ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img

Biased Bowl-Direction of Monofluorosumanene in the Solid State

  • Yumi Yakiyama*
    Yumi Yakiyama
    Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    Innovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    *[email protected]
    More by Yumi Yakiyama
  • Minghong Li
    Minghong Li
    Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    More by Minghong Li
  • Dongyi Zhou
    Dongyi Zhou
    Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    More by Dongyi Zhou
  • Tsuyoshi Abe
    Tsuyoshi Abe
    Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    More by Tsuyoshi Abe
  • Chisato Sato
    Chisato Sato
    Graduate School of Engineering, Tohoku University, 6-6 Aramakiazaaoba, Aoba-ku, Sendai 980-8579, Japan
    More by Chisato Sato
  • Kohei Sambe
    Kohei Sambe
    Graduate School of Engineering, Tohoku University, 6-6 Aramakiazaaoba, Aoba-ku, Sendai 980-8579, Japan
    More by Kohei Sambe
  • Tomoyuki Akutagawa
    Tomoyuki Akutagawa
    Graduate School of Engineering, Tohoku University, 6-6 Aramakiazaaoba, Aoba-ku, Sendai 980-8579, Japan
    Institute of Multidisciplinary Research for Advanced Materials (IMRAM), Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan
  • Teppei Matsumura
    Teppei Matsumura
    Division of Chemical Engineering, Graduate School of Engineering Science, Osaka University, 1-3 Machikaneyama, Toyonaka, Osaka 560-8531, Japan
  • Nobuyuki Matubayasi
    Nobuyuki Matubayasi
    Division of Chemical Engineering, Graduate School of Engineering Science, Osaka University, 1-3 Machikaneyama, Toyonaka, Osaka 560-8531, Japan
  • , and 
  • Hidehiro Sakurai
    Hidehiro Sakurai
    Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    Innovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
Cite this: J. Am. Chem. Soc. 2024, 146, 8, 5224–5231
Publication Date (Web):February 19, 2024
https://doi.org/10.1021/jacs.3c11311

Copyright © 2024 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0.
  • Open Access

Article Views

2457

Altmetric

-

Citations

-
LEARN ABOUT THESE METRICS
PDF (4 MB)
Supporting Info (1)»

Abstract

A new curved π-conjugated molecule 1-fluorosumanene (1) was designed and synthesized that possesses one fluorine atom on the benzylic carbon of sumanene. This compound can exhibit bowl inversion in solution, leading to the formation of two diastereomers, 1endo and 1exo, with different dipole moments. Experimental and theoretical investigation revealed an energetical relationship among 1exo, 1endo, and solvent to realize the various endo:exo ratios in the single crystals of 1 depending on the crystallization solvent. Significantly, the molecular dynamics (MD) simulations revealed that 1exo positively worked for the elongation of the stacking structure and the final endo:exo ratio was affected by the relative stability difference between 1endo and 1exo derived by solvation. Such an arrangeable endo:exo ratio of 1 realized the preparation of unique materials showing a different dielectric response from the same molecule 1 just by changing the crystallization solvent.

This publication is licensed under

CC-BY-NC-ND 4.0.
  • cc licence
  • by licence
  • nc licence
  • nd licence

Introduction

ARTICLE SECTIONS
Jump To

The process of nucleation, the initial step in crystallization, is largely affected by various factors such as temperature, solvent, concentration of the solute, and impurities. (1) These factors significantly affect the molecular arrangements and morphology of the resulting crystals. Understanding these structural aspects is crucial as they are closely linked to the physical properties of crystalline materials. Notably, the contribution of solvents to the nucleation has been investigated recently by combining the experimental approach using 1H NMR and FT-IR spectroscopies and the theoretical one applying density functional theory (DFT) calculations and molecular dynamics (MD) simulations. The outcomes of these studies have unveiled discernible solvent-induced effects on nucleation events in various small aromatic compounds (2−10) and pharmaceutical molecules. (11−14) Despite these advancements, the general rules governing nucleation and subsequent crystallization processes remain elusive. (15−17) Therefore, the control of crystal packings and the physical properties of the crystals remain a challenging issue.
Buckybowls are the partial structures of fullerenes such as C60 and are recognized as distinctive molecular components in the realm of new molecular materials. Bowl-to-bowl inversion is one of the most representative and attractive phenomena in buckybowl chemistry. This unique property of the curved-π system affords a variety of structural and physical properties that respond to external environments. Consequently, these buckybowls hold potential applications as the molecular switches responsive to external stimuli. (18) One representative buckybowl, sumanene (Sum), also shows dynamic phenomenon involved by its bowl inversion in the solution state (Figure 1). (19,20) Sum possesses three benzylic carbons, and their two geminal protons are not geometrically equivalent due to the bowl structure. This characteristic implies that monobenzyl substitution on sumanene potentially gives endo- (concave side) and/or exo- (convex side) substituted derivatives. The resulting endo:exo product ratio is affected by the electronic factor, namely, the stereoelectronic effect, as well as the structural factor, mainly the steric hindrance between the substituent and the bowl. (21) For example, OH group substitution exclusively yields the endo product, while trimethylsilyl substitution gives only the exo product. The above example shows noticeable energy differences between the corresponding diastereomers. Meanwhile, the reaction with comparable exo and endo populations is also reported if the two diastereomers possess similar energy. In such an occasion, it is expected that other environmental factors such as the temperature, concentration, solvent, and the presence of other additives may affect the final endo:exo ratio. Especially in the crystalline state, in which small intermolecular interactions afford different types of stabilization effects from that in the solution state, it is unsurprising to observe a disparate endo:exo ratio.

Figure 1

Figure 1. Bowl inversion of Sum and the conceptual figure of this work.

Recently, our focus has been on studying fluorine-introduced buckybowls, (22−24) especially fluorosumanenes (25−28) from both structural and application viewpoints. The salient feature of these compounds is that they form isostructural single crystals of pristine sumanene. (29,30) Particularly intriguing is the instance of difluorosumanene (F2-Sum), where two fluorine atoms are on a single benzylic carbon site. This configuration induces the anisotropic dielectric response due to its strong dipole moment caused by F-introduction. (26) This attribute has been harnessed to create solid solutions with pristine sumanene and change the activation energy of dielectric response depending on the Sum:F2-Sum ratio. (27) In this context, we further extended our interest to another fluorosumanene family, monofluorosumanene (1), which possesses one fluorine atom at the peripheral benzylic carbon. In this paper, we found that 1 was a diastereomeric mixture of its endo- (1endo) and exo- (1exo) substituted products in solution and solid states. Because of the curved structure of the sumanene skeleton, 1endo and 1exo exhibited distinct electronic structures, leading to different dipole moments (1.96 D for 1endo and 4.57 D for 1exo at B3LYP/6-311+G(d,p)). Noteworthy is that 1 gave the isostructural single crystal packing of Sum, and its endo:exo ratio significantly changed depending on the crystallization solvent from 1endo:1exo = 60:40 to 18:82. This alternation in the endo:exo ratio directly corresponds to a remarkable difference in the dielectric response of each crystal. In addition, we investigated its mechanism through quantum chemistry calculations and molecular dynamics simulations to find that solvation and intermolecular interactions in the aggregated structure significantly affected the final endo:exo ratio in the crystals.

Results and Discussion

ARTICLE SECTIONS
Jump To

Monofluorosumanene (1) was synthesized from direct fluorination of previously reported hydroxysumanene (21) by 135 mol % diethylaminosulfur trifluoride (DAST) in CH2Cl2 in 43% yield (Scheme 1). The fundamental properties of 1 were confirmed by UV–vis spectroscopy and cyclic voltammetry (CV) experiments (Figures S1 and S2). The UV–vis spectrum of 1 showed similar absorption patterns of pristine Sum and other fluorosumanenes, showing an absorption at 276 nm with a broad peak at around 300 nm (Figure S1). These values are in between those of F2-Sum and F6-Sum. (25,26) The voltammogram of 1 in MeCN with tetrabutylammonium perchlorate as a supporting electrolyte showed a clear reduction peak (Epred = −1.81 V (vs Fc0/Fc+), Figure S2) which positively shifted compared with that of Sum (Epred = −2.49 V (vs Fc0/Fc+)). (31)

Scheme 1

Scheme 1. Synthesis of 1-Fluorosumanene (1)
The thermodynamic relationship between 1endo and 1exo was clarified by theoretical calculation and 1H NMR measuremens. The theoretical analysis operated at the B3LYP/6-311+G(d,p) level of theory indicated that 1endo is 0.67 kcal/mol more stable than 1exo and the bowl inversion energy of 1endo was 20.74 kcal/mol while 1exo was 20.06 kcal/mol (Figure 2). These calculation results agreed with the variable temperature 1H NMR measurements in various solvents (Table 1). Regardless of the solvents and the temperature range, the gem-proton signal on the F-attached benzylic carbon of 1endo showed an intensity higher than that of 1exo, appearing at a higher magnetic field (see Supporting Information). This observation strongly indicated that 1endo was the major component of the system. The van’t Hoff plot based on the experimentally obtained endo:exo ratio at each temperature demonstrated no temperature range (193–373 K), in which the equilibrium relationship between 1endo and 1exo could be reversed (Tables 1 and S1, Figure S3). The thermodynamic parameters ΔrH and ΔrS displayed a notable compensatory correlation between enthalpy and entropy, suggesting a consistent inversion mechanism across all solvents (Figure S4). (32,33) It was hypothesized that the positive entropy change, resulting from the flipping from a more crowded environment of 1endo due to the presence of solvent-interactive adjacent H atoms to a more F atom-exposed 1exo structure, serves as the primary driving force behind the bowl inversion process. Although the largest ΔrS value was obtained with acetone, the molecular size (van der Waals volume: CH2Cl2, 56.27 Å3; acetone, 66.60 Å3; DMF, 77.59 Å3) (34) or polarity (ET(30): CH2Cl2, 40.7 kcal/mol; acetone, 42.2 kcal/mol; DMF, 43.2 kcal/mol) (35) did not explain the results, indicating simple consideration of solvent size and polarity does not solve this relationship, and more detailed and careful investigation of the whole bowl flipping process, including transition state through DFT calculation using cluster model of 1 and solvent molecule, is required, (36) which will be our future work.

Figure 2

Figure 2. Structural and energetic difference between 1endo and 1exo. All the calculation data were obtained at the B3LYP/6-311+G(d,p) level of theory.

Table 1. endo:exo Ratio of 1 in CD2Cl2, Acetone-d6, and DMF-d7 at Various Temperaturesa
  1endo:1exo
temperature (K) CD2Cl2 acetone-d6 DMF-d7
373     59:41
348     61:39
323     63:37
298 68:32 64:36 64:36
273 72:28 70:30 66:34
243 75:25 78:22 69:31
193 80:20 90:10  
Thermodynamic Parameter
ΔrH (kJ mol–1) 2.7 7.4 2.5
ΔrS (J mol–1 K–1) 2.5 20 3.3
a

All the samples were prepared in 1.5 mg/mL. Thermodynamic parameters ΔrH and ΔrS obtained via van’t Hoff plot are also shown.

The previous experiments confirmed that no conversion of the major species occurred in the diluted solution state. Nevertheless, considering the slight energetic difference between 1endo and 1exo is small (<1 kcal/mol), it is anticipated that the endo-favored condition is easily replaced by the exo-favored condition when perturbed by intermolecular interactions in a dense system, namely in the solid state. In this context, we next investigated how the endo:exo ratio changes in the solid state by crystallization conditions, such as solvent type and temperature. Single crystal preparation was performed by slow evaporation or vapor diffusion methods using three kinds of solvents, CH2Cl2, acetone, and DMF, with a temperature range from 193 to 323 K (Table 2). All the crystals were isostructural with Sum and already reported other fluorosumanenes, (25,26,29,30) trigonal R3c space group. In all the trials, the temperature effect was not significant, even though the entropic effect seemed to contribute to result in a slightly higher 1endo ratio at more than 298 K. Meanwhile, the endo:exo ratio of 1 in each crystal significantly differed depending on the type of crystallization solvent. Figure 3 shows the typical X-ray analysis results of the single crystals prepared from CH2Cl2 at 193 K. In the structure, we could find no solvent incorporation and observed that only one-third of the sumanene skeleton was independent. Within that, the two-electron densities dangling on the benzylic carbon were not equivalent, indicating the disordering of F atoms in a whole crystal. After refinement of the data, it was found that the occupancy factor of the two F atoms at the endo and the exo sides were 0.154(3) and 0.179(3), respectively; namely, the approximate 1endo:1exo ratio in the present crystal was 46:54 (Figure 3a). The bowl depth, defined by the vertical distance between the bottom hexagonal ring and the peripheral aromatic carbon, was 1.14 Å. This value was in between those of Sum (1.11 Å) (29) and F2-Sum (1.16 Å) (Figure 3b). (26)
Table 2. 1endo:1exo Ratio in a Single Crystal of 1 Prepared in CH2Cl2, Acetone, and DMF at Various Temperatures
  1endo:1exo
temperature (K) CH2Cl2 acetone DMF
323   34:66 22:78
298 45:55 25:75 18:82
243 60:40 27:73 29:71
193 46:54 36:64  

Figure 3

Figure 3. Representative crystal structure of 1. The sample crystal was obtained by the slow evaporation method using CH2Cl2 at 193 K.

In the crystal structure, 1endo and 1exo formed 1D columnar structure mainly stabilized by intracolumnar CH···π interactions between the benzylic C–H and π-plane of the sumanene skeleton and weak π–π interactions to give the bowl to bowl distance to be 3.890 Å (Figure 3b). All of the 1D-columns were connected through CH···π interactions (C–C: 3.895 Å) between benzylic carbon and neighboring aromatic carbons, probably with a slight contribution of CF···π and F···F interactions (2.393 Å) to stabilize the total packing structure with the distance of the central column axes to be 9.625 Å (Figure 3b,c). Notably, the endo:exo ratio difference did not affect the structural parameters much, supporting that the structural difference between 1endo and 1exo was well relaxed in the packing structures (Table S2). It is reasonable to assume that the slight structural and electronic difference between 1endo and 1exo was magnified in the crystal growing process via intermolecular interactions with solvents to achieve the formation of single crystals with such a variety of endo:exo ratios.
Such a biased endo:exo ratio in the crystal of 1, coupled with the significant difference in dipole moments between 1exo and 1endo, prompted us to apply 1 as a modifiable source of dielectric material. We tried to evaluate the dielectric response on the pelletized crystalline (PC) samples made from two kinds of crushed single crystals: recrystallized from DMF (PCDMF) or CH2Cl2 (PCCH2Cl2) (1endo:1exo = 27:73 for PCDMF, 47:53 for PCCH2Cl2). Both samples underwent measurements at several frequencies with the temperature range from 300 to 420 K (Figure 4) under N2 atmosphere. There was a significant difference in the dielectric properties depending on the endo:exo ratio. In the case of PCDMF, both real (ε1) and imaginary (ε2) parts of the dielectric constant were enhanced above ∼360 K at 1 MHz, with a Debye-type dielectric relaxation (Figure 4a,b). This phenomenon was quite similar to that of already reported F2-Sum, (26) indicating that the in-plane directional thermal vibration of polarized 1exo worked as the main contributor to the dielectric response. The ln(τ)–Tp1–1 plots of PCDMF indicated a linear correlation (Figure S5), where Tp1 is the dielectric ε1-peak and the τ-value is the relaxation time of the inverse of the measured f-values of τ = 1/(2πf). The activation energy for the thermally activated motion of PCDMF was evaluated to be Ea = 70 kJ/mol, comparable to previously reported F2-Sum and other reported π-planar organic dielectrics (Figure S5). (37,38) On the contrary, PCCH2Cl2 showed significantly high dielectric constants as a simple organic molecule even after carefully drying up the possibly containing solvents. After the first heating, the dielectric constant decreased significantly (Figure 4c–f). Based on the experimental results, considerable structural relaxation was expected to occur at high temperatures, leading to a more stabilized structure in which the thermally activated motion of 1exo is strongly prohibited.

Figure 4

Figure 4. Temperature dependence of (a, c, e) the real part (ε1) and (b, d, f) the imaginary part (ε2) of the dielectric constant of PCDMF (a, b) and PCCH2Cl2 (c–f) in compressed pellets measured at various frequencies. (e) and (f) represent the 2nd sweep results.

Regarding utilizing the functional molecules, it is a significant benefit that just a change in crystallization solvent gives a substantial property difference. To obtain more profound insight into how each conformer and solvent contribute to the obtained results, we first conducted DFT calculations to investigate how the intermolecular interactions involved in the crystallization process of 1 affect the final 1endo:1exo ratio in the single crystal of 1. Our analysis employed a simple dimer model of 1endo and 1exo to explore the most stable conformer and stacking order. We focused on evaluating the interaction energies between the upper and bottom sides of 1s in the model. Here, interdimer interactions were not considered as intracolumnar interactions were approximately 10 times stronger than intercolumnar ones, especially for sumanene derivatives that pack in the R3c space group. (27) The dimer model preparation was prepared by optimizing 1endo and 1exo at the ωb97XD/6-311+G(d,p) level of theory, with the bowl-to-bowl distance to be 3.9 Å and a rotation angle θ of F-attached benzylic carbons set to 60° or 180° (Figure 5). It should be noted that all the possible stacking orders (upper or lower sides in the dimer) were considered. The simulation performed at the ωb97XD/6-311+G(d,p) level of theory clearly showed that the two dimers (1endo/1endo and 1exo/1endo), in which 1endo was placed at the bottom part of the stacking structure, were less stable (−18.35 to −18.94 kcal/mol) compared to the others (1exo/1exo and 1endo/1exo) with 1exo at the bottom position (−21.03 to −21.34 kcal/mol). These results were reasonable because the F atom of 1endo connected vertically to the bowl, leading to significant steric hindrance when another 1 came from the concave side. In contrast, the C–F bond on 1exo takes a somewhat in-plane direction of the bowl to avoid serious steric repulsion during stacking formation. These results explained that 1exo was more likely to be incorporated into the stacking column, especially when it approached the convex side. Indeed, the increment of 1exo ratio in all the crystals compared with the solution state data suggested that the above discussion is somewhat applicable to explaining the phenomena. However, the significant differences in the final endo:exo ratio in the crystalline state depending on the crystallization solvent indicated the presence of distinct solvent effects during the crystal growth, which assisted or interfered with introducing 1exo into the resulting crystal structure.

Figure 5

Figure 5. Dimer model preparation and the resulting intermolecular energies based on the stacking species and order.

To illustrate the relationship between the solvent and the endo:exo ratio in the crystallization process, we applied molecular dynamics (MD) simulation considering three kinds of solvent systems: acetone, CH2Cl2, and DMF. All the molecules were modeled with a potential function based on the general Amber force field (GAFF) through DFT calculations at the B3LYP/6-31G(d,p) level for geometry optimization. (39,40) All the MD simulations were carried out with GROMACS 2020.6 in the NPT ensemble at 300 K, 1 bar for 30 ns (see Supporting Information for the simulation details). (41) We first investigated the effect of the solvent on the aggregation of 1endo and 1exo. Three systems (100% population of each conformer or a 50:50 mixture of endo and exo conformers) were considered in the simulation. Note that there is no conversion between the endo and exo forms during the simulations. The concentration of 1 was 0.5 M at each run, and the population of the aggregates was examined. Here, we set the position of the center of mass on each conformer and defined the specific component as an “aggregate” when all the molecules are connected with a distance between centers of mass of 8 Å or less (see Supporting Information). The monomer population is larger in the order of DMF > CH2Cl2 > acetone for each conformer and the 50:50 mixture (Figure 6). Conversely, the aggregation tendency is stronger with acetone > CH2Cl2 > DMF and is particularly suppressed in DMF. It is also seen in Figure 6 that the 1endo isomers are less likely to aggregate, indicating that the aggregation is more preferable for the 1exo form.

Figure 6

Figure 6. Aggregation state of 1 expressed as the % population of the number of molecules forming an aggregate with the degree of aggregation in the abscissae for (a) 100% 1exo, (b) 100% 1endo, and (c) 50:50 mixture of them in acetone, CH2Cl2, and DMF. The summed values over the degrees of aggregation of 4 or more are also shown in (d) to illustrate the dependence on the conformational states and solvents more clearly.

The relative preference of the 1exo and 1endo isomers in the monomeric and aggregates states are listed in Table 3. In acetone, for example, the monomeric 1endo and 1exo population is 57:43, implying the preference for the endo form when 1 is monomeric. This agrees with the experimental observation in Table 1, while Table 3 shows that more than half of 1 is in the exo form when the degree of aggregation exceeds 2. The tendency in CH2Cl2 was relatively close to the case in acetone, while in DMF, the 1exo population increment was not as steep as observed in acetone and CH2Cl2, and the average number was 3.3 even at the seventh degree of aggregation.
Table 3. Average Number of 1exo in Each Degree of Aggregation in Acetone, CH2Cl2, and DMFa
  degree of aggregation
solvent 1 2 3 4 5 6 7
acetone 0.43 1.0 1.6 2.2 3.8 3.4 3.8
CH2Cl2 0.44 1.0 1.7 2.3 3.0 3.5 3.9
DMF 0.48 1.0 1.6 2.2 2.6 2.9 3.3
a

The concentration of 1 in the system was 0.5 M, and the endo:exo ratio was 50:50.

We next analyzed the stacking manner of 1 in all of the aggregates, focusing on the neighboring pairs. A pair of 1 is counted as neighbors to each other when their center-of-mass distance is less than 8 Å. The MD simulation in acetone and CH2Cl2 revealed that the population of the 1exo/1exo combination became dominant as the degree of aggregation increased, while others with the endo form were on the decrease (Figure 7). This result matched the conclusion obtained from the DFT calculation that 1exo positively contributed to the elongation of the stacks more than 1endo. The same tendency was observed in the case of DMF, although the population of 1exo/1exo is smaller than those in the other two cases, in agreement with the smaller number for 1exo in Table 3.

Figure 7

Figure 7. Population analysis data for neighboring pairs of 1exo/1exo, 1exo/1endo, 1endo/1exo, and 1endo/1endo in various aggregates in (a) acetone, (b) CH2Cl2, and (c) DMF. The averaged data over the degrees of aggregation of 4 or more are shown in (d) to illustrate the dependence on the solvents more clearly. The definition of the neighboring pairs follows Figure 5.

The above discussion indicated that 1exo was more easily introduced in the aggregates than 1endo, contributing to its elongation, and that aggregation is somehow suppressed in DMF. It should be noted, though, that the endo and exo conformers ratio is 50:50 in their mixture systems in MD. As shown in Table 1, the experimental endo:exo ratio is not 50:50, and this also affects the extent of exo propensity at recrystallization, as discussed next.
The energetic difference in Figure 1 was computed in a vacuum, and the endo:exo ratio is modified due to the solvation effect. This effect is quantified by the solvation free energy Δμ of a monomeric 1, which is shown in Figures 8. (42,43) It is evident that the exo form interacts more favorably with each solvent than does endo and partially cancels the endo preference in Figure 1. The extent of partial cancellation is stronger in the order of DMF > acetone > CH2Cl2, which means that the exo population is larger in this order when the recrystallization process starts. As discussed in Figure 7, on the other hand, the tendency for exo inclusion is more evident in acetone and CH2Cl2. With the combination of the two effects, accordingly, the exo population in the crystal is high with solvent DMF in Table 2 due to enhanced stability of the exo form at the outset of recrystallization, and it is high with acetone due to the higher inclusion efficiencies of 1exo. Indeed, Table 3 and Figures 7 and 8 show for the solvent effects that the exo form is less favorably incorporated into aggregates when it is more strongly solvated. The energetic preference toward exo pairs in aggregates competes against that for the exo monomer at solvation, and the latter effect is more important than the former to determine the exo propensity at recrystallization. Accordingly, a guideline can be proposed that the crystal becomes richer with exo when the exo–solvent interactions are made stronger.

Figure 8

Figure 8. Solvation free energy Δμ of 1endo and 1exo in acetone, CH2Cl2, and DMF. ΔΔμ = Δμ(1endo) – Δμ(1exo) is noted in units of kcal/mol for each solvent.

Conclusion

ARTICLE SECTIONS
Jump To

We successfully synthesized monofluorosumanene 1, which is a diastereomeric mixture of 1exo and 1endo due to bowl inversion. Our experimental and theoretical investigations uncovered the energetic relationship among 1exo, 1endo and solvent, enabling us to arrange the endo:exo ratio in the single crystals of 1 by the appropriate choice of the crystallization solvent. Throughout our project, it was found that the solvation made 1exo more stable, though it was not sufficient to reverse the intrinsic stability of 1endo, resulting in a higher population of 1endo in the solution state. However, the sterically favorable structure of 1exo for aggregation facilitated the inversion of the endo:exo ratio in the solid state. Our DFT calculations and MD simulations played a pivotal role in understanding the underlying mechanisms, highlighting the critical impact of the solvation stabilization effect and the inclusion efficiency of 1exo. Such an arrangeable endo:exo ratio of 1 realized the preparation of unique materials that exhibit entirely different dielectric responses. One showed the Debye type dielectric relaxation, while the other displayed a significantly high dielectric constant, all derived from the same molecule 1, solely by changing the crystallization solvent. This significant achievement opens up new applications for the bowl inversion phenomenon in the curved-π molecules and highlights the remarkable potential of fluorosumanenes as energy materials.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c11311.

  • Synthesis of 1, 1H NMR, 13C NMR, and 19F NMR charts of 1, detailed experimental procedures, single crystal X-ray analysis details, molecular dynamics simulation detail, computational procedures with Cartesian coordinates, and supplemental figures and tables about optical and electrical properties of 1, equilibrium details of 1 in each solvent, structural parameters of single crystal of 1, and Arrhenius plot obtained from PCDMF (PDF)

Accession Codes

CCDC 22889992289010 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Yumi Yakiyama - Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, JapanInnovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, JapanOrcidhttps://orcid.org/0000-0003-4278-2015 Email: [email protected]
  • Authors
    • Minghong Li - Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    • Dongyi Zhou - Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    • Tsuyoshi Abe - Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, Japan
    • Chisato Sato - Graduate School of Engineering, Tohoku University, 6-6 Aramakiazaaoba, Aoba-ku, Sendai 980-8579, Japan
    • Kohei Sambe - Graduate School of Engineering, Tohoku University, 6-6 Aramakiazaaoba, Aoba-ku, Sendai 980-8579, Japan
    • Tomoyuki Akutagawa - Graduate School of Engineering, Tohoku University, 6-6 Aramakiazaaoba, Aoba-ku, Sendai 980-8579, JapanInstitute of Multidisciplinary Research for Advanced Materials (IMRAM), Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, JapanOrcidhttps://orcid.org/0000-0003-3040-1078
    • Teppei Matsumura - Division of Chemical Engineering, Graduate School of Engineering Science, Osaka University, 1-3 Machikaneyama, Toyonaka, Osaka 560-8531, Japan
    • Nobuyuki Matubayasi - Division of Chemical Engineering, Graduate School of Engineering Science, Osaka University, 1-3 Machikaneyama, Toyonaka, Osaka 560-8531, JapanOrcidhttps://orcid.org/0000-0001-7176-441X
    • Hidehiro Sakurai - Division of Applied Chemistry, Graduate School of Engineering, Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, JapanInnovative Catalysis Science Division, Institute for Open and Transdisciplinary Research Initiatives (ICS-OTRI), Osaka University, 2-1 Yamadaoka, Suita, Osaka 565-0871, JapanOrcidhttps://orcid.org/0000-0001-5783-4151
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

This work was supported by Grant-in-Aid for Scientific Research on Innovative Areas “π Space Figuration” (Grant JP26102002), Grant-in-Aid for Transformative Research Areas “Science of 2.5 Dimensional Materials” (Grants JP21H05232 and JP21H05233), “Hyper-Ordered Structures Sciences” (Grants JP21H05563 and JP23H04112), JSPS KAKENHI (Grants JP19H00912 and JP20H00400), and Dynamic Alliance for Open Innovation Bridging Human, Environment and Materials (Grants 20224023 and 20234025). The X-ray diffraction studies of one of the single crystal of 1 by synchrotron radiation were performed at BL02B1 of SPring-8 (Grant 2022A1330). The theoretical calculations were performed at the Research Centre for Computational Science, Okazaki, Japan (Grant 22-IMS-C068). We thank Prof. Gaku Fukuhara for the fruitful discussions about the solvation phenomena of 1. Y.Y. is grateful for support from the program “Initiative for Realizing Diversity in the Research Environment”, Osaka University. M.L. thanks the China Scholarship Council for support (Grants CSC 201708310115 and CSC 201908050056). N.M. is grateful to the Fugaku Supercomputer Project (Grants JPMXP1020230325 and JPMXP1020230327) and the Data-Driven Material Research Project (Grant JPMXP1122714694) from the Ministry of Education, Culture, Sports, Science, and Technology.

References

ARTICLE SECTIONS
Jump To

This article references 43 other publications.

  1. 1
    Davey, R. J.; Schroeder, S. L. M.; ter Horst, J. H. Nucleation of Otganic Crystals-A molecular Perspective. Angew. Chem., Int. Ed. 2013, 52, 21662179,  DOI: 10.1002/anie.201204824
  2. 2
    Gavezzotti, A.; Filippini, G.; Kroon, J.; van Eijck, B. P.; Klewinghaus, P. The Crystal Polymorphism of Tetolic Acid (CH3C═CCOOH): A Molecular Dynamics Study of Precurosrs in Solution, and a Crystal Structure Generation. Chem.─Eur. J. 1997, 3, 893899,  DOI: 10.1002/chem.19970030610
  3. 3
    Davey, R. J.; Blagden, N.; Righini, S.; Alison, H.; Quayle, M. J.; Fuller, S. Crystal Polymorphism as a Probe for Molecular Self-Assembly during Nucleation from Solutions: The Case of 2,6-Dihidroxybenzoic Acid. Cryst. Growth Des. 2001, 1, 5965,  DOI: 10.1021/cg000009c
  4. 4
    Spitaleri, A.; Hunter, C. A.; McCabe, J. F.; Packer, M. J.; Cockroft, S. L. A 1H NMR study of crystal nucleation in solution. CrystEngComm 2004, 6, 489493,  DOI: 10.1039/b407163h
  5. 5
    Parveen, S.; Davey, R. J.; Dent, G.; Pritchard, R. G. Linking solution chemistry to crystal nucleation: the case of tetrolic acid. Chem.Commun. 2005, 15311533,  DOI: 10.1039/b418603f
  6. 6
    Davey, R. J.; Dent, G.; Mughal, R. K.; Parveen, S. Concerning the Relationship between Structural and Growth Synthons in Crystal Nucleation: Solution and Crystal Chemistry of Calboxylic Acids As Revealed thorough IR Spectroscopy. Cryst. Growth Des. 2006, 6, 17881796,  DOI: 10.1021/cg060058a
  7. 7
    Chen, J.; Trout, B. L. Computational Study of Solvent Effects on the Molecular Self-Assembly of Tetrolic Acid in Solution and Implications for the Polymorph Formed from Crystallization. J. Phys. Chem. B 2008, 112, 77947802,  DOI: 10.1021/jp7106582
  8. 8
    Habgood, M. Solution and nanoscale structure selection: implications for the crystal energy landscape of tetrolic acid. Phys. Chem. Chem. Phys. 2012, 14, 91959203,  DOI: 10.1039/c2cp40644f
  9. 9
    Sullivan, R. A.; Davey, R. J.; Sadiq, G.; Dent, G.; Back, K. R.; ter Horst, J. H.; Toroz, D.; Hammond, R. B. Revealing the Roles of Desoluvation and Molecular Self-Assembly in Crystal Nucleation from Solution: Benzoic and p-Aminobenzoic Acids. Cryst. Growth Des. 2014, 14, 26892696,  DOI: 10.1021/cg500441g
  10. 10
    Gaines, E.; Maisuria, K.; Di Tommaso, D. The role of solvent in the self-assembly of m-aminobenzoic acid: a density functional theory and molecular dynamics study. CrystEngComm 2016, 18, 29372948,  DOI: 10.1039/C6CE00130K
  11. 11
    Hunter, C. A.; McCabe, J. F.; Spitaleri, A. Solvent effects of rrthe structures of prenucleation aggregates of carbamazepine. CrystEngComm 2012, 14, 71157117,  DOI: 10.1039/c2ce25941a
  12. 12
    Zeglinski, J.; Kuhs, M.; Khamar, D.; Hegarty, A. C.; Devi, R. K.; Rasmuson, Å. C. Crystal Nucleation of Tolbutamide in Solution: Relationship to Solvent, Solute Conformation, and Solution Structure. Chem.─Eur. J. 2018, 24, 49164926,  DOI: 10.1002/chem.201705954
  13. 13
    Chai, Y.; Wang, L.; Bao, Y.; Teng, R.; Liu, Y.; Xie, C. Investigating the Solvent Effect on Crystal Nucleation of Etoricoxib. Cryst. Growth Des. 2019, 19, 16601667,  DOI: 10.1021/acs.cgd.8b01571
  14. 14
    Simões, R. G.; Melo, P. L. T.; Bernardes, C. E.; Heilmann, M. T.; Emmerling, F.; Minas da Piedade, M. E. Linking Aggregation in Solution, Solvation, and Solubility of Simvastatin: An Experimental and MD Simulation Study. Cryst. Growth Des. 2021, 21, 544551,  DOI: 10.1021/acs.cgd.0c01325
  15. 15
    Burton, R. C.; Ferrari, E. S.; Davey, R. J.; Finney, J. L.; Bowron, D. T. The Relationship between Solution Structure and Crystal Nucleation: A Neutron Scattering Study of Supersaturated Methanolic Solutions of Benzoic Acid. J. Phys. Chem. B 2010, 114, 88078816,  DOI: 10.1021/jp103099j
  16. 16
    Joseph, A.; Rodrigues Alves, J. S.; Bernardes, C. E. S.; Piedade, M. F. M.; Minas da Piedade, M. E. Tautomer selection through solvate formation: the case of 5-hydroxynicotinic acid. CrystEngComm 2019, 21, 22202233,  DOI: 10.1039/C8CE02108B
  17. 17
    Nakamuro, T.; Sakakibara, M.; Nada, H.; Harano, K.; Nakamura, E. Capturing the Moment of Emergence of Crystal Nucleus from Disorder. J. Am. Chem. Soc. 2021, 143, 17631767,  DOI: 10.1021/jacs.0c12100
  18. 18
    Fujii, S.; Ziatdinov, M.; Higashibayashi, S.; Sakurai, H.; Kiguchi, M. Bowl Inversion and Electric Switching of Buckybowls on Gold. J. Am. Chem. Soc. 2016, 138, 1214212149,  DOI: 10.1021/jacs.6b04741
  19. 19
    Sakurai, H.; Daiko, T.; Hirao, T. A Synthesis of Sumanene, a Fullerene Fragment. Science 2003, 301, 1878,  DOI: 10.1126/science.1088290
  20. 20
    Amaya, T.; Sakane, H.; Muneishi, T.; Hirao, T. Bowl-to-bowl inversion of sumanene derivatives. Chem. Commun. 2008, 765767,  DOI: 10.1039/B712839H
  21. 21
    Higashibayashi, S.; Onogi, S.; Srivastava, H. K.; Sastry, G. N.; Wu, Y.-T.; Sakurai, H. Stereoelectronic Effect of Courved Aromatic Structures: Favoring the Unexpxcted endo Confomration of Benzylic-Substituted Sumanene. Angew. Chem., Int. Ed. 2013, 52, 73147316,  DOI: 10.1002/anie.201303134
  22. 22
    Kuvychko, I. V.; Dubceac, C.; Deng, S. H. M.; Wang, X.; Granovsky, A. A.; Popov, A. A.; Petrukhina, M. A.; Strauss, S. H.; Boltalina, O. V. C20H4(C4F8)3: A Fluorine-Containing Annulated Corannulene that Is a Better Electron Acceptor Than C60. Angew. Chem., Int. Ed. 2013, 52, 75057508,  DOI: 10.1002/anie.201300796
  23. 23
    Dubceac, C.; Sevryugina, Y.; Kuvychko, I. V.; Boltalina, O. V.; Strauss, S. H.; Petrukhina, M. A. Self-Assembly of Aligned Hybrid One-Dimensional Stacks from Two Complementary π-Bowls. Cryst. Growth Des. 2018, 18, 307311,  DOI: 10.1021/acs.cgd.7b01258
  24. 24
    Haupt, A.; Lentz, D. Corranulenes with Electron-Withdrawing Substituents: Synthetic Approaches and Resulting Structural and Electronic Properties. Chem.─Eur. J. 2019, 25, 34403454,  DOI: 10.1002/chem.201803927
  25. 25
    Schmidt, B. M.; Topolinski, B.; Higashibayashi, S.; Kojima, T.; Kawano, M.; Lentz, D.; Sakurai, H. The Synthesis of Hexafluorosumanene and Its Congeners. Chem.─Eur. J. 2013, 19, 32823286,  DOI: 10.1002/chem.201204622
  26. 26
    Li, M.; Wu, J.; Sambe, K.; Yakiyama, Y.; Akutagawa, T.; Kajitani, T.; Fukushima, T.; Matsuda, K.; Sakurai, H. Dielectric response of 1,1-difluorosumanene caused by an in-plane motion. Mater. Chem. Front. 2022, 6, 17521758,  DOI: 10.1039/D2QM00134A
  27. 27
    Li, M.; Chen, X.; Yakiyama, Y.; Wu, J.; Akutagawa, T.; Sakurai, H. Tuning the dielectric response by co-crystallization of sumanene and its fluorinated derivative. Chem. Commun. 2022, 58, 89508953,  DOI: 10.1039/D2CC02766F
  28. 28
    Yakiyama, Y.; Li, M.; Sakurai, H. Fluorosumanenes as building blocks for organic crystalline dielectrics. Pure Appl. Chem. 2023, 95, 421430,  DOI: 10.1515/pac-2023-0211
  29. 29
    Sakurai, H.; Daiko, T.; Sakane, H.; Amaya, T.; Hirao, T. Structural Elucidation os Sumanene and Generation of Its Benzylic Anions. J. Am. Chem. Soc. 2005, 127, 1158011581,  DOI: 10.1021/ja0518169
  30. 30
    Mebs, S.; Weber, M.; Luger, P. A.; Schmidt, B. M.; Sakurai, H.; Higashibayashi, S.; Onogi, S.; Lentz, D. Experimental electron density of sumanene, a bowl-shaped fullerene fragment; comparison with the related corannulene hydrocarbon. Org. Biomol. Chem. 2012, 10, 22182222,  DOI: 10.1039/c2ob07040e
  31. 31
    Zanello, P.; Fedi, S.; de Biani, F. F.; Giorgi, G.; Amaya, T.; Sakane, H.; Hirao, T. The electrochemical inspection of the redox activity of sumanene and its concave CpFe complex. Dalton Trans. 2009, 91929197,  DOI: 10.1039/b910711h
  32. 32
    Liu, L.; Guo, Q. Isokinetic Relationship, Isoequilibrium Relationship, and Enthalpy-Entropy Compensation. Chem. Rev. 2001, 101, 673696,  DOI: 10.1021/cr990416z
  33. 33
    Pan, A.; Biswas, T.; Rakshit, A. K.; Moulik, S. P. Enthalpy-Entropy Compensation (EEC) Effect: A Recisit. J. Phys. Chem. B 2015, 119, 1587615884,  DOI: 10.1021/acs.jpcb.5b09925
  34. 34
    Zhao, Y. H.; Abraham, M. H.; Zissimos, A. M. Fast Calculation of van der Waals Volume as a Sum of Atomic and Bond Contributions and Its Application to Drug Compounds. J. Org. Chem. 2003, 68, 73687373,  DOI: 10.1021/jo034808o
  35. 35
    Reichardt, C. Solvents and Solvent Effects in Organic Chemistry; Wiley-VCH: Weinheim, Germany, 2003.
  36. 36
    Bruno, C.; Benassi, R.; Passalacqua, A.; Paolucci, F.; Fontanesi, C.; Marcaccio, M.; Jackson, E. A.; Scott, L. T. Electrochamical and Theoretical Investigation of Corannulene Reduction Processes. J. Phys. Chem. B 2009, 113, 19541962,  DOI: 10.1021/jp8045092
  37. 37
    Harada, J.; Yoneyama, N.; Sato, S.; Takahashi, Y.; Inabe, T. Crystals of Charge-Transfer Complexes with Reorienting Polar Molecules: Dielectric Properties and Order-Disorder Phase Transitions. Cryst. Growth Des. 2019, 19, 291299,  DOI: 10.1021/acs.cgd.8b01418
  38. 38
    Harada, J.; Ohtani, M.; Takahashi, Y.; Inabe, T. Molecular Motion, Dielectric Response, and Phase Transition of Charge-Transfer Crystals: Acquired Dynamic and Dielectric Properties of Polar Molecules in Crystals. J. Am. Chem. Soc. 2015, 137, 44774486,  DOI: 10.1021/jacs.5b00412
  39. 39
    Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 11571174,  DOI: 10.1002/jcc.20035
  40. 40
    Bayly, C. I.; Cieplak, P.; Cornell, W. D.; Kollman, P. A. A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model. J. Phys. Chem. 1993, 97, 1026910280,  DOI: 10.1021/j100142a004
  41. 41
    Abraham, M. J.; Murtola, T.; Schulz, R.; Ṕall, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–2, 1925,  DOI: 10.1016/j.softx.2015.06.001
  42. 42
    Matubayasi, N.; Nakahara, M. Theory of Solutions in the Energy Representation. II. Functional for the Chemical Potential. J. Chem. Phys. 2002, 117, 36053616,  DOI: 10.1063/1.1495850
  43. 43
    Sakuraba, S.; Matubayasi, N. ERmod: Fast and Versatile Computation Software for Solvation Free Energy with Approximate Theory of Solutions. J. Comput. Chem. 2014, 35, 15921608,  DOI: 10.1002/jcc.23651

Cited By

ARTICLE SECTIONS
Jump To

This article has not yet been cited by other publications.

  • Abstract

    Figure 1

    Figure 1. Bowl inversion of Sum and the conceptual figure of this work.

    Scheme 1

    Scheme 1. Synthesis of 1-Fluorosumanene (1)

    Figure 2

    Figure 2. Structural and energetic difference between 1endo and 1exo. All the calculation data were obtained at the B3LYP/6-311+G(d,p) level of theory.

    Figure 3

    Figure 3. Representative crystal structure of 1. The sample crystal was obtained by the slow evaporation method using CH2Cl2 at 193 K.

    Figure 4

    Figure 4. Temperature dependence of (a, c, e) the real part (ε1) and (b, d, f) the imaginary part (ε2) of the dielectric constant of PCDMF (a, b) and PCCH2Cl2 (c–f) in compressed pellets measured at various frequencies. (e) and (f) represent the 2nd sweep results.

    Figure 5

    Figure 5. Dimer model preparation and the resulting intermolecular energies based on the stacking species and order.

    Figure 6

    Figure 6. Aggregation state of 1 expressed as the % population of the number of molecules forming an aggregate with the degree of aggregation in the abscissae for (a) 100% 1exo, (b) 100% 1endo, and (c) 50:50 mixture of them in acetone, CH2Cl2, and DMF. The summed values over the degrees of aggregation of 4 or more are also shown in (d) to illustrate the dependence on the conformational states and solvents more clearly.

    Figure 7

    Figure 7. Population analysis data for neighboring pairs of 1exo/1exo, 1exo/1endo, 1endo/1exo, and 1endo/1endo in various aggregates in (a) acetone, (b) CH2Cl2, and (c) DMF. The averaged data over the degrees of aggregation of 4 or more are shown in (d) to illustrate the dependence on the solvents more clearly. The definition of the neighboring pairs follows Figure 5.

    Figure 8

    Figure 8. Solvation free energy Δμ of 1endo and 1exo in acetone, CH2Cl2, and DMF. ΔΔμ = Δμ(1endo) – Δμ(1exo) is noted in units of kcal/mol for each solvent.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 43 other publications.

    1. 1
      Davey, R. J.; Schroeder, S. L. M.; ter Horst, J. H. Nucleation of Otganic Crystals-A molecular Perspective. Angew. Chem., Int. Ed. 2013, 52, 21662179,  DOI: 10.1002/anie.201204824
    2. 2
      Gavezzotti, A.; Filippini, G.; Kroon, J.; van Eijck, B. P.; Klewinghaus, P. The Crystal Polymorphism of Tetolic Acid (CH3C═CCOOH): A Molecular Dynamics Study of Precurosrs in Solution, and a Crystal Structure Generation. Chem.─Eur. J. 1997, 3, 893899,  DOI: 10.1002/chem.19970030610
    3. 3
      Davey, R. J.; Blagden, N.; Righini, S.; Alison, H.; Quayle, M. J.; Fuller, S. Crystal Polymorphism as a Probe for Molecular Self-Assembly during Nucleation from Solutions: The Case of 2,6-Dihidroxybenzoic Acid. Cryst. Growth Des. 2001, 1, 5965,  DOI: 10.1021/cg000009c
    4. 4
      Spitaleri, A.; Hunter, C. A.; McCabe, J. F.; Packer, M. J.; Cockroft, S. L. A 1H NMR study of crystal nucleation in solution. CrystEngComm 2004, 6, 489493,  DOI: 10.1039/b407163h
    5. 5
      Parveen, S.; Davey, R. J.; Dent, G.; Pritchard, R. G. Linking solution chemistry to crystal nucleation: the case of tetrolic acid. Chem.Commun. 2005, 15311533,  DOI: 10.1039/b418603f
    6. 6
      Davey, R. J.; Dent, G.; Mughal, R. K.; Parveen, S. Concerning the Relationship between Structural and Growth Synthons in Crystal Nucleation: Solution and Crystal Chemistry of Calboxylic Acids As Revealed thorough IR Spectroscopy. Cryst. Growth Des. 2006, 6, 17881796,  DOI: 10.1021/cg060058a
    7. 7
      Chen, J.; Trout, B. L. Computational Study of Solvent Effects on the Molecular Self-Assembly of Tetrolic Acid in Solution and Implications for the Polymorph Formed from Crystallization. J. Phys. Chem. B 2008, 112, 77947802,  DOI: 10.1021/jp7106582
    8. 8
      Habgood, M. Solution and nanoscale structure selection: implications for the crystal energy landscape of tetrolic acid. Phys. Chem. Chem. Phys. 2012, 14, 91959203,  DOI: 10.1039/c2cp40644f
    9. 9
      Sullivan, R. A.; Davey, R. J.; Sadiq, G.; Dent, G.; Back, K. R.; ter Horst, J. H.; Toroz, D.; Hammond, R. B. Revealing the Roles of Desoluvation and Molecular Self-Assembly in Crystal Nucleation from Solution: Benzoic and p-Aminobenzoic Acids. Cryst. Growth Des. 2014, 14, 26892696,  DOI: 10.1021/cg500441g
    10. 10
      Gaines, E.; Maisuria, K.; Di Tommaso, D. The role of solvent in the self-assembly of m-aminobenzoic acid: a density functional theory and molecular dynamics study. CrystEngComm 2016, 18, 29372948,  DOI: 10.1039/C6CE00130K
    11. 11
      Hunter, C. A.; McCabe, J. F.; Spitaleri, A. Solvent effects of rrthe structures of prenucleation aggregates of carbamazepine. CrystEngComm 2012, 14, 71157117,  DOI: 10.1039/c2ce25941a
    12. 12
      Zeglinski, J.; Kuhs, M.; Khamar, D.; Hegarty, A. C.; Devi, R. K.; Rasmuson, Å. C. Crystal Nucleation of Tolbutamide in Solution: Relationship to Solvent, Solute Conformation, and Solution Structure. Chem.─Eur. J. 2018, 24, 49164926,  DOI: 10.1002/chem.201705954
    13. 13
      Chai, Y.; Wang, L.; Bao, Y.; Teng, R.; Liu, Y.; Xie, C. Investigating the Solvent Effect on Crystal Nucleation of Etoricoxib. Cryst. Growth Des. 2019, 19, 16601667,  DOI: 10.1021/acs.cgd.8b01571
    14. 14
      Simões, R. G.; Melo, P. L. T.; Bernardes, C. E.; Heilmann, M. T.; Emmerling, F.; Minas da Piedade, M. E. Linking Aggregation in Solution, Solvation, and Solubility of Simvastatin: An Experimental and MD Simulation Study. Cryst. Growth Des. 2021, 21, 544551,  DOI: 10.1021/acs.cgd.0c01325
    15. 15
      Burton, R. C.; Ferrari, E. S.; Davey, R. J.; Finney, J. L.; Bowron, D. T. The Relationship between Solution Structure and Crystal Nucleation: A Neutron Scattering Study of Supersaturated Methanolic Solutions of Benzoic Acid. J. Phys. Chem. B 2010, 114, 88078816,  DOI: 10.1021/jp103099j
    16. 16
      Joseph, A.; Rodrigues Alves, J. S.; Bernardes, C. E. S.; Piedade, M. F. M.; Minas da Piedade, M. E. Tautomer selection through solvate formation: the case of 5-hydroxynicotinic acid. CrystEngComm 2019, 21, 22202233,  DOI: 10.1039/C8CE02108B
    17. 17
      Nakamuro, T.; Sakakibara, M.; Nada, H.; Harano, K.; Nakamura, E. Capturing the Moment of Emergence of Crystal Nucleus from Disorder. J. Am. Chem. Soc. 2021, 143, 17631767,  DOI: 10.1021/jacs.0c12100
    18. 18
      Fujii, S.; Ziatdinov, M.; Higashibayashi, S.; Sakurai, H.; Kiguchi, M. Bowl Inversion and Electric Switching of Buckybowls on Gold. J. Am. Chem. Soc. 2016, 138, 1214212149,  DOI: 10.1021/jacs.6b04741
    19. 19
      Sakurai, H.; Daiko, T.; Hirao, T. A Synthesis of Sumanene, a Fullerene Fragment. Science 2003, 301, 1878,  DOI: 10.1126/science.1088290
    20. 20
      Amaya, T.; Sakane, H.; Muneishi, T.; Hirao, T. Bowl-to-bowl inversion of sumanene derivatives. Chem. Commun. 2008, 765767,  DOI: 10.1039/B712839H
    21. 21
      Higashibayashi, S.; Onogi, S.; Srivastava, H. K.; Sastry, G. N.; Wu, Y.-T.; Sakurai, H. Stereoelectronic Effect of Courved Aromatic Structures: Favoring the Unexpxcted endo Confomration of Benzylic-Substituted Sumanene. Angew. Chem., Int. Ed. 2013, 52, 73147316,  DOI: 10.1002/anie.201303134
    22. 22
      Kuvychko, I. V.; Dubceac, C.; Deng, S. H. M.; Wang, X.; Granovsky, A. A.; Popov, A. A.; Petrukhina, M. A.; Strauss, S. H.; Boltalina, O. V. C20H4(C4F8)3: A Fluorine-Containing Annulated Corannulene that Is a Better Electron Acceptor Than C60. Angew. Chem., Int. Ed. 2013, 52, 75057508,  DOI: 10.1002/anie.201300796
    23. 23
      Dubceac, C.; Sevryugina, Y.; Kuvychko, I. V.; Boltalina, O. V.; Strauss, S. H.; Petrukhina, M. A. Self-Assembly of Aligned Hybrid One-Dimensional Stacks from Two Complementary π-Bowls. Cryst. Growth Des. 2018, 18, 307311,  DOI: 10.1021/acs.cgd.7b01258
    24. 24
      Haupt, A.; Lentz, D. Corranulenes with Electron-Withdrawing Substituents: Synthetic Approaches and Resulting Structural and Electronic Properties. Chem.─Eur. J. 2019, 25, 34403454,  DOI: 10.1002/chem.201803927
    25. 25
      Schmidt, B. M.; Topolinski, B.; Higashibayashi, S.; Kojima, T.; Kawano, M.; Lentz, D.; Sakurai, H. The Synthesis of Hexafluorosumanene and Its Congeners. Chem.─Eur. J. 2013, 19, 32823286,  DOI: 10.1002/chem.201204622
    26. 26
      Li, M.; Wu, J.; Sambe, K.; Yakiyama, Y.; Akutagawa, T.; Kajitani, T.; Fukushima, T.; Matsuda, K.; Sakurai, H. Dielectric response of 1,1-difluorosumanene caused by an in-plane motion. Mater. Chem. Front. 2022, 6, 17521758,  DOI: 10.1039/D2QM00134A
    27. 27
      Li, M.; Chen, X.; Yakiyama, Y.; Wu, J.; Akutagawa, T.; Sakurai, H. Tuning the dielectric response by co-crystallization of sumanene and its fluorinated derivative. Chem. Commun. 2022, 58, 89508953,  DOI: 10.1039/D2CC02766F
    28. 28
      Yakiyama, Y.; Li, M.; Sakurai, H. Fluorosumanenes as building blocks for organic crystalline dielectrics. Pure Appl. Chem. 2023, 95, 421430,  DOI: 10.1515/pac-2023-0211
    29. 29
      Sakurai, H.; Daiko, T.; Sakane, H.; Amaya, T.; Hirao, T. Structural Elucidation os Sumanene and Generation of Its Benzylic Anions. J. Am. Chem. Soc. 2005, 127, 1158011581,  DOI: 10.1021/ja0518169
    30. 30
      Mebs, S.; Weber, M.; Luger, P. A.; Schmidt, B. M.; Sakurai, H.; Higashibayashi, S.; Onogi, S.; Lentz, D. Experimental electron density of sumanene, a bowl-shaped fullerene fragment; comparison with the related corannulene hydrocarbon. Org. Biomol. Chem. 2012, 10, 22182222,  DOI: 10.1039/c2ob07040e
    31. 31
      Zanello, P.; Fedi, S.; de Biani, F. F.; Giorgi, G.; Amaya, T.; Sakane, H.; Hirao, T. The electrochemical inspection of the redox activity of sumanene and its concave CpFe complex. Dalton Trans. 2009, 91929197,  DOI: 10.1039/b910711h
    32. 32
      Liu, L.; Guo, Q. Isokinetic Relationship, Isoequilibrium Relationship, and Enthalpy-Entropy Compensation. Chem. Rev. 2001, 101, 673696,  DOI: 10.1021/cr990416z
    33. 33
      Pan, A.; Biswas, T.; Rakshit, A. K.; Moulik, S. P. Enthalpy-Entropy Compensation (EEC) Effect: A Recisit. J. Phys. Chem. B 2015, 119, 1587615884,  DOI: 10.1021/acs.jpcb.5b09925
    34. 34
      Zhao, Y. H.; Abraham, M. H.; Zissimos, A. M. Fast Calculation of van der Waals Volume as a Sum of Atomic and Bond Contributions and Its Application to Drug Compounds. J. Org. Chem. 2003, 68, 73687373,  DOI: 10.1021/jo034808o
    35. 35
      Reichardt, C. Solvents and Solvent Effects in Organic Chemistry; Wiley-VCH: Weinheim, Germany, 2003.
    36. 36
      Bruno, C.; Benassi, R.; Passalacqua, A.; Paolucci, F.; Fontanesi, C.; Marcaccio, M.; Jackson, E. A.; Scott, L. T. Electrochamical and Theoretical Investigation of Corannulene Reduction Processes. J. Phys. Chem. B 2009, 113, 19541962,  DOI: 10.1021/jp8045092
    37. 37
      Harada, J.; Yoneyama, N.; Sato, S.; Takahashi, Y.; Inabe, T. Crystals of Charge-Transfer Complexes with Reorienting Polar Molecules: Dielectric Properties and Order-Disorder Phase Transitions. Cryst. Growth Des. 2019, 19, 291299,  DOI: 10.1021/acs.cgd.8b01418
    38. 38
      Harada, J.; Ohtani, M.; Takahashi, Y.; Inabe, T. Molecular Motion, Dielectric Response, and Phase Transition of Charge-Transfer Crystals: Acquired Dynamic and Dielectric Properties of Polar Molecules in Crystals. J. Am. Chem. Soc. 2015, 137, 44774486,  DOI: 10.1021/jacs.5b00412
    39. 39
      Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. A.; Case, D. A. Development and testing of a general amber force field. J. Comput. Chem. 2004, 25, 11571174,  DOI: 10.1002/jcc.20035
    40. 40
      Bayly, C. I.; Cieplak, P.; Cornell, W. D.; Kollman, P. A. A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model. J. Phys. Chem. 1993, 97, 1026910280,  DOI: 10.1021/j100142a004
    41. 41
      Abraham, M. J.; Murtola, T.; Schulz, R.; Ṕall, S.; Smith, J. C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–2, 1925,  DOI: 10.1016/j.softx.2015.06.001
    42. 42
      Matubayasi, N.; Nakahara, M. Theory of Solutions in the Energy Representation. II. Functional for the Chemical Potential. J. Chem. Phys. 2002, 117, 36053616,  DOI: 10.1063/1.1495850
    43. 43
      Sakuraba, S.; Matubayasi, N. ERmod: Fast and Versatile Computation Software for Solvation Free Energy with Approximate Theory of Solutions. J. Comput. Chem. 2014, 35, 15921608,  DOI: 10.1002/jcc.23651
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.3c11311.

    • Synthesis of 1, 1H NMR, 13C NMR, and 19F NMR charts of 1, detailed experimental procedures, single crystal X-ray analysis details, molecular dynamics simulation detail, computational procedures with Cartesian coordinates, and supplemental figures and tables about optical and electrical properties of 1, equilibrium details of 1 in each solvent, structural parameters of single crystal of 1, and Arrhenius plot obtained from PCDMF (PDF)

    Accession Codes

    CCDC 22889992289010 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect