 (62%) |
S32 | 358.6 | 3.46 | 0.567 | 6.48 × 10−4 | H−7 → L+1 (22%), H−6 → L (21%), H−2 → L+4 (18%), H−1 → L+5 (18%) | |
6 | S2 | 473.2 | 2.62 | 0.093 | 7.01 × 10−5 | H−1 → L+2 (14%), H → L (70%) |
S29 | 363.1 | 3.41 | 0.6117 | 8.12 × 10−4 | H−3 → L+2 (29%), H−2 → L+4 (12%), H−1 → L+5 (12%), H → L+3 (42%) |
Complex | 〈R2〉 | Functionals | α tot | γ tot |
---|---|---|---|---|
1 | 56198.4 | BHandHLYP | 928.3 | 271774.0 |
CAM-B3LYP | 928.1 | 278630.9 | ||
2 | 60036.2 | BHandHLYP | 959.2 | 511368.5 |
CAM-B3LYP | 959.7 | 559226.1 | ||
3 | 66504.6 | BHandHLYP | 988.4 | 724360.4 |
CAM-B3LYP | 989.5 | 807513.3 | ||
4 | 73926.3 | BHandHLYP | 1056.8 | 374262.2 |
CAM-B3LYP | 1058.6 | 389005.6 | ||
5 | 81643.6 | BHandHLYP | 1089.8 | 737074.0 |
CAM-B3LYP | 1092.8 | 817738.6 | ||
6 | 87560.0 | BHandHLYP | 1122.3 | 1074495.1 |
CAM-B3LYP | 1126.2 | 1216516.9 |
Fig. 6 (a) The αtot values for the complexes. (b) The γtot values for the complexes calculated using the CAM-B3LYP and BHandHLYP methods. |
The total polarizability (αtot) values of the complexes gradually increased as 928.3 a.u. (1) < 959.2 a.u. (2) < 988.4 a.u. (3), which indicated that the packing arrangement with double convex–convex interactions exhibits the largest linear optical response. The same conclusion was obtained from the results of C70 complexes, that is, 1056.8 a.u. (4) < 1089.8 a.u. (5) < 1122.3 a.u. (6). The relative electronic spatial extent 〈R2〉 values are physical properties that represent the electron density volume around the molecule and usually represent the size of the molecule.33 The 〈R2〉 values obtained by the B3LYP/6-31G level in series increase as 56198.4 a.u. (1) < 60036.2 a.u. (2) < 66504.6 a.u. (3) and 73926.3 a.u. (4) < 81643.6 a.u. (5) < 87560.0 a.u. (6). They are in agreement with the increasing order of the αtot values, which is due to the fact that more diffuse electron clouds may lead to larger 〈R2〉 values, resulting in a larger αtot value.
The γtot values were gradually increased from 271774.0 a.u. (1) to 724360.4 a.u. (3) as well as 724360.4 a.u. (4) to 1074495.1 a.u. (6), which show that the packing arrangement with the double convex–convex interactions possesses the largest third order nonlinear optical response. It was found that the convex–convex interaction was more likely to be excited and cause intermolecular CT as compared to the concave–convex interaction. The results show that the larger NLO response is related to the packing arrangement, which is of great significance for the further study of such systems and their derivatives. To understand the origin of the third-order NLO response, a two-level model of the γtot value was considered for the studied complexes, which is the linkage between the γtot value and electronic transition(s) in low–lying crucial excited states. Therefore, using the two-level models is a reliable method to analyze second hyperpolarizabilities, and the expression is as follows: fos2/ΔE5, where fos the oscillator strength, and ΔE the transition energy.34
In the two-level expression, the γtot value is proportional to the quadratic power of fos but inversely proportional to the fifth power of ΔE. The fos2/ΔE5 of low-lying crucial excited states has been calculated as 1.89 × 10−8 (1) < 1.37 × 10−4 (2) < 7.42 × 10−4 (3), which is in quantitative agreement with the regular γtot values. For C70 complexes, the low-lying crucial excited states were chosen as S30 for 4, S32 for 5, and S29 for 6. The fos2/ΔE5 of these excited states was calculated as 1.02 × 10−4 (4) < 6.48 × 10−4 (5) < 8.12 × 10−4 (6), which also agree with the regular γtot values. This indicates that the γtot value depends on the transition energy and oscillator strength of the low excited state. A smaller transition energy and larger oscillator strength are conducive to the larger γtot values. The low-lying crucial excited state of complex 3 and 6 have the largest fos (0.727 for 3 and 2.62 for 6) and smallest ΔE (3.72 eV for 3 and 0.093 eV for 6), and thus, the largest γtot values are for complexes 3 and 6.
Field (10−4 a.u.) | Energy (a.u.) | D z + | D z − | E gap |
---|---|---|---|---|
0 | −3823.0219 | 3.469 | 3.469 | 2.68 |
10 | −3823.0220 | 3.482 | 3.459 | 2.66 |
20 | −3823.0233 | 3.492 | 3.447 | 2.57 |
30 | −3823.0254 | 3.506 | 3.433 | 2.36 |
40 | −3823.0284 | 3.516 | 3.418 | 2.14 |
50 | −3823.0323 | 3.524 | 3.401 | 1.93 |
The polarizabilities (αtot) and second hyperpolarizabilities (γtot) of complex 1 were also calculated on the basis of the resulting structures with the external electric fields including 10 × 10−4, 20 × 10−4, 30 × 10−4, 40 × 10−4, and 50 × 10−4 a.u. at the direction of +z to study the effect of an electric field on NLO properties (Tables S6 and S7, ESI†). The results show that as the intensity of the electric field increases from 0 to 50 × 10−4 a.u., the αtot values range from 928.1 a.u. to 928.6 a.u., which indicates that the intensity of the external electric field has little influence on the polarizability. However, with regard to the second hyperpolarizabilities, the influence of the strength of the external electric field is enormous. It is due to the α value being the first derivative, but the γ value is the third derivative of the electric field according to eqn (4) and (5). Therefore, the α value remains unchanged under the electric field. However, the γ value exhibits non-monotonic behaviour under the electric field. The γtot values of complex 1 with external electric fields are 280082.1 a.u. (F = 10 × 10−4 a.u.), 285384.6 (F = 20 × 10−4 a.u.), 281285.3 (F = 30 × 10−4 a.u.), 282167.2 (F = 40 × 10−4 a.u.), and 283268.6 (F = 50 × 10−4 a.u.), respectively. The variation tendency of the γtot values is shown in Fig. 7. The addition of an external electric field will increase the γtot value by 1451.2–6753.64 a.u., and the γtot value sharply increases when the external electric field intensity is 20 × 10−4 a.u.
The total polarizability (αtot) values of the complexes were gradually increased in the order 1 < 2 < 3 < 4 < 5 < 6, which indicated that the packing arrangement with double convex–convex interactions exhibited the largest linear optical response. The same conclusion can be drawn from the results of the second hyperpolarizabilities (γtot). This shows that the packing arrangement with double convex–convex interactions possesses the largest third-order nonlinear optical response. Thus, the convex–convex interaction is more likely to be excited and cause intermolecular CT rather than the concave–convex interaction. This work offers new methods for searching for hosts for fullerenes via convex–convex shape complementarity and for exploring excellent nonlinear optical materials.
Footnote |
† Electronic supplementary information (ESI) available: The components of the total polarizabilities α (a.u.) and the second hyperpolarizabilities γ (a.u.) for the studied complexes. See DOI: 10.1039/d0cp03253k |
This journal is © the Owner Societies 2021 |