Abstract
[n]Peri-acenes ([n]PA) have attracted great interest as promising candidates for nanoelectronics and spintronics. However, the synthesis of large [n]PA (n > 4) is extremely challenging due to their intrinsic open-shell radical character and high reactivity. Herein, we report the successful synthesis and isolation of a derivative (1) of peri-hexacene in crystalline form. The structure of 1 is unequivocally confirmed by X-ray crystallographic analysis. Its ground state, aromaticity and photophysical properties are systematically studied by both experimental methods and theoretical calculations. Although the parent peri-hexacene is calculated to have a very large diradical character (y0 = 94.5%), 1 shows reasonable stability (t1/2 = 24 h under ambient conditions) due to the kinetic blocking. 1 exhibits an open-shell singlet ground state with a small singlet-triplet energy gap (−1.33 kcal/mol from SQUID measurements). 1 has also a narrow HOMO-LUMO energy gap (1.05 eV) and displays amphoteric redox behavior. This work opens new avenues for the design and synthesis of stable zigzag-edged graphene-like molecules with significant diradical character.
Subject terms: Electronic properties and devices, Magnetic properties and materials, Synthesis of graphene
[n]Peri-acenes have attracted interest as promising candidates for nanoelectronics and spintronics but the synthesis of large (n > 4) [n]peri-acenes is extremely challenging due to their intrinsic open-shell radical character and high reactivity. Herein, the authors report the synthesis and isolation of a derivative of peri-hexacene in crystalline form.
Introduction
Atomically precise nanographenes (NGs) have gained intensive attention in the past decade because of their fascinating optoelectronic properties and high promise for next-generation semiconductor materials1–10. Both experimental and computational studies have demonstrated that widths and edge topologies of NGs play a pivotal role in determining their electronic and magnetic properties as well as their chemical reactivities11–13. Among various NG families, [n]peri-acenes ([n]PAs), where n defines the number of benzene rings formed along the zigzag axis, are a unique class of rectangle-shaped NGs consisting of two zigzag-edges and two arm-chair-edges, which can exhibit, if n exceeds a certain threshold (n > 3), diradical and even polyradical character upon the lateral π-extension due to the gain of additional Clar’s aromatic sextets in their open-shell resonance forms (Fig. 1)14–21. They can also be considered as ideal model molecules to explore the electronic structures of the large-size NGs with zigzag edges22–24. However, their synthesis is a huge challenge owing to their intrinsic open-shell radical character, which is related to the presence of a localized non-bonding π-state around parallel zigzag edges. To date, only a few [n]PA molecules have been successfully synthesized. For example, in 2018, Feng and Wu’s groups independently reported the in-solution synthesis of peri-tetracene derivatives ([4]PA, Fig. 1a), showing large diradical character with a half-life of 3–7 h in solution under ambient conditions25,26. In 2021, Feng’s group described the in-solution synthesis and in-situ characterization of hitherto the largest peri-heptacene derivative ([7]PA, Fig. 1c), which demonstrated a significant tetraradical character with a half-life of ~25 min under inert conditions27. It should be mentioned that parent peri-pentacene ([5]PA, Fig. 1b) has been accomplished on Au (III) surface under ultra-high vacuum conditions and its singlet open-shell nature was revealed by scanning tunneling spectroscopy28, but all attempts so far to synthesize [5]PA or its derivatives in solution have been unsuccessful due to the poor solubility and undesired side reactions29,30. Therefore, the solution-phase synthesis of large [n]PAs (n > 4) or their appropriate derivatives depends heavily on the development of more effective design and synthetic strategies, which need to overcome the challenges associated with the low solubility and the intrinsic high reactivity of the target molecules.
Peri-hexacene ([6]PA, Fig. 1d), which has been achieved difficultly by solution-mediated or surface-assisted synthesis, is the next important target of synthesis and isolation, and its realization is also crucial for elucidating the structure-property relationship of this family from both fundamental and applied perspectives. In this context, herein we report the solution-phase synthesis and characterization of a stable peri-hexacene derivative (1, Fig. 1d). To obtain preliminary insights into the stability and properties of 1, Clar’s aromatic π-sextet rule was utilized. The Clar’s rule is a fundamental concept in the chemistry of polycyclic aromatic hydrocarbons (PAHs) and posits that the Kekulé resonance structure with the highest number of disjoint benzene-like aromatic π-sextets—referred to as “Clar’s sextets”—has the most contribution to the properties of the PAH. Generally, an isomeric PAH with more Clar’s sextets is expected to exhibit higher stability14–16. Recent studies have highlighted the importance of additional aromatic sextet rings on stabilizing the diradical structure25–28,31–38. Specifically, the gain of one or more additional Clar’s sextets in the open-shell form is crucial for achieving a stable ground-state singlet diradicaloid with high diradical character. According to this rule, the molecule can be drawn in at least three different resonance forms: a closed-shell form with two aromatic sextet rings (the hexagons shaded in blue, form A), an open-shell diradical form with five aromatic sextet rings (form B), and an open-shell tetraradical form with six aromatic sextet rings (form C). The gain of three/four Clar’s aromatic sextet rings in the open-shell resonance forms B/C implies a large diradical character and moderate tetraradical character. Our spin-unrestricted theoretical calculations (UB3LYP/6-31 G(d,p) on the parent [6]PA indicated that the frontier SOMO-α and SOMO-β orbitals have significant disjoint features, and the high spin densities are mainly distributed along the two zigzag edges (Fig. 1e). Thus, the diradical character (y0) and tetraradical character (y1) of the parent [6]PA were calculated to be 94.5% and 24.5%, respectively, by the Natural orbital occupation number (NOON) calculations (UCAM-B3LYP/6-31G(d,p))39, which is much larger than [4]PA (y0 = 52.5%) and [5]PA (y0 = 75.4%) under the same level of theory. The significant open-shell radical character of [6]PA suggests the need to kinetically block the most reactive zigzag edges by bulky groups and to develop an innovative strategy to build up the framework. In response to these challenges, a soluble and stable [6]PA derivative 1 was successfully synthesized and isolated in a crystalline form. Its ground state, aromaticity and photophysical properties were systematically explored by both experimental methods and theoretic calculations and compared with those of [4]PA and [7]PA derivatives.
Results
To realize the synthesis of [6]PA with sufficient stability and solubility, our strategy is to introduce four bulky and electron-withdrawing 2,6-dichlorophenyl groups at the most reactive sites of the two zigzag edges and to incorporate C(sp3)-bridged cyclopenta (CP) rings at the bay regions of [6]PA where each sp3 carbon substituted by two 3,5-di-tert-butylphenyl groups. The incorporation of CP rings in these bay regions not only enhances solubility but also restrain the Diels-Alder reaction in these areas, a phenomenon observed in the previous report concerning the synthesis of peri-tetracene derivative25. In addition to these substituents, four additional 3,5-di-tert-butylphenyl groups are also attached onto the π-conjugated framework to further improve solubility. As depicted in Fig. 2, the key intermediate for the synthesis of the target molecule 1 is the partially fused dibromo-intermediate 7, which is prepared via multiple cross-coupling, regioselective cyclization and oxidative dehydrogenation reactions. The CP ring-fused perylene 2 was synthesized from parent perylene according to our reported procedure40,41. The monobromo CP-fused perylene 3 was firstly synthesized by regio-selective bromination with bromine under controlled conditions at the bay region, and then Miyaura borylation reaction of 3 with bis(pinacolato)diboron (B2(pin)2)in the presence of palladium catalyst provided monoboronic ester 4 in 63% yield. After that, the Suzuki coupling of 5 with 4 yielded the intermediate 6 in 68% yield. Next, the key intermediate 7 was obtained by the treatment of 6 with CuBr2 in the presence of base where a one-pot reaction of benzannulation and intramolecular oxidative cyclodehydrogenation was achieved. This method differs from the stepwise syntheses typically reported for peri-acene derivatives25–27. Our one-pot strategy not only streamlines the procedure but also enhances overall efficiency, thereby providing an alternative pathway for the construction of other graphene fragments. The structure of 7 was confirmed by single-crystal analysis (Supplementary Fig. 22). Afterwards, lithium-halogen exchange reaction by treatment of 7 with n-BuLi proceeds and subsequent quenching with 2,6-dichlorobenzaldehyde to generate the diol, which underwent an intramolecular Friedel-Crafts cyclization in the presence of trifluoromethanesulfonic acid (TfOH) to afford the dihydro- precursor 8 in 53% yield over three steps from 7. Finally, the deprotonation of 8 with tBuOK and 18-crown-6 in THF followed by the oxidation with p-chloranil furnished the target peri-hexacene derivative 1 in 73 % yield. Because two cyclopenta (CP) ring was attached onto the arm-chair edges and the four bulky and electron-withdrawing 2,6-dichlorophenyl groups are attached onto the most reactive zig-zag edges, compound 1 shows good solubility and stability, and can be purified by triethylamine deactivated silica gel column. Stability test of the solution of 1 in carbon tetrachloride under ambient air and light conditions followed by UV/Vis-NIR absorption measurements revealed a half-life time (t1/2) of about 24 h when monitored at 617 nm (Supplementary Fig. 1). During the submission of our work, Wu et al. reported the synthesis of both peri-pentacene and peri-hexacene derivatives utilizing a similar synthetic strategy42. Their studies indicated that the peri-hexacene derivative exhibited remarkable stability, with its absorbance at the maximum decreasing by only 1% after 72 hours of exposure to ambient air and light, which surpassed the stability of our peri-hexacene derivative 1 under similar conditions. The exceptional stabilities observed in their work can be ascribed to the protection of the zigzag edges with bulky groups through a synergistic captodative effect, wherein the electron donor and acceptor collaboratively stabilize the radicals by forming a zwitterionic radical structure. In contrast to the stabilization strategy based on kinetic blocking combined with a synergistic captodative effect, our design employed kinetic protection of both zigzag edges and bay regions by bulky groups with similar electronic characteristics, resulting in reasonable stability. These findings offer valuable insights for the synthesis of a broader range of graphene fragments with zigzag edges in the future.
Single crystals of 1 suitable for X-ray crystallographic analysis were grown by slow diffusion of acetonitrile into its solution in toluene. The molecule has a centrosymmetric geometry with a slightly distorted skeleton (Fig. 3a, b). The dihedral angles between the π-conjugated skeleton and its 2,6-dichlorophenyl substituents range from 83.1 to 84.5o (Fig. 3a). Bond length analysis of the 1 backbone revealed that the bond lengths of the CC bonds (the bonds shown in red, Fig. 3c) linking the two hexacene units were in the range of 1.409-1.429 Å, which are shorter than that in [4]PA derivative (1.442–1.459 Å)26, indicating enhanced electronic coupling between two hexacene units. The CC bonds shown in blue (Fig. 3c) are much longer (1.464 Å) like a typical C(sp2)-C(sp2) single bond, which could be attributed to the elongation induced by the steric hindrance between the neighboring 2,6-dichlorophenyl substituenets. The calculated harmonic oscillator model of aromaticity (HOMA) values43 based on the X-ray structures revealed that the four benzenoid rings at the termini (rings A/F/L/Q) have largest HOMA values (0.87/0.84, Fig. 3c), indicating a large aromatic character. The rings C/D/N/O have much smaller HOMA values (0.38 ~ 0.44, Fig. 3c) while the other benzenoid rings have a HOMA value of 0.65–0.80 (Fig. 3c), indicating significant contribution of the open-shell diradical form B with five localized aromatic sextet rings shown in Fig. 1d to the ground-state structure. On the other hand, nucleus independent chemical shift (NICS)44 calculations show that these five benzenoid rings have large negative NICS(1)zz values (−20.49 ppm to −22.20 ppm), the rings B/E/H/J/M/P have moderate negative NICS(1)zz values (−11.93 ppm to −16.00 ppm) while the rings C/D/N/O are almost non-aromatic (Fig. 4a). All these suggest that [6]PA possesses a large diradical character, which can be stabilized by five aromatic sextet rings. In addition, no intermolecular π-π interaction was observed in the 3D packing structure (Supplementary Fig. 23).
To gain a deeper understanding of the electronic structure and aromaticity of compound 1, we carried out anisotropy of induced current-density (ACID)45 and isochemical shielding surface (ICSS)46,47 calculations (B3LYP/6-31 G(d,p)). ACID plot shows obvious clockwise diatropic ring current circuit along the periphery (Fig. 4b), indicating that the molecule is globally aromatic and the two hexacene units are coupled well with each other, in accordance with the bond length analysis. Notably, the central anthracene units show clockwise diatropic ring current (Fig. 4b), in agreement with the HOMA value and NICS analysis. Furthermore, the calculated 2D ICSS map also shows that the rings A/F/L/Q have the strongest shielding, the rings B/E/H/J/M/P have moderate shielding, and there is the least shielding in other benzenoid rings (Fig. 4c), consistent with the NICS and ACID calculations. For comparison, although the open-shell diradical form of [4]PA exhibits a localized aromatic character (Fig. 1a), the central benzenoid ring connecting the two tetracene units in[4]PA is almost non-aromatic26. The central benzenoid rings connecting the two hexacene/heptacene units in [6]PA and [7]PA27 are aromatic. This difference could be ascribed to the much larger diradical character of [6]PA and [7]PA compared to [4]PA. Therefore, the lateral extension in [6]PA and [7]PA provides the opportunity to form local aromatic sextets along the central row of the backbone, resulting in different electronic structures.
To further understand the ground-state electronic structure of compound 1, we conducted magnetic measurements such as variable temperature (VT) NMR, electron spin resonance (ESR) and superconducting quantum interference device (SQUID) measurements. The 1H NMR spectrum of 1 in THF-d8 at room temperature exhibited one almost flat baseline and cooling of sample to −80 oC also did not obtain full resolution (Supplementary Fig. 3). The NMR broadening can be explained by the thermal population of triplet diradical species at elevated temperatures, as expected for species with strong diradical character. Further experimental evidence was obtained by EPR measurements on 1 in solid and solution. The EPR intensities of compound 1 in powder state and dilute toluene solution decreased with the lowering of temperature (Fig. 5b and Supplementary Fig. 6a), and fitting of the data in both solid state and frozen solution by using Bleaney–Bowers equation48 gave a singlet-triplet energy gap (∆ES-T) of −1.41 kcal/mol and −1.74 kcal/mol, respectively (Fig. 5c and Supplementary Fig. 6b). In the solid states, we could observe the weak axially symmetric triplet state EPR signals corresponding to the ΔMs = ±1 transition (Fig. 5a), and the spectral simulation showed the spin-Hamiltonian parameters of S = 1, g = 2.0022, |D/hc| = 0.00825 cm−1, and |E/hc| = 0.00191 cm−1. Accordingly, the spin distance was determined to be 6.9 Å, which close to distance of the two zigzag edges of hexacene (7.06 Å) in the crystals. Rabi cycles at 70 K by an echo-detected nutation experiment of compound 1 was conducted (Fig. 5d), and a stable tris(2,4,6-trichlorophenyl)methyl (TTM, S = 1/2) radical is employed as an external standard. The results show that the slope of the linear dependence of Rabi frequency on the B1 magnetic field is approximately times that of a doublet species (S = 1/2), indicating the triplet species derived by thermal excitation (Fig. 5e)49,50. The spin-lattice relaxation time (T1) and spin coherence time (Tm) of 1 were obtained using the inversion recovery method and Hahn-echo sequences (Supplementary Fig. 4 and Supplementary Fig. 5). The spin-echo decay curves of 1 were fitted with an exponential decay to extract T1 of 16.7 ms and Tm of 0.85 μs at 70 K. Compared to most reported radicaloids (Supplementary Table 2)51–55, peri-hexacene 1 exhibits a longer spin-lattice relaxation time (T1 = 16.7 ms) and a comparable spin coherence time (Tm = 0.85 μs), with Tm significantly exceeding the duration of spin manipulation pulses (on the order of 10 ns) by nearly two orders of magnitude. These characteristics highlight its potential applications in quantum information processing and two-qubit systems56–60. Furthermore, we also carried out superconducting quantum interference device (SQUID) measurements for the microcrystalline sample of 1 at 2–300 K. The magnetic susceptibility increased with increasing temperature, and the singlet-triplet energy gap (∆ES-T) was estimated to be −1.33 kcal/mol by careful fitting of the data by using the Bleaney–Bowers48 equation (Fig. 5f), consistent with the above EPR results. All these indicate that compound 1 is an open-shell singlet diradicaloid and exhibits a large diradical character with a small singlet-triplet energy gap. For comparison, the [4]PA derivative (−2.5 kcal/mol)26 showed a larger ∆ES-T value, in consistency with its smaller diradical character. On the other hand, broken-symmetry DFT calculations (UB3LYP/6-31 G(d,p)) also predict that the open-shell singlet biradical state of 1 has lower energy compared to the triplet biradical state, and the singlet-triplet energy gap (∆ES-T) is calculated to be −2.08 kcal/mol, which is consistent with the above experimental results. In addition, DFT calculations also gave a triplet-quintet energy gap (∆ET-Q) of −21.79 kcal/mol for 1. This large ∆ET-Q value is in accordance with its small tetraradical character.
The UV-vis absorption spectra of 1 was recorded in carbon tetrachloride and compared to that of precursor 8 in DCM, as illustrated in Fig. 6a. Compound 8 exhibits a well-resolved absorption maximum at 632 nm along with a small shoulder band peaking at 580 nm. In contrast, the absorption spectrum of 1 displays a broad band in the near-infrared (NIR) region with a maximum at 789 nm, which is significantly redshifted compared with that of precursor 8. Moreover, 1 shows a weak absorption band in the NIR region with maximum (λmax) at 1090 nm along with a shoulder (λsh) at 1197 nm (Fig. 6a). 1 also displays multiple intense absorption bands in the visible region. The NIR absorption band has a similar structure to that in [4]PA derivative26 and [7]PA derivative27, however, the intensity is much weaker. Furthermore, the characteristic long-wavelength shoulder peak for 1 is strong indication that it has an open-shell singlet ground state, and the band is originated from the ground-state HOMO,HOMO → LUMO,LUMO double excitation61. The optical band gap (Egopt) of 1 from the onset of its UV/Vis absorption is roughly estimated to be 0.99 eV, which is smaller than that of [4]PA derivative (1.12 eV)26 due to the π-extension.
The electrochemical property of 1 was investigated by differential pulse voltammetry (DPV) and cyclic voltammetry (CV) measurements in dichloromethane (Fig. 6b). Compound 1 displays two reversible oxidation waves with halfwave potentials E1/2ox at 0.04 V and 0.32 V, and two reversible reduction waves with half-wave potentials E1/2red at −1.03 V and −1.24 V (vs Fc/Fc+). Compared with [4]PA derivative, the neighboring oxidation or reduction waves of 1 show a smaller segregation, which can be explained by the larger π-conjugated backbone that reduces the intramolecular Coulomb charge repulsion. In addition, according to the onset potentials of the first oxidation/reduction waves, the HOMO and LUMO energy levels of 1 are estimated to be −4.82 eV and −3.76 eV, respectively. Thus, the corresponding electrochemical energy gap (EgEC) is estimated to be 1.06 eV, which is consistent with the optical energy gap. On the other hand, compound 1 can be chemically oxidized into its corresponding radical cation and dication with NO•SbF6 in anhydrous DCM. The radical cation exhibits multiple intense absorption bands in the NIR region with λmax at 858 nm, 1210 nm and 1590 nm, respectively (Supplementary Fig. 2), and a strong EPR signal was observed for the radical cation (Supplementary Fig. 6c). The dication shows an intense absorption band with λmax at 845 nm in the NIR region (Supplementary Fig. 2). All absorption spectra are in agreement with the time-dependent (TD) DFT calculations (Supplementary Figs. 8 to 21).
To reveal the excited-state dynamics of compound 1, its femtosecond transient absorption (fs-TA) spectra have been recorded in degassed DCM (Fig. 7). As shown in Fig. 7a, upon photoexcitation, two photo-induced bleaching (PIB) bands centered at ~600 and ~644 nm were observed for fs-TA spectra of 1. The proximity of these two bands to the position of the ground-state absorption peak suggests they are originated from ground-state bleaching (GSB). Furthermore, two excited-state absorption (ESA) bands centered at ~470 and ~520 nm at blue region, as well as a broad band after 660 nm at the red region were observed, which should be originated from its singlet state. With the delay of the time, the TA signal recovered into the ground state without no further spectral evolution, suggesting that no new excited state was formed. From the fitting result of the single-wavelength dynamics of 470 nm (Fig. 7b), we can see that the singlet state decays rapidly into the ground state with a time constant of 26.4 ps. Notably, no triplet state signal was observed in the evolution process of the singlet state, which is consistent with the uphill thermodynamic driving force (1.33 kcal/mol).
Discussion
In summary, a soluble and stable peri-hexacene derivative 1 was successfully synthesized and isolated in crystalline form. It exhibits large diradical character and moderate tetra radical character but is still reasonably stable as a result of kinetic blocking of the most reactive sites at the two zigzag edges with the bulky and electron-withdrawing 2,6-dichlorophenyl groups. Analysis shows that it displays a unique electronic structure with five localized aromatic sextets. It exhibits characteristic absorption of singlet open-shell diradicaloids in the NIR region and has a smaller singlet-triplet energy gap. It also shows a global aromatic character that is similar to those of peri-tetracene and peri-heptacene analogues. Notably, a long spin-lattice relaxation time of 16.7 ms and quantum phase memory time of 0.85 µs was realized in a 1 solution, suggesting its potential application in organic spintronics. On the other hand, it has a small energy gap, showing amphoteric redox behavior and magnetic activity. Our synthetic strategy and physical studies not only reveal the fundamental change of electronic structures and aromaticity of peri-acenes upon lateral extension but also provide some insights into the synthesis of larger-size peri-acene molecules or other zigzag-edged graphene-like molecules with significant diradical character.
Methods
Synthesis of 1
In a nitrogen-filled glove box, under argon atmosphere and in the dark, a 20 ml vial containing a magnetic stir bar was charged with 8 (50 mg, 0.018 mmol), tert-butoxide potassium (20.4 mg, 0.182 mmol), 18-crown-6 (48.0 mg, 0.182 mmol) and THF (10 mL). The resulting mixture was stirred at room temperature for 6 h, then p-chloranil (11.6 mg, 0.047 mmol) was added. The mixture was kept stirred for 20 min and then the solvent was removed under reduced pressure at room temperature. The resulting residue was purified by flash chromatography (silica gel was neutralized with Et3N, Hexane/DCM = 3:1) to give 1 (36.5 mg) as a cyan solid in 73% yield. Full experiment details can be found in the Supplementary Information.
Supplementary information
Source data
Acknowledgements
W. Z. acknowledges the financial support from National Natural Science Foundation of China (22175061), Excellent Youth Foundation of Hunan Scientific Committee (2022JJ10025), Key research and development projects of Hunan Province (2022GK2035) and Key research and development projects of Xiangtan City (ZX-ZD20221006). Q. J. thanks the construct program of applied characteristic discipline in Hunan Province, the High-level Talent Program from Hunan University of Science and Engineering and the Science and Technology Innovation Program of Hunan Province (2024RC3224) for financial support. G. L. acknowledges financial support from the National Natural Science Foundation of China (22205117), the Fundamental Research Funds for the Central Universities (63223054, 63233092, 63243139).
Author contributions
J. Z., X-J. F. and W. N. contributed equally to this work. W. Z., Q. J. and B. Z. conceived and designed the experiments. J. Z. and X. F. performed synthetic works, physical characterizations, and computational studies. Y. Y., Y. H., J. S., Y. X. and Z. Z. helped to perform part of synthetic works and physical characterizations. G. L. and W. N. performed the EPR measurements. X-N. F. and H. L. performed the excited-state dynamics study. J. Z., W. Z. and Q. J. wrote the manuscript with input from all authors. All authors contributed to the scientific discussion.
Peer review
Peer review information
Nature Communications thanks Igor Alabugin, who co-reviewed with Michael Commodore and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Data availability
All data needed to evaluate the conclusions of this study are available in the main text or Supplementary Information. The X-ray crystallographic data for the structures reported in this study have been deposited at the Cambridge Crystallographic Data Centre under deposition numbers CCDC 2337978 (1) and 2327038 (7). Copies of the data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/. Source Data are provided with this manuscript. All data are available from the corresponding author upon request. Source data are provided with this paper.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
These authors contributed equally: Jinji Zhang, Xiaojing Fang, Weiwei Niu.
Contributor Information
Baishu Zheng, Email: zbaishu@163.com.
Qing Jiang, Email: qjiang198@163.com.
Guangwu Li, Email: ligw@nankai.edu.cn.
Wangdong Zeng, Email: wangdong.zeng@hnust.edu.cn.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-55556-5.
References
- 1.Nakada, K., Fujita, M., Dresselhaus, G. & Dresselhaus, M. S. Edge state in graphene ribbons: nanometer size effect and edge shape dependence. Phys. Rev. B.54, 17954–17961 (1996). [DOI] [PubMed] [Google Scholar]
- 2.Son, Y. W., Cohen, M. L. & Louie, S. G. Half-metallic graphene nanoribbons. Nature444, 347–349 (2006). [DOI] [PubMed] [Google Scholar]
- 3.Kumar, S. et al. Electronic, transport, magnetic, and optical properties of graphene nanoribbons and their optical sensing applications: a comprehensive review. Luminescence38, 909–953 (2023). [DOI] [PubMed] [Google Scholar]
- 4.Mousavi, S. T., Badehian, H. A. & Gharbavi, K. Optical spectra of zigzag graphene nanoribbons: a first-principles study. Phys. Scr.97, 105803 (2022). [Google Scholar]
- 5.Wang, Z., Sun, M., Zhao, Y., Xiao, J. & Dai, X. The electronic and transport properties of the folded zigzag graphene nanoribbon. Surf. Interfaces5, 72–75 (2016). [Google Scholar]
- 6.Wang, X. S. et al. Unique magnetic and thermodynamic properties of a zigzag graphene nanoribbon. Phys. A.527, 121356 (2019). [Google Scholar]
- 7.Son, Y. W., Cohen, M. L. & Louie, S. G. Energy gaps in graphene nanoribbons. Phys. Rev. Lett.97, 216803 (2006). [DOI] [PubMed] [Google Scholar]
- 8.Kan, E. J., Li, Z., Yang, J. & Hou, J. Half-metallicity in edge-modified zigzag graphene nanoribbons. J. Am. Chem. Soc.130, 4224–4225 (2008). [DOI] [PubMed] [Google Scholar]
- 9.Zhang, D. B. & Wei, S. H. Inhomogeneous strain-induced half-metallicity in bent zigzag graphene nanoribbons. NPJ Comput. Mater.3, 32–37 (2017). [Google Scholar]
- 10.Shao, J., Paulus, B. & Tremblay, J. C. Local current analysis on defective zigzag graphene nanoribbons devices for biosensor material applications. J. Comput. Chem.42, 1475–1485 (2021). [DOI] [PubMed] [Google Scholar]
- 11.Martín-Martínez, F. J. et al. Inducing aromaticity patterns and tuning the electronic transport of zigzag graphene nanoribbons via edge design. J. Phys. Chem. C.117, 26371–26384 (2013). [Google Scholar]
- 12.Feng, J. et al. How size, edge shape, functional groups and embeddedness influence the electronic structure and partial optical properties of graphene nanoribbons. Phys. Chem. Chem. Phys.23, 20695–20701 (2021). [DOI] [PubMed] [Google Scholar]
- 13.Li, Z., Huang, B. & Duan, W. The half-metallicity of zigzag graphene nanoribbons with asymmetric edge terminations. J. Nanosci. Nanotechno.10, 5374–5378 (2010). [DOI] [PubMed] [Google Scholar]
- 14.Clar, E. Polycyclic hydrocarbons. Angew. Chem. Int. Ed. Engl.77, 875–876 (1964). [Google Scholar]
- 15.Clar, E. The aromatic sextet. Z. f.ür. Chem.13, 200–200 (1972). [Google Scholar]
- 16.Clar, E., Zander, M. 378. 1: 12-2: 3-10: 11-Tribenzoperylene. J. Chem. Soc., 1861–1865 https://pubs.rsc.org/en/content/articlelanding/1958/jr/jr9580001861 (1958).
- 17.Jiang, D. E., Sumpter, B. G. & Dai, S. First principles study of magnetism in nanographenes. J. Chem. Phys.127, 124703 (2007). [DOI] [PubMed] [Google Scholar]
- 18.Jiang, D. & Dai, S. Circumacenes versus periacenes: HOMO-LUMO gap and transition from nonmagnetic to magnetic ground state with size. Chem. Phys. Lett.466, 72–75 (2008). [Google Scholar]
- 19.Mullinax, J. W. et al. Heterogeneous CPU + GPU algorithm for variational two-electron reduced-density matrix-driven complete active-space self-consistent field theory. J. Chem. Theory Comput.15, 6164–6178 (2019). [DOI] [PubMed] [Google Scholar]
- 20.Konishi, A., Hirao, Y., Kurata, H. & Kubo, T. Investigating the edge state of graphene nanoribbons by a chemical approach: Synthesis and magnetic properties of zigzag-edged nanographene molecules. Solid. State Commun.175–176, 62–70 (2013). [Google Scholar]
- 21.Ajayakumar, M. R., Ma, J. & Feng, X. π-Extendedperi-acenes: recent progress in synthesis and characterization. Eur. J. Org. Chem.2022, 01428 (2022). [Google Scholar]
- 22.Tombros, N., Jozsa, C., Popinciuc, M., Jonkman, H. T. & van Wees, B. J. Electronic spin transport and spin precession in single graphene layers at room temperature. Nature448, 571–574 (2007). [DOI] [PubMed] [Google Scholar]
- 23.Biel, B., Triozon, F., Blase, X. & Roche, S. J. N. l. Chemically induced mobility gaps in graphene nanoribbons: a route for upscaling device performances. Nano. Lett.9, 2725–2729 (2009). [DOI] [PubMed] [Google Scholar]
- 24.Xiao, Y., Ye, Q., Liang, J., Yan, X. & Zhang, Y. Different noncollinear magnetizations on two edges of zigzag graphene nanoribbons. Chin. Phys. B.29, 127201 (2020). [Google Scholar]
- 25.Ajayakumar, M. R. et al. Toward full zigzag-edged nanographenes: peri-tetracene and its corresponding circumanthracene. J. Am. Chem. Soc.140, 6240–6244 (2018). [DOI] [PubMed] [Google Scholar]
- 26.Ni, Y. et al. A peri-tetracene diradicaloid: synthesis and properties. Angew. Chem. Int. Ed. Engl.57, 9697–9701 (2018). [DOI] [PubMed] [Google Scholar]
- 27.Ajayakumar, M. R. et al. Persistent peri-heptacene: synthesis and in situ characterization. Angew. Chem. Int. Ed. Engl.60, 13853–13858 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Rogers, C. et al. Closing the nanographene gap: surface-assisted synthesis of peripentacene from 6,6’-bipentacene precursors. Angew. Chem. Int. Ed. Engl.54, 15143–14146 (2015). [DOI] [PubMed] [Google Scholar]
- 29.Zhang, X., Li, J., Qu, H., Chi, C. & Wu, J. Fused bispentacenequinone and its unexpected Michael addition. Org. Lett.12, 3946–3949 (2010). [DOI] [PubMed] [Google Scholar]
- 30.Zophel, L. et al. Toward the peri-pentacene framework. Chem. Eur. J.19, 17821–17826 (2013). [DOI] [PubMed] [Google Scholar]
- 31.Konishi, A. et al. Synthesis and characterization of teranthene: a singlet biradical polycyclicaromatic hydrocarbon having Kekule´ structures. J. Am. Chem. Soc.132, 11021–11023 (2010). [DOI] [PubMed] [Google Scholar]
- 32.Kuriakose, F. et al. Design and synthesis of Kekulè and non-Kekulè diradicaloids via the radical periannulation strategy: the power of seven Clar’s sextets. J. Am. Chem. Soc.144, 23448–23464 (2022). [DOI] [PubMed] [Google Scholar]
- 33.Konishi, A. et al. Synthesis and characterization of quarteranthene: elucidating the characteristics of the edge state of graphene nanoribbons at the molecular level. J. Am. Chem. Soc.135, 1430–1437 (2013). [DOI] [PubMed] [Google Scholar]
- 34.Sun, Z. et al. Dibenzoheptazethrene isomers with different biradical characters: an exercise of Clar’s aromatic sextet rule in singlet biradicaloids. J. Am. Chem. Soc.135, 18229–18236 (2013). [DOI] [PubMed] [Google Scholar]
- 35.Jiang, Q., Tang, H., Peng, Y., Hu, Z. & Zeng, W. Helical polycyclic hydrocarbons with open-shell singlet ground states and ambipolar redox behaviors. Chem. Sci.15, 10519–10528 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Zeng, W. et al. Rylene ribbons with unusual diradical character. Chem2, 81–92 (2017). [Google Scholar]
- 37.Jiang, Q., Wei, H., Hou, X. & Chi, C. Circumpentacene with open‐shell singlet diradical character. Angew. Chem. Int. Ed.62, 06938 (2023). [DOI] [PubMed] [Google Scholar]
- 38.Shen, J. J. et al. A stable [4,3]peri‐acene diradicaloid: synthesis, structure, and electronic properties. Angew. Chem. Int. Ed.60, 4464–4469 (2020). [DOI] [PubMed] [Google Scholar]
- 39.Doehnert, D. & Koutecky, J. Occupation numbers of natural orbitals as a criterion for biradical character. Different kinds of biradicals. J. Am. Chem. Soc.102, 1789–1794 (1980). [Google Scholar]
- 40.Zeng, W., Qi, Q. & Wu, J. Cyclopenta-fused perylene: a new soluble, stable and functionalizable rylene building block. Sci. Bull.60, 1266–1271 (2015). [Google Scholar]
- 41.Shen, T. et al. Bis‐peri‐dinaphtho‐rylenes: facile synthesis via radical‐mediated coupling reactions and their distinctive electronic structures. Angew. Chem. Int. Ed.62, 11928 (2023). [DOI] [PubMed] [Google Scholar]
- 42.Zou, Y. et al. Peri-pentacene and peri-hexacene diradicaloids. J. Am. Chem. Soc.146, 27293–27298 (2024). [DOI] [PubMed] [Google Scholar]
- 43.Ostrowski, S. & Dobrowolski, J. C. What does the HOMA index really measure? RSC Adv.4, 44158–44161 (2014). [Google Scholar]
- 44.Chen, Z., Wannere, C. S., Corminboeuf, C., Puchta, R. & Schleyer, Pv. R. Nucleus-independent chemical shifts (NICS) as an aromaticity criterion. Chem. Rev.105, 3842–3888 (2005). [DOI] [PubMed] [Google Scholar]
- 45.Geuenich, D., Hess, K., Köhler, F. & Herges, R. Anisotropy of the induced current density (ACID), a general method to quantify and visualize electronic delocalization. Chem. Rev.105, 3758–3772 (2005). [DOI] [PubMed] [Google Scholar]
- 46.Klod, S. & Kleinpeter, E. Ab initio calculation of the anisotropy effect of multiple bonds and the ring current effect of arenes—application in conformational and configurational analysis. J. Chem. Soc. Perkin Trans.2, 1893–1898 (2001). [Google Scholar]
- 47.Lu, T. & Chen, F. Multiwfn: a multifunctional wavefunction analyzer. J. Comput. Chem.33, 580–592 (2011). [DOI] [PubMed] [Google Scholar]
- 48.Bleaney, B. & Bowers, K. D. Anomalous paramagnetism and exchange interaction in copper acetate. R. Soc. Lond. Ser. A214, 451–465 (1952). [Google Scholar]
- 49.Hu, Z. et al. Endohedral metallofullerene as molecular high spin qubit: diverse Rabi cycles in Gd2@C79N. J. Am. Chem. Soc.140, 1123–1130 (2018). [DOI] [PubMed] [Google Scholar]
- 50.Sato, K. et al. Polycationic high-spin states of one- and two-dimensional(diarylamino) benzenes, prototypical model units for purelyorganic ferromagnetic metals as studied by pulsed ESR/electron spin transient nutation spectroscopy. J. Am. Chem. Soc.119, 6607–6613 (1997). [Google Scholar]
- 51.Wang, Z. Y. et al. A stable triplet-ground-state conjugated diradical based on a diindenopyrazine skeleton. Angew. Chem. Int. Ed.60, 4594–4598 (2021). [DOI] [PubMed] [Google Scholar]
- 52.Zhu, Z. et al. Rational design of an air-stable, high-spin diradical with diazapyrene. Angew. Chem. Int. Ed.62, e202314900 (2023). [DOI] [PubMed] [Google Scholar]
- 53.Komeda, J. et al. Selective transition enhancement in a g-engineered diradical. Chem. Eur. J.30, e202400420 (2024). [DOI] [PubMed] [Google Scholar]
- 54.Lombardi, F. et al. Quantum units from the topological engineering of molecular graphenoids. Science366, 1107–1110 (2019). [DOI] [PubMed] [Google Scholar]
- 55.Lombardi, F. et al. Synthetic tuning of the quantum properties of open-shell radicaloids. Chem7, 1363–1378 (2021). [Google Scholar]
- 56.Wolfowicz, G. et al. Author correction: quantum guidelines for solid-state spin defects. Nat. Rev. Mater.6, 1191–1191 (2021). [Google Scholar]
- 57.Graham, M. J., Yu, C. J., Krzyaniak, M. D., Wasielewski, M. R. & Freedman, D. E. Synthetic approach to determine the effect of nuclear spin distance on electronic spin decoherence. J. Am. Chem. Soc.139, 3196–3201 (2017). [DOI] [PubMed] [Google Scholar]
- 58.Winpenny, R. E. P. Quantum information processing using molecular nanomagnets as qubits. Angew. Chem. Int. Ed.47, 7992–7994 (2008). [DOI] [PubMed] [Google Scholar]
- 59.Valenta, L. et al. Trimesityltriangulene: a persistent derivative of Clar’s hydrocarbon. Chem. Commun.58, 3019–3022 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Mayländer, M., Thielert, P., Quintes, T., Vargas Jentzsch, A. & Richert, S. Room temperature electron spin coherence in photogenerated molecular spin qubit candidates. J. Am. Chem. Soc.145, 14064–14069 (2023). [DOI] [PubMed] [Google Scholar]
- 61.Di Motta, S., Negri, F., Fazzi, D., Castiglioni, C. & Canesi, E. V. Biradicaloid and polyenic character of quinoidal oligothiophenes revealed by the presence of a low-lying double-exciton state. J. Phys. Chem. Lett.1, 3334–3339 (2010). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
All data needed to evaluate the conclusions of this study are available in the main text or Supplementary Information. The X-ray crystallographic data for the structures reported in this study have been deposited at the Cambridge Crystallographic Data Centre under deposition numbers CCDC 2337978 (1) and 2327038 (7). Copies of the data can be obtained free of charge via https://www.ccdc.cam.ac.uk/structures/. Source Data are provided with this manuscript. All data are available from the corresponding author upon request. Source data are provided with this paper.