Katsuma
Matsui
a,
Yasutomo
Segawa
*a,
Tomotaka
Namikawa
bc,
Kenji
Kamada
*bc and
Kenichiro
Itami
*a
aDepartment of Chemistry, Graduate School of Science, Nagoya University, Nagoya 464-8602, Japan. E-mail: ysegawa@nagoya-u.jp; itami.kenichiro@a.mbox.nagoya-u.ac.jp; Fax: +81 52-788-6098; Tel: +81 52-788-6098
bResearch Institute for Ubiquitous Energy Devices, National Institute of Advanced Industrial Science and Technology (AIST), Ikeda, Osaka 563-8577, Japan. E-mail: k.kamada@aist.go.jp; Fax: +81 72-751-9637; Tel: +81 72-751-9523
cDepartment of Chemistry, School of Science and Technology, Kwansei Gakuin University, Sanda, Hyogo 669-1337, Japan
First published on 29th August 2012
The first synthesis of an all-benzene carbon nanocage 1, which represents a junction unit of branched carbon nanotubes, has been achieved. A stepwise assembly of six L-shaped units [cis-di(p-bromophenyl)cyclohexane derivative] and two three-way units (1,3,5-triborylbenzene) by cross-coupling and homocoupling provided the unstrained cyclic precursor. Acid-mediated aromatization of cyclohexane moieties in the precursor resulted in the formation of carbon nanocage 1. Photophysical measurements and DFT studies revealed the unique properties of 1, such as D3 symmetry with degenerate HOMO/HOMO − 1 and LUMO/LUMO + 1, high fluorescence quantum yield (ΦF = 0.87), and a relatively large two-photon absorption cross section (500 GM at 590 nm), which are attractive for various applications.
We became interested in all-benzene carbon nanocages as part of our ongoing efforts to synthesize structurally uniform carbon nanotubes using carbon nanorings, such as cycloparaphenylene (a short sidewall segment of armchair carbon nanotubes), as templates (Fig. 1, top). Recently, the chemical synthesis of cycloparaphenylene (CPP)6 has been accomplished by the groups of Bertozzi,7 Itami,8–13 Yamago,14,15 and Jasti7,16,17 after many years.6,18 The modular,10 size-selective,8,10,13–17 and scalable11,12 synthesis of [n]CPPs is now possible, and [12]CPP has become commercially available.19 CPP is an interesting and unique molecular entity not only because it represents the shortest sidewall segment of armchair carbon nanotube structures but also because of its aesthetic appeal,6 unique cyclic conjugation, photophysical properties,7,14,15,20 and guest-encapsulating properties.11,12,21 More recently, short sidewall segments of chiral carbon nanotubes,22,23 pyridine-inserted CPP,24 and π-extended CPP (cyclo-1,4-naphthylene)25 have been also synthesized.
Fig. 1 Carbon nanoring and carbon nanocage as segments of straight and branched carbon nanotubes. |
In addition to straight carbon nanotubes, branched or so-called Y-junction carbon nanotubes have also attracted attention as a small logic gates or small transistors (Fig. 1, bottom).26 Our simple retrosynthetic analysis of these structures led to the identification of a tri-bridged carbon nanocage consisting solely of benzene rings as a junction unit for branched carbon nanotubes. Moreover, the synthesis of such carbon nanocages is of great interest, as they would possess unique photophysical properties due to the hitherto unknown symmetric structures with a mix of linear and branched conjugations.27
For the synthesis of carbon nanocages, we used our synthesis platform for CPPs as a blueprint (Fig. 2). We previously reported that U-shaped units synthesized from L-shaped cis-diphenylcyclohexane units and a linear phenylene unit could be coupled by Suzuki–Miyaura coupling or nickel-mediated homocouplings to afford strain-free cyclic precursors (Fig. 2, left). The cyclohexane moieties in the precursors were aromatized to afford [14]–[16]CPP. We envisaged that the carbon nanocage 1, consisting of 20 benzene rings, could be synthesized by using 1,3,5-triborylbenzene as a “three-way” unit to assemble the L-shaped unit (Fig. 2, right).
Fig. 2 Our strategy for the synthesis of all-benzene macrocycles. |
Scheme 1 Synthesis of carbon nanocage 1. Reaction conditions: (a) PdCl2(dppf), K2CO3, DMF, 90 °C; (b) Ni(cod)2, 2,2′-bipyridyl, DMF, 90 °C; (c) NaHSO4·H2O, o-chloranil, m-xylene/DMSO, reflux. |
Characteristic signals of one singlet and three sets of doublets observed in the 1H NMR spectrum of 1 clearly support its D3h symmetric structure in solution state. The structure of 1 was also confirmed by the observation of 14 independent signals in the 13C NMR spectrum as well as HH COSY, HMQC, and high-resolution MALDI-TOF MS. Similarly to CPPs, carbon nanocage 1 has high solubility in most organic solvents such as CHCl3, CH2Cl2, EtOAc, acetone, THF, Et2O, benzene, and toluene.
To gain insight into the structural properties of 1, DFT calculations were performed (B3LYP/6-31G(d)). The optimized structure of 1 is shown in Fig. 3. The conformation with D3 symmetry was found to be the ground state of 1. As expected, 1 forms a beautiful cage-shape structure with a spherical void inside. The two tri-substituted benzene rings deviate from each other at an angle of 12.5°. The distance between these two benzene rings is 18.4 Å, a value which is between the diameters of [13]CPP (17.9 Å) and [14]CPP (19.4 Å).9 Other di-substituted benzene rings in 1 are slightly bent to form three arches. The strain energies of carbon nanocage 1 and a model of cyclic precursor 5 (5-H: all methoxymethoxy groups in 5 are removed) were determined to be 67.2 kcal mol−1 (1) and 2.1 kcal mol−1 (5-H) by using the hypothetical homodesmotic reaction (see the ESI for details†).9 It is clear from these values that the six cyclohexane moieties of 5 effectively help in attenuating the build-up of strain energy during the macrocyclization (4 → 5).
Fig. 3 Optimized structure of 1. (a) Top view. (b) Side view. |
Fig. 4 (a) UV-vis absorption (solid line) and fluorescence spectra (broken line) of 1. (b) Blue fluorescence emission of the chloroform solution of 1. |
We further performed time-dependent DFT (TD-DFT) studies of the optimized structure of 1 (B3LYP/6-31G(d)). The energy diagrams and pictorial representations of frontier MOs of 1 are shown in Fig. 5. Four orbitals, HOMO − 1, HOMO, LUMO, and LUMO + 1, have e-symmetry whereas HOMO – 2 and LUMO + 2 have a2- and a1-symmetries, respectively. Degeneracy in energy levels was found for the relevant pairs of frontier orbitals, HOMO − 1/HOMO (−5.39 eV) and LUMO/LUMO + 1 (−1.58 eV). The HOMO–LUMO gap of 3.81 eV is slightly higher than that of [18]CPP (3.71 eV (ref. 20)). The HOMO → LUMO and HOMO − 1 → LUMO + 1 transitions mainly contribute to the transition to the S3-state, which has a high oscillator strength (f = 3.04). This high oscillator strength value should be the reason for the high kr value of 1. On the other hand, the HOMO − 1 → LUMO and HOMO → LUMO + 1 transitions leading to the S4-state were found to be symmetrically forbidden (A1, f = 0.00).
Fig. 5 Energy diagrams and pictorial representations of the frontier MOs of 1 calculated at the B3LYP/6-31G(d) level of theory. Excitation energies were computed by TD-DFT at the same level. Values in parentheses represent oscillator strengths (f). |
To investigate more about the nature of the electronic excited states, thereby exploring new applications of 1, we measured the two-photon absorption (TPA) spectrum by the femtosecond open-aperture Z-scan method (Fig. 6). The TPA spectrum of 1, taken in CHCl3, is broad, spanning 640 nm to 480 nm, with a peak value of the TPA cross section of 500 ± 60 GM at 590 nm, beside the maximum value of 540 ± 80 GM at the shorter wavelength edge of the measurements (480 nm). These values are relatively large for a hydrocarbon molecule without any electron donor or acceptor substituents.30 The peak wavelength (590 nm) of 1 shows a large bathochromic shift compared to that of didecyl-substituted p-septiphenyl, which is a linear paraphenylene and can be a model compound of each arc of 1, measured under the same conditions (also in Fig. 6). This shows the exciton coupling among the arcs occurs in 1. Previously the exciton coupling among the arms was reported to enhance the TPA for a benzene-cored tribranched molecule.31 However, the peak cross section of 1 (500 GM) is close to or less than three times that of the p-septiphenyl (190 GM), suggesting the exciton coupling has little effect on the magnitude of TPA beside the drawback due to the curved shape of each arc. Moreover, the broad TPA spectrum of 1 is considered to consist of many TPA transitions in the wavelength range. This is supported by the TD-DFT calculations, which show that there are many TPA-allowed excited states [A1-states: S4 (3.57 eV), S7 (3.66 eV), and S13 (3.81 eV)], which can show strong TPA transitions beside E-states, which can show weak transitions in the wavelength range.
Fig. 6 Two-photon absorption cross-section spectra of 1 (circles) and 2′′′,5′′′-didecyl-p-septiphenyl (rhombi) in CHCl3. 1 GM = 10−50 cm4 s per molecule per photon. |
Footnote |
† Electronic supplementary information (ESI) available: Experimental procedures, characterization of data for all new compounds, and details of photophysical and computational studies. See DOI: 10.1039/c2sc21322b |
This journal is © The Royal Society of Chemistry 2013 |