Dong Meng, Guogang Liu, Chengyi Xiao, Yanjun Shi, Lei Zhang, Lang

π-π stacking of the more planar components consisting of down PDI petal and part of the corannulene core (Figure S9a). Charge Transport Property. To...
1 downloads 0 Views 1MB Size
Subscriber access provided by ECU Libraries

Article

Corannurylene Pentapetalae Dong Meng, Guogang Liu, Chengyi Xiao, Yanjun Shi, Lei Zhang, Lang Jiang, Kim K. Baldridge, Yan Li, Jay S Siegel, and Zhaohui Wang J. Am. Chem. Soc., Just Accepted Manuscript • Publication Date (Web): 08 Mar 2019 Downloaded from http://pubs.acs.org on March 8, 2019

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

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

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

Journal of the American Chemical Society

Dong Meng,†,‡ Guogang Liu,‡,§ Chengyi Xiao,┴ Yanjun Shi,‡ Lei Zhang,┴ Lang Jiang,‡ Kim K. Baldridge,# Yan Li,*,‡,§ Jay S. Siegel*,# and Zhaohui Wang*,†,‡ †

Key Laboratory of Organic Optoelectronics and Molecular Engineering, Department of Chemistry, Tsinghua University, Beijing 100084, China ‡

Beijing National Laboratory for Molecular Sciences, CAS Key Laboratory of Organic Solids, CAS Center of Excellence in Molecular Science, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China #

School of Pharmaceutical Science and Technology, Tianjin University, 92 Weijin Road, Nankai District, Tianjin 300072, China ┴

College of Energy, Beijing University of Chemical Technology, Beijing 100029, China

§

School of Chemistry and Chemical Engineering, University of Chinese Academy of Sciences, Beijing 100049, China

ABSTRACT: Despite the great advances in the synthesis of diverse nonplanar graphenoids, investigations into the relationship between structural features and intermolecular interactions still present significant challenges. Herein, the novel nonplanar graphenoid structure, corannurylene pentapetalae (CRP), obtained via bottom-up syntheses of hybridization between perylene diimide (PDI) planar fragments and a corannulene curved core, is presented. Single crystal studies reveal a D5-symmetric as well as a C2symmetric graphenoid corannurylene pentapetalae. The D5-symmetric structure has a unique honeycomb lattice with two chiral honeycomb layers alternately stacked, whereas the C2-symmetric CRP forms dimer units via π-π stacking. Transistor devices demonstrate that, without any π-π stacking, the honeycomb lattice of the D5-symmetric CRP has the potential to also facilitate electron transport.

Graphenoids with planar sheets of varying size, shape, and periphery have made great progress in practical applications of electronics, energy storage, catalysis, and biosensing.1 In recent years, a growing number of graphenoid studies focus on unusual curved and warped structures such as bowls, helices, and saddles on account of their structural variety and unique physical and chemical properties.2,3 Nonplanar graphenoids are typically arrived at using one of two strategies of embeding four-, five-, seven-, or eight-membered rings into the hexagonal lattice,4 and introducing large bulky groups to the periphery.5 Recently, a new area in which curved and flat graphenoids are hybridized has attracted the attention of chemists and material scientists. Such methods are expected to enable construction of novel nonplanar graphenoids with unique structural features, intermolecular interactions, and eventually distinctive applications as functional materials. Perylene diimides (PDIs), which are considered as one of the most promising n-type semiconductors because of their high electron affinities, good chemical and thermal stabilities, and excellent electron-transporting characteristics, are typical planar graphenoids, the single crystals of which are inclined to arrange along the π-stacking direction of the perylene core even with functional groups in their bay regions.6 Corannulene as the smallest subunit of C60 is a typical curved graphenoid with concave and convex π-surfaces.2 Until now, some nonplanar graphenoids containing corannulene subunits have been reported.4(a,b),7 However, investigations into the relationships

between structural features of hybrid nonplanar graphenoids and intermolecular interactions are still challenging, features that hold significance to their potential applications. Herein is discussed the design and synthesis of the novel nonplanar graphenoid, corannurylene pentapetalae (CRP, Figure 1).8 Hybridization of five PDI fragments around a corannulene core followed by cycloaromatization, yields D5- and C2-symmetric corannurylene pentapetalae flowers. Bottom up syntheses that use modular methods to arrive quickly at atomically defined hybrids of different graphenoid fragments yield unexpected structures. This high symmetry chiral propeller manifests a broad spectrum of complex structural and physical behavior thus providing insight to the molecular design and chemical synthesis of tailor-made atomically defined nonplanar nanomolecular graphenoid materials. This advance in non-planar graphenoids will enable the synthesis and study of super molecular propellers and there materials applications.

Synthesis and Stereoisomerism. The synthetic route to corannurylene pentapetalae (CRP-n) is shown in Figure 1a. CRP-n containing 36 aromatic rings and 540 atoms were constructed by two steps. The intermediate compound 3 was synthesized via five-fold Suzuki-Miyaura coupling reaction between monobromo-PDI 19 and 1,3,5,7,9pentakis(pinacolatoboryl) corannulene 210 using PdCl2(dppf) as the catalyst and Na2CO3 as the base in 60% yield. CRP-n

ACS Paragon Plus Environment

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

Page 2 of 8

Figure 1. (a) Synthetic route to corannurylene pentapetalae and (b) eight possible stereoisomers containing four pairs of enantiomers caused by steric hindrance effect during five PDI petals.

were obtained via photocyclic aromatization under irradiation in 62% yield. Considering the large steric hindrance effect during five PDI petals, there are eight possible stereoisomers containing four pairs of enantiomers as shown in Figure 1b. CRP-1 ((P,P,P,P,P)-(M,M,M,M,M) enantiomers) with ten imide groups of five PDI petals arranged up and down has the highest D5 symmetry, whereas CRP-2 ((P,P,P,P,M)(M,M,M,M,P)) with six imide groups of three successive PDI petals arranged up and down and one whole PDI up and one whole PDI down has the reduced C2 symmetry. C2-symmetric CRP-3 ((P,P,P,M,M)-(M,M,M,P,P)) has two PDI petals in meta-position up and down, and CRP-4 ((P,M,P,M,M)(M,P,M,P,P)) has two PDI petals in meta-position up and another two PDI petals in meta-position down. The C2 axes of CRP-2, CRP-3, and CRP-4 are shown in Figure S3. Two pairs of enantiomers, CRP-1 and CRP-2, were obtained with ratio ~1:1 under irradiation condition of 40W LED lamp. CRP-1 and CRP-2 are predicted to be lower in energy (∆G = 0 and 5.53 kcal.mol-1, respectively), than CRP-3 (∆G = 6.03 kcal.mol-1) and CRP-4 (∆G = 11.42 kcal.mol-1) via B3LYP-D3/6-311G(2d,p) calculations (Table S1). The con-

version from CRP-2 to CRP-1 was attempted by heating its diphenyl ether solution at 250 °C for 6 hours. However, no changes were observed from its 1H NMR spectra (Figure S4). All of these compounds were unambiguously characterized by 1 H and 13C NMR spectra, and high-resolution mass spectra. CRP-1 and CRP-2 are soluble in common organic solvents such as dichloromethane, chloroform, tetrahydrofuran, toluene and ortho-dichlorobenzene (o-DCB) at room temperature due to their non-planar molecular scaffold and the branched alkyl chains. The thermal properties were evaluated by thermal gravimetric analysis (TGA) performed under nitrogen. CRP-n possess excellent thermal stability with decomposition temperature of 5% weight loss over 390 °C (Figure S5). In order to investigate the chirality of CRPs with different symmetry, enantiomerically pure CRP-1 and CRP-2 were achieved by using chiral high-performance liquid chromatography (HPLC) (Figure S6). Two enantiomers of CRP-1 and CRP-2 were characterized, respectively, by circular dichroism (CD) spectroscopy in CHCl3 between which there are subtle distinctions in shapes and big differences in the intensity of Δɛ (Figure S7).

ACS Paragon Plus Environment

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

Journal of the American Chemical Society

Figure 2. Structural conformations and crystal packing arrangements of CRPs. (a) Single-crystal X-ray molecular structure of CRP-1 (top view and side view) and the molecular packing arrangement of honeycomb structure. (b) Single-crystal X-ray molecular structure of CRP2 (top view and side view) and corresponding arrangement. The alkyl chains and hydrogen atoms are omitted for clarity.

Molecular Conformation and Crystal Packing Arrangement. The molecular conformations of CRP-1 and CRP-2 were confirmed by X-ray diffraction analysis of single crystals obtained by slow vapor diffusion method (THF/MeOH), which exhibit pentapetalae-shaped conformations with different structural features (Figure 2). In order to clarify CRPs skeletons, branched alkyl chains in imide groups of PDI petals are omitted in Figure 2 and displayed in Figure S8 and S9. The single crystal of CRP-1 was characterized as racemates of (P,P,P,P,P)-(M,M,M,M,M)-enantiomers, and the single crystal of CRP-2 was revealed as racemates of (P,P,P,P,M)(M,M,M,M,P)-enantiomers. The size of CRP-1 skeleton is ~21 Å in length and width and 8.0 Å in altitude. Because of two up and down PDI petals, CRP-2 has small difference in width and a higher altitude of 9.67 Å. The average splay angle (α value in Figure 2a and Figure S10) of CRPs is about 31°for CRP-1, 32°and 42°for CRP-2 (two molecules in a cell). The bowl depth of pristine corannulene is 0.87 Å.11 The central corannulene moiety of CRPs is stretched and adopts a shallow bowl-shaped geometry with bowl depth of ~0.32 Å for CRP-1 and ~0.63 Å for CRP-2 attributing to the outward tensile force from the steric repulsion of five PDI petals. The bowl depth of CRP-1 was also much shallower than that (~0.57 Å) of quintuple [6]helicene with a corannulene core.12 Two up and down PDI units (red and green ones) in CRP-2 relieve the repulsion to adjacent PDI units endowing them with deeper bowl depth than CRP-1. Honeycomb structures in polymers by self-organization have been reported.13 However, honeycomb lattice selfassembled by nonplanar graphenoid structure with chirality has not yet been reported. From Figure 2a it can be seen that

three CRP-1 molecules with the same chirality self-assemble into a tubular structure with ~21.3 Å in height and ~18.4 Å in diameter, which can be considered as the smallest unit to construct honeycomb lattice. Two types of chiral honeycomb layers are found in the crystal lattice, which alternately stack to form a three-dimensional achiral honeycomb lattice. From Figure S8 it can be seen that the branched alkyl chains in imide groups of PDI petals serve to support the tubular units and the lamellar arrangement. In order to investigate the effect of structural features of CRPs backbones to the honeycomb arrangement, the crystal packing arrangement of CRP-1 was compared with that of CRP-2. From Figure 2b and Figure S9 it can be seen that CRP-2 is inclined to form dimer units via π-π stacking of the more planar components consisting of down PDI petal and part of the corannulene core (Figure S9a). Charge Transport Property. To gain insight into the charge transport property of peculiar honeycomb lattice, the micro/nanocrystals of CRP-1 were prepared by in situ dropcoating method (see Supplementary Section 10). As shown in Figure 3a and 3b, highly regular hexagonal crystals possessing smooth surface were obtained. Theoretical simulations using the growth morphology algorithm of Materials Studio (MS) suggests a thermodynamically stable trigonal crystalline structure, as shown in Figure 3d, which coincides well with the observed hexagonal morphology. 1D out-of-plane X-ray diffraction (XRD) (Figure 3c) confirms the formation of single crystalline microsheets with only a triple Bragg diffraction peaks. The first diffraction peak at 2θ = 4.1°was found with the distance of 21.6 Å corresponding to the thickness of one

ACS Paragon Plus Environment

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

Page 4 of 8

Figure 3. Single-crystal transistor characterization of CRP-1. (a) Optical microscopy (OM) image of micro/nanocrystals in large area selfassembled on OTS treated Si/SiO2 substrate in dilute CHCl3 (1 mg/ml) at room temperature. (b) Representative highly regular hexagonal crystal with six edges at ~9 micron. (c) 1D out-of-plane XRD pattern of single crystal data. (d) Predicted morphology based on attachment energies principle, viewed perpendicular to the (002) plane and side view. (e) TEM image and its corresponding SAED pattern. (f) SEM image of a transistor with six electrodes to probe the charge transport properties along different crystal planes.

honeycomb layer formed by (M,M,M,M,M) or (P,P,P,P,P) enantiomers associated with the branched alkyl chains and could be calibrated as (002) lattice plane in accordance with the single crystal diffraction data. A typical transmission electron microscopy (TEM) image of hexagonal microcrystal with ~0.78 micrometers of the edge length and its corresponding selected-area electron diffraction (SAED) pattern are shown in Figure 3e. The lattice constants were calculated to be a = b = 25.4 Å and γ = 120°by analysis of the diffraction spots. The symmetric six in-plane lattice planes calculated from the SAED analysis were also in accordance with the MS calculation. 1D out-of-plane XRD data together with SAED pattern and single-crystal XRD data revealed that the micro/nanometer-sized single crystals possess the same phase with previously obtained single crystals. SEM image of Figure 3f shows the transistor devices based on individual hexagonal microcrystal of CRP-1 with six electrodes and channels fabricated by “two-dimensional organic-ribbon mask” technique14, which is the best device configuration to examine the charge transport properties of the crystal along different crystal planes. The transfer curves of six different crystal planes are shown in Figure S11 and the mobility data are summarized in Table S3. The transistors along the channels 1↔2, 2↔3, 3↔4, 4↔5, 5↔6, 6↔1 all show electron mobility at 10-4 cm2V-1s-1 (μ1-2 = 5.92×10-4, μ2-3 = 5.69×10-4, μ3-4 = 7.32×10-4, μ4-5 = 6.69×10-4, μ5-6 = 9.64×10-4, μ6-1 = 6.92×10-4). This result implies that without any π-π stacking, the honeycomb lattice of CRP-1 could also facilitate electron transport with a very small difference during six crystal planes.15 Optical and Electronic Properties. The UV-vis absorption spectra of CRPs in chloroform (1 × 10-5 M) are shown in Figure 4a, which exhibit broad absorption in the whole 350 to 600 nm region along with much stronger absorbance (εmax = 2.6 × 105 M-1 cm-1 for CRP-1 and 2.3 × 105 M-1 cm-1 for CRP-2) than monomer PDI and known PDI oligomers.3a CRP-n in

Figure 4. UV-vis absorption spectra of CRP-1 and CRP-2 in CHCl3 (1 × 10-5 M) (a) and in films (b), and normalized excitation and emission spectra (480 nm for CRP-1 and 474 nm for CRP-2) at room temperature (RT) and low temperature (LT) (77 K) in methylcyclohexane.

Table 1. Electrochemical Properties and Energy Levels of CRP-1 and CRP-2 in Solution E1ra [V]

E2ra [V]

ELUMOb [eV]

EHOMOc [eV]

Egd [eV]

CRP-1

-1.09

-1.19

-3.86

-6.01

2.15

CRP-2

-1.13

-1.22

-3.83

-5.98

2.15

a

+

Half-wave reductive potentials (in V vs Fc/Fc ) measured in CH2Cl2 at a scan rate of 0.1 V/s with ferrocene as an internal potential mark. b Estimated from the onset potential of the first reduction wave and calculated according to ELUMO = -(4.8 + Eonsetre) eV. c Calculated according to EHOMO = (ELUMO - Eg) eV. d Obtained from the edge of the absorption spectra in CHCl3 according to Eg = (1240/λonset).

ACS Paragon Plus Environment

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

Journal of the American Chemical Society film (Figure 4b) have similar absorption spectra to that in solution, demonstrating that there is no obvious aggregation in film based on CRP backbones. The emission spectra of CRPs in methylcyclohexane at room and low temperature (77 K) are shown in Figure 4c and 4d. The fluorescence quantum yield and lifetime of D5-symmetric CRPs (ΦF = 7.36%, τF = 4.26 ns for CRP-1) were decreased in comparison to C2-symmetric CRPs (ΦF = 11.85%, τF = 6.36 ns for CRP-2). The electrochemical properties of CRPs with different structure features were studied by cyclic voltammetry in CH2Cl2 (Figure S12 and Table 1). Both CRP molecules exhibit eight obvious reversible reduction waves with the first half waves at -1.09 V for CRP-1 and -1.13 V for CRP-2 versus Fc/Fc+ implying strong electron-accepting abilities. The LUMO values of CRP-1 and CRP-2 in CH2Cl2 are estimated to be −3.86 eV and −3.83 eV from the onset reduction potential, and are comparable to several fullerene derivatives recently used as acceptors in organic solar cells.

fragments around a corannulene core. Two types of chiral honeycomb layers are found in the crystal lattice of D5symmetric CRP, which alternately stack to form a threedimensional achiral honeycomb lattice, whereas C2-symmetric CRP forms dimer units via π-π stacking of the relatively planar components. Transistor devices demonstrate that, without any π-π stacking, the honeycomb lattice could also facilitate electron transport. The unique example of hybrid graphenoid with a honeycomb crystal lattice expands the scope of complex supramolecular architectures, piquing interest in applications in organic electronics.

The Supporting Information is available free of charge on the ACS Publications website. Details of synthetic methods and characterizations, spectroscopic details, OFET device fabrication and characterization, and theoretical calculation methods, including supporting Figures and Tables (PDF file).

* [email protected] * [email protected] * [email protected]

The authors declare no competing financial interest.

Figure 5. Comparison of B3LYP/6-311+G(2d,p) calculated NICS values along the symmetry axis for CRP-1 D5 symmetry (blue), corannulene D5h (green), and benzene D6h symmetry (red).

Aromaticity. Computing a 3-D plot of the nuclear independent chemical shift (NICS)16 can provide an impression of the aromatic character of a molecule in comparison to a representative known system. In the case of the systems discussed in this work, in particular CRP-1 D5, comparison with corannulene D5h17 and benzene D6h as computed along the central axis perpendicular to the molecular plane (e.g., from the plane of the molecule to 5.5 Å above the plane)18,19 establishes the basic trends. In particular, the generally accepted NICS(XX-in-plane), NICS(ZZ) and NICS(Iso) curves are depicted in Figure 5. By symmetry, the π part of the NICS(ZZ) term must vanish in the plane. In contrast, the σ component is neglected at distances around 1.0–1.5 Å above the plane; thus the choice to use these as the classical “aromatic” NICS positions. The comparison among CRP-1 D5, corannulene D5h, and benzene D6h NICS(Iso) curves shows that the nuclear magnetic shift profile for CRP-1 and corannulene are essentially identical, and are magnified compared to that of benzene.

The novel nonplanar graphenoid structure, corannurylene pentapetalae, has been synthesized via hybridization of five PDI

We greatly appreciate the following students: Y. Xia of Tianjin University for the excitation and emission spectra characterization; C. Liu of ICCAS for the assistance with OFET measurements. This work was financially supported by the National Natural Science Foundation of China (NSFC) (No. 21672221, 21734009, 21790361, 91427303), the 973 Program (2015CB856502), and the Youth Innovation Promotion Association of Chinese Academy of Sciences. K.K.B. and J.S.S. are grateful to the National Basic Research Program of China (2015CB856500), the Qian Ren Scholar Program of China, and the Synergetic Innovation Center of Chemical Science and Engineering (Tianjin) for support of this work. The computations were done on the Arran cluster supported by the Health Sciences Platform (HSP) at Tianjin University, P. R. China. We are grateful to the staff of BL17B beamline at National Center for Protein Science Shanghai and Shanghai Synchrotron Radiation Facility, for assistance during data collection.

(1) (a) Chen, Z.; Wang, H. I.; Teyssandier, J.; Mali, K. S.; Dumslaff, T.; Ivanov, I.; Zhang, W.; Ruffieux, P.; Fasel, R.; Räder, H. J.; Turchinovich, D.; Feyter, S. D.; Feng, X.; Kläui, M.; Narita, A.; Bonn, M.; Müllen, K. Chemical Vapor Deposition Synthesis and Terahertz Photoconductivity of Low-Band-Gap N = 9 Armchair Graphene Nanoribbons. J. Am. Chem. Soc. 2017, 139, 3635-3638. (b) Zhou, F.; Huang, H.; Xiao, C.; Zheng, S.; Shi, X.; Qin, J.; Fu, Q.; Bao, X.; Feng, X.; Müllen, K.; Wu, Z.-S. Electrochemically Scalable Production of Fluorine-Modified Graphene for Flexible and High-Energy IonogelBased Microsupercapacitors. J. Am. Chem. Soc. 2018, 140, 81988205. (c) Koga, Y.; Kaneda, T.; Saito, Y.; Murakami, K.; Itami, K. Synthesis of Partially and Fully Fused Polyaromatics by Annulative Chlorophenylene Dimerization. Science 2018, 359, 435-439. (d) Sei-

ACS Paragon Plus Environment

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

fert, S.; Shoyama, K.; Schmidt, D.; Würthner, F. An Electron-Poor C64 Nanographene by Palladium-Catalyzed Cascade C-C Bond Formation: One-Pot Synthesis and Single-Crystal Structure Analysis. Angew. Chem. Int. Ed. 2016, 55, 6390-6395. (e) Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. MoS2 Nanoparticles Grown on Graphene: An Advanced Catalyst for the Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2011, 133, 7296-7299. (f) Heerema, S. J.; Dekker, C. Graphene Nanodevices for DNA Sequencing. Nat. Nanotechnol. 2016, 11, 127-136. (2) (a) Wu, Y.-T.; Siegel, J. S. Aromatic Molecular-Bowl Hydrocarbons: Synthetic Derivatives, Their Structures, and Physical Properties. Chem. Rev. 2006, 106, 4843-4867. (b) Tsefrikas, V. M.; Scott, L. T. Geodesic Polyarenes by Flash Vacuum Pyrolysis. Chem. Rev. 2006, 106, 4868-4884. (c) Juríček, M.; Strutt, N. L.; Barnes, J. C.; Butterfield, A. M.; Dale, E. J.; Baldridge, K. K.; Stoddart, J. F.; Siegel, J. S. Induced-fit Catalysis of Corannulene Bowl-to-bowl Inversion. Nat. Chem. 2014, 6, 222-228. (d) Valenti, G.; Bruno, C.; Rapino, S.; Fiorani, A.; Jackson, E. A.; Scott, L. T.; Paolucci, F.; Marcaccio, M. Intense and Tunable Electrochemiluminescence of Corannulene. J. Phys. Chem. C. 2010, 114, 19467-19472. (e) Seiders, T. J.; Baldridge, K. K.; Siegel, J. S. Synthesis and Characterization of the First Corannulene Cyclophane. J. Am. Chem. Soc. 1996, 118, 2754-2755. (3) (a) Ball, M.; Zhong, Y.; Wu, Y.; Schenck, C.; Ng. F.; Steigerwald, M.; Xiao, S.; Nuckolls, C. Contorted Polycyclic Aromatics. Acc. Chem. Res. 2015, 48, 267-276. (b) Pun, S. H.; Miao, Q. Toward Negatively Curved Carbons. Acc. Chem. Res. 2018, 51, 1630-1642. (c) Schuster, N. J.; Sánchez, R. H.; Bukharina, D.; Kotov, N. A.; Berova, N.; Ng, F.; Steigerwald, M. L.; Nuckolls, C. A Helicene Nanoribbon with Greatly Amplified Chirality. J. Am. Chem. Soc. 2018, 140, 62356239. (d) Fan, W.; Winands, T.; Doltsinis, N. L.; Li, Y.; Wang, Z. A Decatwistacene with an Overall 170°Torsion. Angew. Chem. Int. Ed. 2017, 56, 15373-15377. (e) Liu, G.; Koch, T.; Li, Y.; Doltsinis, N. L.; Wang, Z. Nanographene Imides Featuring Dual-Core Sixfold [5]Helicenes. Angew. Chem. Int. Ed. 2019, 58, 178-183. (f) Wang, Y.; Yin, Z.; Zhu, Y.; Gu, J.; Li, Y.; Wang, J. Hexapole [9]Helicene. Angew. Chem. Int. Ed. 2019, 58, 587-591. (4) (a) Kawasumi, K.; Zhang, Q.; Segawa, Y.; Scott, L. T.; Itami, K. A Grossly Warped Nanographene and the Consequences of Multiple Odd-membered-ring Defects. Nat. Chem. 2013, 5, 739-744. (b) Lin, H.-A.; Sato, Y.; Segawa, Y.; Nishihara, T.; Sugimoto, N.; Scott, L. T.; Higashiyama, T.; Itami, K. A Water-Soluble Warped Nanographene: Synthesis and Applications for Photoinduced Cell Death. Angew. Chem. Int. Ed. 2018, 57, 2874-2878. (c) Cheung, K. Y.; Xu, X.; Miao, Q. Aromatic Saddles Containing Two Heptagons. J. Am. Chem. Soc. 2015, 137, 3910-3914. (d) Cheung, K. Y.; Chan, C. K.; Liu, Z.; Miao, Q. A Twisted Nanographene Consisting of 96 Carbon Atoms. Angew. Chem. Int. Ed. 2017, 56, 9003-9007. (e) Feng, C.-N.; Kuo, M.-Y.; Wu, Y.-T. Synthesis, Structural Analysis, and Properties of [8]Circulenes. Angew. Chem. Int. Ed. 2013, 52, 7791-7794. (f) Sakamoto, Y.; Suzuki, T. Tetrabenzo[8]circulene: Aromatic Saddles from Negatively Curved Graphene. J. Am. Chem. Soc. 2013, 135, 14074-14077. (5) (a) Tan, Y.-Z.; Yang, B.; Parvez, K.; Narita, A.; Osella, S.; Beljonne, D.; Feng, X.; Müllen, K. Atomically Precise Edge Chlorination of Nanographenes and Its Application in Graphene Nanoribbons. Nat. Commun. 2013, 4, 2646. (b) Wang, Z.; Dötz, F.; Enkelmann, V.; Müllen, K. “Double-Concave” Graphene: Permethoxylated Hexaperi-hexabenzocoronene and Its Cocrystals with Hexafluorobenzene and Fullerene. Angew. Chem. Int. Ed. 2005, 44, 1247-1250. (c) Wasserfallen, D.; Kastler, M.; Pisula, W.; Hofer, W. A.; Fogel, Y.; Wang, Z.; Müllen, K. Suppressing Aggregation in a Large Polycyclic Aromatic Hydrocarbon. J. Am. Chem. Soc. 2006, 128, 1334-1339. (6) (a) Jiang, W.; Li, Y.; Wang, Z. Tailor-Made Rylene Arrays for High Performance n-Channel Semiconductors. Acc. Chem. Res. 2014, 47, 3135-3147. (b) Usta, H.; Facchetti, A.; Marks, T. J. n-Channel Semiconductor Materials Design for Organic Complementary Circuits. Acc. Chem. Res. 2011, 44, 501-510. (c) Gsänger, M.; Oh, J. H.; Könemann, M.; Höffken, H. W.; Krause, A.-M.; Bao, Z.; Würthner, F. A Crystal-Engineered Hydrogen-Bonded Octachloroperylene Diimide with a Twisted Core: An n-Channel Organic Semiconductor. Angew. Chem. Int. Ed. 2010, 49, 740-743. (d) Liu, C.; Xiao, C.; Li, Y.; Hu, W.; Li, Z.; Wang, Z. High Performance, Air Stable n-Type Single

Page 6 of 8

Crystal Transistors Based on Core-tetrachlorinated Perylene Diimides. Chem. Commun. 2014, 50, 12462-12464. (e) Li, Y.; Tan, L.; Wang, Z.; Qian, H.; Shi, Y.; Hu, W. Air-Stable n-Type Semiconductor: Core-Perfluoroalkylated Perylene Bisimides. Org. Lett. 2008, 10, 529-532. (7) (a) Jackson, E. A.; Steinberg, B. D.; Bancu, M.; Wakamiya, A.; Scott, L. T. Pentaindenocorannulene and Tetraindenocorannulene: New Aromatic Hydrocarbon π Systems with Curvatures Surpassing That of C60. J. Am. Chem. Soc. 2007, 129, 484-485. (b) Scott, L. T.; Jackson, E. A.; Zhang, Q.; Steinberg, B. D.; Bancu, M.; Li, B. A Short, Rigid, Structurally Pure Carbon Nanotube by Stepwise Chemical Synthesis. J. Am. Chem. Soc. 2012, 134, 107-110. (c) Dutta, A. K.; Linden, A.; Zoppi, L.; Baldridge, K. K.; Siegel, J. S. Extended Corannulenes: Aromatic Bowl/Sheet Hybridization. Angew. Chem. Int. Ed. 2015, 54, 10792-10796. (d) Fernández-Garcí a, J. M.; Evans, P. J.; Rivero, S. M.; Fernández, I.; García-Fresnadillo, D.; Perles, J.; Casado, J.; Martín, N. π-Extended Corannulene-Based Nanographenes: Selective Formation of Negative Curvature. J. Am. Chem. Soc. 2018, 140, 17188-17196. (e) Shi, K.; Lei, T.; Wang, X.-Y.; Wang, J.Y.; Pei, J. A Bowl-shaped Molecule for Organic Field-effect Transistors: Crystal Engineering and Charge Transport Switching by Oxygen Doping. Chem. Sci. 2014, 5, 1041-1045. (f) Fujikawa, T.; Preda, D. V.; Segawa, Y.; Itami, K.; Scott, L. T. Corannulene−Helicene Hybrids: Chiral π-Systems Comprising Both Bowl and Helical Motifs. Org. Lett. 2016, 18, 3992-3995. (g) Yanney, M.; Fronczek, F. R.; Henry, W. P.; Beard, D. J.; Sygula, A. Cyclotrimerization of Corannulyne: Steric Hindrance Tunes the Inversion Barriers of Corannulene Bowls. Eur. J. Org. Chem. 2011, 2011, 6636-6639. (h) Sygula, A.; Fronczek, F. R.; Sygula, R.; Rabideau, P. W.; Olmstead, M. M. A Double Concave Hydrocarbon Buckycatcher. J. Am. Chem. Soc. 2007, 129, 38423843. (i) Rabideau, P. W.; Abdourazak, A. H.; Folsom, H. E.; Marcinow, Z.; Sygula, A.; Sygula, R. Buckybowls: Synthesis and ab Initio Calculated Structure of the First Semibuckminsterfullerene. J. Am. Chem. Soc. 1994, 116, 7891-7892. (8) The term ‘pentapetalae’, an existing taxonomic clad, refers to a morphology with 5-fold symmetry among 5 petals. In this context, the morphology of molecules of interest fit well with this descriptor. The term ‘corannurylene’ combines an abbreviated corannulene as prefix to rylene in order to convey the hybridization of these structural motifs in these molecules. Although it may be more rigorous to use corannul- instead of corannu- as the combining form, dropping the ‘l’ makes the combination much easier to read and to say. (9) Rajasingh, P.; Cohen, R.; Shirman, E.; Shimon, L. J. W.; Rybtchinski, B. Selective Bromination of Perylene Diimides under Mild Conditions. J. Org. Chem. 2007, 72, 5973-5979. (10) Eliseeva, M. N.; Scott, L. T. Pushing the Ir-Catalyzed C−H Polyborylation of Aromatic Compounds to Maximum Capacity by Exploiting Reversibility. J. Am. Chem. Soc. 2012, 134, 15169-15172. (11) Petrukhina, M. A.; Andreini, K. W.; Mack, J.; Scott, L. T. X-ray Quality Geometries of Geodesic Polyarenes from Theoretical Calculations: What Levels of Theory Are Reliable? J. Org. Chem. 2005, 70, 5713-5716. (12) Kato, K.; Segawa, Y.; Scott, L. T.; Itami, K. A Quintuple [6]Helicene with a Corannulene Core as a C5-Symmetric PropellerShaped π-System. Angew. Chem. Int. Ed. 2018, 57, 1337-1341. (13) Zhang, Q.; Yang, X.; Li, P.; Huang, G.; Feng, S.; Shen, C.; Han, B.; Zhang, X.; Jin, F.; Xu, F.; Lu, T. J. Bioinspired Engineering of Honeycomb Structure – Using Nature to Inspire Human Innovation. Prog. Mater. Sci. 2015, 74, 332-400. (14) Li, R.; Jiang, L.; Meng, Q.; Gao, J.; Li, H.; Tang, Q.; He, M.; Hu, W.; Liu, Y.; Zhu, D. Micrometer-Sized Organic Single Crystals, Anisotropic Transport, and Field-Effect Transistors of a Fused-Ring Thienoacene. Adv. Mater. 2009, 21, 4492-4495. (15) Bronshtein, I.; Iron, M. A.; Rybtchinski, B. Organic Phototransistors Based on Perylene Diimide Nanocrystals Lacking π-π Interactions. J. Mater. Chem. C, 2018, 6, 10597-10602. (16) Schleyer, P. v. R.; Maerker, C.; Dransfeld, A.; Jiao, H.; Hommes, N. J. R. v. E. Nucleus-Independent Chemical Shifts: A Simple and Efficient Aromaticity Probe. J. Am. Chem. Soc. 1996, 118, 6317-6318.

ACS Paragon Plus Environment

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

Journal of the American Chemical Society (17) Steiner, E.; Fowler, P. W.; Jenneskens, L. W. Counter-Rotating Ring Currents in Coronene and Corannulene. Angew. Chem. Int. Ed. 2001, 40, 362-366. (18) Schleyer, P. v. R.; Jiao, H.; Hommes, N. J. R. v. E.; Malkin, V. G.; Malkina, O. L. An Evaluation of the Aromaticity of Inorganic Rings: Refined Evidence from Magnetic Properties. J. Am. Chem. Soc. 1997, 119, 12669-12670. (19) Schleyer, P. v. R.; Manoharan, M.; Wang, Z.-X.; Kiran, B.; Jiao, H.; Puchta, R.; Hommes, N. J. R. v. E. Dissected NucleusIndependent Chemical Shift Analysis of π-Aromaticity and Antiaromaticity. Org. Lett. 2001, 3, 2465-2468.

ACS Paragon Plus Environment

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

Page 8 of 8

Table of Contents

ACS Paragon Plus Environment

8