Robust free-base and metalated corrole radicals with reduction-induced emission

Pengfei Li Chulin Qu Fan Wu Hu Gao Chengyan Zhao Yue Zhao Zhen Shen

Citation:  Pengfei Li, Chulin Qu, Fan Wu, Hu Gao, Chengyan Zhao, Yue Zhao, Zhen Shen. Robust free-base and metalated corrole radicals with reduction-induced emission[J]. Chinese Chemical Letters, 2025, 36(2): 110292. doi: 10.1016/j.cclet.2024.110292 shu

Robust free-base and metalated corrole radicals with reduction-induced emission

English

  • Organic radicals have attracted considerable interest in various application fields because of their distinct redox and magnetic properties offered by the unpaired electrons [1-3]. However, the active nature of radicals makes their synthesis and manipulation very challenging. Porphyrinoids are typically aromatic macrocycles with [4n + 2] π-conjugation systems (e.g., 18 π tetrapyrrolic porphyrin, corrole, and phthalocyanine) (Fig. 1a). The redox activity of porphyrinoids enables the generation of [4n + 1]/[4n + 3] π radical ions via one-electron redox reactions. The porphyrinoid radicals could be stabilized via spin delocalization on their large π-conjugated systems [4], but most of them only exist in their coordinated forms with the insertion of boron [5,6], phosphorus [7], and various metal ions [8-10]. The absence of free amines in coordinated porphyrins allows for easier radical stabilization. The direct redox reactions on free-base porphyrins usually involve complicated deprotonation-protonation processes and other side reactions [11-13], making the identification and isolation of desired radicals extremely difficult. Free-base porphyrinoid radicals are assumed to be flexible, sensitive to outer stimulus, and easy to be functionalized via coordination, but reports of such molecules are still rare [14-18].

    Figure 1

    Figure 1.  (a) Structures of free-base aromatic porphyrin, corrole, and phthalocyanine. (b) Structure of the free-base benzocorrole radical and schematic of its coordination performance and reduction-induced luminescence.

    Corroles are one-carbon-short porphyrins that typically have three inner NHs in their free-base forms. The electron-rich nature and squeezed inner cavity of corrole provide non-innocent coordination properties for sustaining radicals [19]. When coordinated with some metal ions such as Cu(Ⅱ), Ni(Ⅱ), and Pt(Ⅳ) [20-23], corroles could undergo ring-oxidation to function as dianion radical ligands. The unique coordination performances of corroles lead to a wide range of diversified applications including biomedical theranostics [24-26], catalysis [27-33], and spintronics [34-37]. Benzo-fusion is a highly effective method for raising the energy of the π HOMO of corrole and regulating the d-π interaction between the corrole ligand and the central ion. Paolesse and co-workers reported the first free-base tetrabenzocorrole via a cross-coupling method, which is a 3NH closed-shell molecule bearing eight peripheral -CO2CH3 groups. In addition, they have effectively synthesized a variety of benzocorrole complexes with the insertion of Cu, Co, Sn and Al ions [38,39]. We also successfully synthesized Cu(Ⅱ) and Ni(Ⅱ) benzocorrole complexes that were metalated using one-pot procedures. These complexes demonstrated intriguing capabilities in sustaining stable π-radicals and have been employed in the study of spin modulation [34,36], aromatic-antiaromatic conversion [20], and photothermal therapy [25]. Although ligand-noninnocence was commonly found in metallocorroles, the synthesis of metal-free corrole radicals remained difficult, and Bröring et al. reported the first and only such free monoradical that was stabilized by β-chlorination [15]. It is still desirable to develop stable free-base corrole radicals for controllable coordination to create a variety of radical complexes with unique features.

    In this work, we report a free-base benzocorrole (BC) that is highly stable in its neutral radical state at ambient conditions (Fig. 1b). BC consists of elements C, N, and H only, and contains two inner NHs. As a radical ligand, BC demonstrates easy coordination with both divalent and trivalent metal ions (Zn(Ⅱ), Pd(Ⅱ), and Ga(Ⅲ)), and prefers to stabilize the complexes in their radical states. Red fluorescence or near-infrared phosphorescence could be activated on BC and its complexes via ligand reduction to aromatic anions under various treatments.

    The benzo-fused BC was prepared using the retro-Diels-Alder approach [40]. The bicyclo[2.2.2]octadiene-fused 3H-corrole 1 was newly synthesized as a diastereomeric mixture via the direct acid-catalyzed condensation between 4, 7-dihydro-4, 7-ethano-2H-isoindole and 3, 5-di-tert-butylbenzaldehyde in a 3:1 (v: v) methanol-water mixture solvent, followed by the oxidative cyclization by p-chloranil in chloroform. Heating solid 1 at 250 ℃ in vacuo eliminated its ethylene bridges and afforded radical BC spontaneously (Fig. 2a). The high-resolution ESI mass spectrometry for BC provided the exact mass at m/z = 1061.6427 consistent with a 2H formula C77H81N4 (m/z 1061.6456) (Fig. S2 in Supporting information). Compound 1 was 1H NMR active, indicative of its closed-shell character. In contrast, BC was nearly 1H NMR silent, and only broad signals of tert-butyl protons of BC could be found (Fig. S15 in Supporting information). The electron spin resonance (ESR) measurement of BC provided an intense and sharp peak at g = 2.0029, clearly indicating its free organic radical character (Fig. 2b). The toluene solution of BC stored at ambient conditions exhibited unchanged absorption during 30-day monitoring, indicative of its intriguing stability (Fig. S25 in Supporting information).

    Figure 2

    Figure 2.  (a) Synthesis of BC: (i) CH3OH, H2O, HCl, r.t.; (Ⅱ) CHCl3, p-chloranil, reflux; (Ⅲ) 250 ℃, in vacuo. (b) ESR spectrum of BC in CH2Cl2. (c) Top and (d) side view of the single-crystal structure of BC with thermal ellipsoids at 30% probability. (e) Packing view of BC with space-fill representation.

    Single crystals of BC were obtained from the slow diffusion of methanol into its chloroform solution and analyzed by X-ray diffraction (Figs. 2c-e). The structure of BC contains only two NH moieties with trans arrangement. Compared to 3H-corroles [41-43], the benzocorrole plane of BC demonstrates higher planarity with a mean plane deviation of only 0.056 Å due to smaller steric repulsion within the inner cavity. The amino-type pyrroles of BC have wider Ca-N-Ca angles (111.74° and 112.45°) than the imino-type pyrroles (109.22° and 109.74°), but imino-type pyrroles have more significant long-and-short alternation in Cα-N bond lengths (Table S3 in Supporting information). The three meso-aryl groups are nearly perpendicular to the corrole plane. The directly linked Cα-Cα bond length is 1.395(6) Å, which is significantly shorter than those of the reported free-base corroles ranging between 1.41 Å and 1.44 Å [15,41-43]. There is no close intermolecular stacking observed owing to the bulky 3, 5-tert-butyl-phenyl groups.

    Cyclic voltammetry and differential pulse voltammetry of BC reveal its reversible redox properties with oxidation peaks at Eox1 = −0.12 V, Eox2 = 0.71 V, and a reduction peak at Ered1 = −0.65 V (vs. Fc/Fc+) (Fig. 3a). There is no protonation-deprotonation observed during the redox processes. The gap between Eox1 and Ered1 is only 0.53 V, which is typical for a free π-radical and significantly lower than the approximately 2 V gaps found in 18 π porphyrins and corroles [44,45].

    Figure 3

    Figure 3.  (a) Electrochemical measurements of BC in CH2Cl2. (b) Absorption and fluorescence spectra of 1 in CH2Cl2ex = 415 nm), and absorption spectra of BC in various solvents. (c) Images of BC solutions displaying the color change upon protonation and the fluorescence after reduction. (d) Absorption spectral change of BC upon adding HCl in the CH2Cl2 solution. (e) Absorption spectral changes of BC during oxidation and reduction recorded by spectroelectrochemistry. (f) Absorption, fluorescence (λex = 458 nm), and MCD spectra of BC in DMF (10 µmol/L, 3 mL) upon adding DBU (0.5 µL).

    The absorption spectrum of corrole 1 displays a sharp Soret band at 415 nm in CH2Cl2, with its emission spectrum showing a fluorescence band peaking at 654 nm (Fig. 3b). Comparatively, BC is fluorescence-silent in CH2Cl2, and exhibits an intense red-shifted absorption band at 454 nm and several weak broad bands in the range of 500–750 nm. BC was loaded into water-soluble nanoparticles (BC-NPs) by mixing with amphiphilic polymer DSPE-PEG, which exhibited near identical absorption in water as in CH2Cl2 and toluene. The solution color of BC in CH2Cl2 was light green and rapidly changed to pink upon HCl addition (Fig. 3c), and by adding TEA, the original color was reversibly recovered. Stepwise titration of BC by adding HCl incrementally revealed spectral changes with unchanged isobestic points, thus a one-step diprotonation process forming [H]2+ is purposed (Fig. S16 in Supporting information). The optimized structure of [H]2+ is highly saddle-distorted owing to the steric effect of the four inner pyrrole N-H protons. Such distortion is commonly observed in other tetrapyrrolic macrocycles that are fully protonated [46]. The absorbance maxima of [H]2+ shifts to 519 nm (Fig. 3d). The notable redshift is verified by TD-DFT calculations. The detailed compositions of the electronic transitions of [H]2+ are presented in Table S8 (Supporting information). Similar protonation-induced bathochromic shifts of absorption bands have also been observed in aromatic porphyrins and phthalocyanines [47].

    The redox behaviors of BC associated with absorption changes were further investigated using thin-layer spectroelectrochemistry (Fig. 3e). BC was gradually oxidized to its cation BC+ and reduced to anion BC by shifting the applying potential positively and negatively, respectively. Isobestic points were observed on the spectral change, indicating no side reactions occurred during the redox processes. The band at 454 nm gradually decreases from BC to BC+, corresponding to the conjugation change from [4n + 1] to [4n] π. In contrast, aromatic characteristic Soret bands at 456 and 465 nm, as well as the Q band at 641 nm, intensify from BC to BC, indicating the formation of a [4n + 2] π system after reduction.

    After adding trace 1, 8-diazabicyclo[5.4.0]undec-7-ene (DBU) into the DMF solution of BC, the same spectral change as the reduction in spectroelectrochemistry was observed, indicating the formation of anion BC (Fig. 3f). The Soret band splitting of BC reflects the x and y polarizations (Table S7 in Supporting information) [48]. The chemical reduction also occurred by using THF, CH3CN and DMSO as solvents or hydrazine, NaBH4 and SnCl2 as reductant (Figs. S18–S24 in Supporting information). The DMF solution of BC exhibits red fluorescence, with a maximum at 648 nm and a shoulder at 707 nm in its emission spectrum. DBU has been used as a reductant in a few reactions, including the reduction of a nickel corrole radical [21,49-52]. We assume that there is a direct electron transfer from DBU to the corrole radical. forming a DBU radical that is easily quenched (Fig. S13 in Supporting information). The Hückel aromaticity of BC was identified by magnetic circular dichroism (MCD) spectroscopy [53]. The coupled intense MCD Faraday B terms with a + / - sign at 645 and 628 nm in ascending energy are assigned to the Q bands of 18 π aromatic porphyrinoids, corresponding to an angular orbital momentum change of ±9. The weak MCD signals with a – / + sign at 466 and 455 nm are ascribed to the angular orbital momentum change of ±1.

    Density functional theory (DFT) calculations (B3LYP) were performed on BC and its redox ions (tert-butyl groups omitted). Two common 18 π aromatic free-base porphyrinoids, 5, 10, 15-triphenylcorrole (TPC) and 5, 10, 15, 20-tetraphenylporphyrin (TPP) were also calculated for comparison. The bond lengths of DFT-optimized BC show good consistency with those obtained from the single-crystal structure (Fig. S28 in Supporting information). Free-base porphyrinoids exhibit tautomerism with core-hydrogen alteration [54,55], and the trans-NH state of BC is preferred with lower energy than the possible cis-NH tautomers (Fig. S44 in Supporting information). The spin density of BC exhibits very small distribution on peripheral-benzo- and meso-carbons, but large coefficients on the pyrrolic carbons (Fig. 4a). The molecular orbital (MO) energy level diagram indicates that BC exhibits better one-electron donating and accepting capabilities in comparison to TPC and TPP (Fig. 4a). Both the a2-type lowest singly unoccupied MO (SUMO) and highest singly occupied MO (SOMO) of BC closely resemble the distribution of its spin density (Fig. S36 in Supporting information) [56]. Notably, the SUMO of BC (−3.28 eV) is more than 1 eV lower than the lowest (doubly) unoccupied MO (LUMO) of TPC (−2.12 eV) and TPP (−2.21 eV), indicating the strong electron-accepting ability of BC that facilitates electron pairing. Thus, the electron-transfer process could easily occur on BC upon adding the strong Lewis base DBU. The SOMO-SUMO gap of BC (1.08 eV), which is much smaller than the highest (doubly) occupied MO (HOMO)-LUMO gaps of TPC and TPP. Upon a one-electron reduction, the HOMO-LUMO gap of BC is determined to be 2.44 eV, a value that closely resembles that of TPC (2.47 eV) (Fig. 4b). The solvent effect on the stabilization of the anion of BC was further evaluated by employing SMD solvent models in the calculations. The HOMO and LUMO energy levels, as well as the total energy of BC in DMF, are significantly lower than those in toluene and gas phase, whereas their HOMO-LOMO gaps remain consistent (Fig. S43 in Supporting information). The time-dependent DFT (TD-DFT) calculations (B3LYP) nicely reproduce the absorption spectra of BC, BC+, and BC, as well as the protonated [H2BC]2+ (Figs. S49-S51, S56 in Supporting information). The calculated Q and Soret bands of BC completely originate from its frontier four orbitals with ML = ±4 and ±5 (Table S7). The S1 state of BC was optimized, predicting a fluorescence band at 647.5 nm corresponding to the S1 → S0 transition, which exactly matched the experimental finding at 648 nm (Fig. 4c). The electron and hole distributions on the S1 state of BC were analyzed by Multiwfn, revealing no spatial separation of the π-excitation (Fig. 4d) [47,57]. The Sr index, which quantifies the degree of electron-hole overlap, exhibits a value of 0.94 within the range of 0 to 1, suggesting the S1 state of BC is a locally-excited (LE) state.

    Figure 4

    Figure 4.  (a, b) Frontier molecular orbital diagrams of BC, TPC, TPP, and BC, with the spin density map of BC. (c) TD-DFT simulated emission spectra of BC. (d) The electron (blue) and hole (yellow) distribution of the S1 excited state on BC. (e) Molecular electrostatic potential and ICSS maps (1 Å above the XY plane) of BC, TPC, and TPP.

    DFT calculations were also performed on a variety of inner-3NH corroles with the same phenyl substituents at meso positions, including TPC, dibenzo-fused DBC-3H, tetrabenzo-fused TBC-3H, and octacarboxymethyl-substituted tetrabenzo-fused OCTBC-3H, to address the reason for the energy level and vertical ionization potential due to the π-extension (Table S5 in Supporting information). This indicates that it is the most unstable among the four and may tends to lose one electron and one proton under ambient conditions [13], resulting in the generation of a neutral radical.

    Electrostatic potential (ESP) and Iso-chemical shielding surface (ICSS) calculations [58,59] were conducted on BC, TPC, and TPP, using their optimized structures (Fig. 4e). The ESP maps show that the central coordination atmosphere of BC is closer to that of TPP rather than TPC. BC has a high concentration of negative charges (blue) surrounding its two imine N atoms, indicating its Lewis base nature, which allows for further protonation or coordination. In contrast, TPC has significant positive charge (red) on the amine N—H, which is located opposite the negatively charged imine N atom, indicating that TPC is more acidic than TPP and BC [60]. The ICSS maps clearly show that both TPC and TPP are typically aromatic with the shielded region depicted in red and yellow. The nonaromatic character of BC is revealed by the green-cyan region with close-to-zero nucleus-independent chemical shifts (NICS) values inside the corrole ring [61]. The one-electron reduced anion BC changes to aromatic, exhibiting an inner shielding area similar to neutral TPC and TPP. The NICS(1) values of BC and TPP at the molecular center are −11.6, −11.3, and −13.3 ppm, respectively. The induced ring currents calculated by anisotropy of the induced current density are clockwise for BC+, and anticlockwise for BC, but interrupted on the meso carbons of BC (Fig. S57 in Supporting information) [62].

    The zinc, gallium, and palladium complexes of BC were synthesized (Fig. 5a), and all were identified as π-radicals by their sharp single-line ESR signals and spin density maps with near-zero density on central metals (Figs. S28 and S44 in Supporting information). The direct insertion of Zn(Ⅱ) into 3H-corrole was difficult, which easily led to ring degradation [21,63]. Treating BC with zinc acetate in pyridine under reflux afforded stable Zn(Ⅱ) complex Zn-BC in high yield. Ga(Ⅲ) corroles were typically fluorescent molecules with axial pyridine coordination and have shown good tumor imaging and therapeutic performance [64-66]. Intriguingly, treating BC with GaCl3 resulted in the formation of non-emitting, chloride-ligated radical Ga-BC, as evidenced by the single-crystal X-ray diffraction. Such radical could only be obtained by treating common aromatic gallium corroles with the additional chemical oxidant N(4-BrC6H4)3SbCl6 [67]. The five-coordinate Ga(Ⅲ) center in Ga-BC has a distorted square pyramidal geometry. The Ga(Ⅲ) ion is 0.487 Å above the 4 N plane, and the Ga-Cl length is 2.2208(17)Å. By adding pyridine incrementally into the toluene solution of Ga-BC, the fluorescence appears and gradually enhances at 637 and 700 nm (Fig. 5b), corresponding to its ligand reduction to aromatic anion. However, the reduction was not observed for BC, Zn-BC, and Pd-BC in pure pyridine. The mass spectrum shows that the axial chloride of Ga-BC could be substituted by treating it with excess pyridine (Fig. S12 in Supporting information), which may facilitate ligand reduction. The palladium insertion of 3H-corrole could afford a closed-shell anionic complex accompanied by a counterion [68]. A neutral palladium radical complex Pd-BC was synthesized for the first time via the reflux of BC and Pd(OAc)2 in pyridine. Unlike Ga-BC, the single-crystal structure of Pd-BC reveals no axial ligand and the Pd(Ⅱ) ion perfectly lies on the highly planar corrole plane. After adding DBU into the DMF solution of Pd-BC, the appearance of sharp B and Q bands indicates the formation of anion Pd-BC (Fig. 5c). Interestingly, Pd-BC displays near-infrared phosphorescence in the anoxic solution at room temperature owing to the promoted intersystem crossing (ISC) to triplet by palladium. The long-lived emission shows a peak at 930 nm with a microsecond-level lifetime (9.82 µs) and extends beyond 1100 nm, exhibiting a "mega" shift of 306 nm relative to the lowest energy Q band. The quantum yields and lifetimes of all reduced compounds are listed in Table S4 (Supporting information). The electrochemical measurements were performed on the complexes in CH2Cl2 (Fig. S31 in Supporting information). The difference in potential between Eox1 and Ered1 is 0.49 V for Zn-BC, 0.59 V for Ga-BC, and 0.46 V for Pd-BC. These small gaps indicate the radical nature of the complexes.

    Figure 5

    Figure 5.  (a) Synthesis and chemical structures of the radical complexes. (i) Zn-BC: Zn(OAc)2·2H2O, pyridine, reflux; (Ⅱ) Ga-BC: GaCl3, pyridine, reflux; (Ⅲ) Pd-BC: Pd(OAc)2, pyridine, reflux. Single-crystal structures of Ga-BC and Pd-BC are presented (thermal ellipsoids at 30% probability) with central coordination structures highlighted. (b) Fluorescence spectra of Ga-BC in toluene as a function of the volume proportion of the added pyridine (λex = 467 nm). (c) Absorption and emission spectra, and time-resolved photoluminescence (PL) decay curve of Pd-BC in deaerated DMF upon adding DBU (λex = 624 nm).

    To get a better understanding of the room-temperature phosphorescence caused by Pd coordination, the excited-state energy gaps, and spin-orbit coupling matrix elements (SOCME) between singlet state S1 and triplet states Tn were further calculated for BC and Pd-BC (B3LYP) (Fig. 6) [69,70]. BC exhibits four triplet states, and Pd-BC exhibits three triplet states that are located below their respective S1 state. The T4 state of Pd-BC has nearly the same energy as its S1 state. BC and Pd-BC have large S1-T1 gaps (ΔES1-T1) of 0.58 eV and 0.60 eV, respectively, while their T3 and T4 states have small S1-Tn gaps (< 0.3 eV) that facilitate the ISC processes. All the S1-Tn (n = 1–4) SOCMEs of BC are smaller than 0.65 cm−1. For Pd-BC, although the S1-T1 and S1-T4 SOCMEs are minor (0.17 and 0.13 cm−1, respectively), the S1-T2 and S1-T3 SOCMEs exhibit significant values of 24.44 and13.83 cm−1, respectively. The ISC process is supposed to efficiently take place in the S1→T3 channel of Pd-BC due to the narrow energy gap (ΔES1-T3 = 0.17 eV) and large SOCME. The higher excited triplet states Tn (n > 1) can undergo fast internal conversion (IC) to T1, subsequently deactivate to the ground state S0 and emit phosphorescence [71].

    Figure 6

    Figure 6.  Schematic diagram of the calculated energy gaps, spin-orbital coupling constants between S1 and Tn (n = 1–4), and proposed energy transfer processes for fluorescence (Fluo.) and phosphorescence (Phos.).

    In summary, we have prepared a free-base benzocorrole radical BC and its various radical complexes, which were fully studied by crystallographic analysis, spectroscopic measurements, and theoretical calculations. The protonation at the inner imines of BC largely red-shifted its main absorption band. The fluorescence turn-on response of BC was discovered under the mild reducing atmosphere, corresponding to a reductive conversion to its aromatic anion. The Zn(Ⅱ), Pd(Ⅱ), and Ga(Ⅲ) complexes of BC are all stable radicals and readily reduced to their emissive anions in a manner similar to BC. Ga-BC was ligated by an axial chloride, which could be replaced by pyridine accompanied by enhanced fluorescence. The anion of Pd-BC exhibits NIR phosphorescence beyond 900 nm at room temperature. The reduction-driven luminescence of radicals offers their potential applications in responsive imaging. This radical ligand will be used to produce new complexes with tunable redox, magnetic and emissive properties, thus broadening the application of functional radical materials.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Pengfei Li: Writing – original draft, Validation, Methodology, Investigation, Formal analysis, Data curation, Conceptualization. Chulin Qu: Writing – review & editing, Investigation, Formal analysis, Data curation. Fan Wu: Writing – review & editing, Writing – original draft, Methodology, Funding acquisition, Conceptualization. Hu Gao: Methodology, Investigation. Chengyan Zhao: Investigation, Formal analysis, Data curation. Yue Zhao: Visualization, Software, Resources, Methodology, Data curation. Zhen Shen: Writing – review & editing, Resources, Project administration, Funding acquisition, Conceptualization.

    This work was supported by the National Natural Science Foundation of China (Nos. 22271140, 22071103, and 22371118). The theoretical calculations are performed using the supercomputing resources at the High-Performance Computing Center of Nanjing University.

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.110292.


    1. [1]

      Z.X. Chen, Y. Li, F. Huang, Chem 7 (2021) 288–332. doi: 10.1016/j.chempr.2020.09.024

    2. [2]

      L. Ji, J. Shi, J. Wei, et al., Adv. Mater. 32 (2020) 1908015. doi: 10.1002/adma.201908015

    3. [3]

      I. Ratera, J. Veciana, Chem. Soc. Rev. 41 (2012) 303–349. doi: 10.1039/C1CS15165G

    4. [4]

      D. Shimizu, A. Osuka, Chem. Sci. 9 (2018) 1408–1423. doi: 10.1039/c7sc05210c

    5. [5]

      D. Shimizu, J. Oh, K. Furukawa, et al., Angew. Chem. Int. Ed. 54 (2015) 6613–6617. doi: 10.1002/anie.201501592

    6. [6]

      W. Deng, Y. Liu, D. Shimizu, et al., Chem. Eur. J. 29 (2023) e202203484. doi: 10.1002/chem.202203484

    7. [7]

      T. Yoshida, W. Zhou, T. Furuyama, et al., J. Am. Chem. Soc. 137 (2015) 9258–9261. doi: 10.1021/jacs.5b05781

    8. [8]

      D. Shimizu, K. Fujimoto, A. Osuka, Angew. Chem. Int. Ed. 57 (2018) 9434–9438. doi: 10.1002/anie.201805385

    9. [9]

      N. Fukui, W. Cha, D. Shimizu, et al., Chem. Sci. 8 (2017) 189–199. doi: 10.1039/C6SC02721K

    10. [10]

      X.-S. Ke, Y. Hong, P. Tu, et al., J. Am. Chem. Soc. 139 (2017) 15232–15238. doi: 10.1021/jacs.7b09167

    11. [11]

      J. Shen, J. Shao, Ou Z, et al., Inorg. Chem. 45 (2006) 2251–2265. doi: 10.1021/ic051729h

    12. [12]

      Y. Yamamoto, A. Yamamoto, Shin-ya Furuta, et al., J. Am. Chem. Soc. 127 (2005) 14540–14541. doi: 10.1021/ja052842+

    13. [13]

      Y. Yamamoto, Y. Hirata, M. Kodama, et al., J. Am. Chem. Soc. 132 (2010) 12627–12638. doi: 10.1021/ja102817a

    14. [14]

      D. Shimizu, J. Oh, K. Furukawa, et al., J. Am. Chem. Soc. 137 (2015) 15584–15594. doi: 10.1021/jacs.5b11223

    15. [15]

      P. Schweyen, K. Brandhorst, R. Wicht, et al., Angew. Chem. Int. Ed. 54 (2015) 8213–8216. doi: 10.1002/anie.201503624

    16. [16]

      S. Hiroto, H. Shinokubo, A. Osuka, J. Am. Chem. Soc. 128 (2006) 6568–6569. doi: 10.1021/ja061621g

    17. [17]

      S. Ooi, D. Shimizu, K. Furukawa, et al., Angew. Chem. Int. Ed. 57 (2018) 14916–14920. doi: 10.1002/anie.201810200

    18. [18]

      S. Ooi, B. Adinarayana, D. Shimizu, et al., Angew. Chem. Int. Ed. 59 (2020) 9423–9427. doi: 10.1002/anie.202002976

    19. [19]

      A. Ghosh, Chem. Rev. 117 (2017) 3798–3881. doi: 10.1021/acs.chemrev.6b00590

    20. [20]

      H. Gao, F. Wu, Y. Zhao, et al., J. Am. Chem. Soc. 144 (2022) 3458–3467. doi: 10.1021/jacs.1c11716

    21. [21]

      M.L. Naitana, W.R. Osterloh, L. Di Zazzo, et al., Inorg. Chem. 61 (2022) 17790–17803. doi: 10.1021/acs.inorgchem.2c03099

    22. [22]

      C.M. Lemon, M. Huynh, A.G. Maher, et al., Angew. Chem. Int. Ed. 55 (2016) 2176–2180. doi: 10.1002/anie.201509099

    23. [23]

      A.B. Alemayehu, H. Vazquez-Lima, C.M. Beavers, et al., Chem. Commun. 50 (2014) 11093–11096. doi: 10.1039/C4CC02548B

    24. [24]

      R.D. Teo, J.Y. Hwang, J. Termini, et al., Chem. Rev. 117 (2017) 2711–2729. doi: 10.1021/acs.chemrev.6b00400

    25. [25]

      H. Gao, X. Zhi, F. Wu, et al., Angew. Chem. Int. Ed. 62 (2023) e202309208. doi: 10.1002/anie.202309208

    26. [26]

      L. Shi, H.Y. Liu, L.P. Si, et al., Chin. Chem. Lett. 21 (2010) 373–375. doi: 10.1016/j.cclet.2009.11.027

    27. [27]

      H. Shi, R. Liang, D.L. Phillips, et al., J. Am. Chem. Soc. 144 (2022) 7588–7593. doi: 10.1021/jacs.2c02506

    28. [28]

      Q.-C. Chen, S. Fite, N. Fridman, et al., ACS Catal. 12 (2022) 4310–4317. doi: 10.1021/acscatal.1c05243

    29. [29]

      N.C. Ng, M.H.R. Mahmood, H.Y. Liu, et al., Chin. Chem. Lett. 25 (2014) 571–574. doi: 10.1016/j.cclet.2014.01.027

    30. [30]

      L. Yu, Q. Wang, L. Dai, et al., Chin. Chem. Lett. 24 (2013) 447–449. doi: 10.1016/j.cclet.2013.03.029

    31. [31]

      C. Wu, X. Li, M. Shao, et al., Chin. Chem. Lett. 33 (2022) 4559–4562. doi: 10.1016/j.cclet.2022.01.065

    32. [32]

      Q. Zhang, Y. Wang, Y. Wang, et al., Chin. Chem. Lett. 32 (2021) 3807–3810. doi: 10.1016/j.cclet.2021.04.048

    33. [33]

      Y. Zhang, K. Ren, L. Wang, et al., Chin. Chem. Lett. 33 (2022) 33–60. doi: 10.1016/j.cclet.2021.06.013

    34. [34]

      F. Wu, J. Liu, P. Mishra, et al., Nat. Commun. 6 (2015) 7547. doi: 10.1038/ncomms8547

    35. [35]

      J. Xu, L. Zhu, H. Gao, et al., Angew. Chem. Int. Ed. 60 (2021) 11702–11706. doi: 10.1002/anie.202016674

    36. [36]

      F. Wu, H. Gao, Y. Zhao, et al., Chin. Chem. Lett. 34 (2023) 107994. doi: 10.1016/j.cclet.2022.107994

    37. [37]

      C. Yu, H. Xiao, H. Lu, et al., Chin. Chem. Lett. 35 (2024) 108883. doi: 10.1016/j.cclet.2023.108883

    38. [38]

      G. Pomarico, S. Nardis, R. Paolesse, et al., J. Org. Chem. 76 (2011) 3765–3773. doi: 10.1021/jo200026u

    39. [39]

      G. Pomarico, S. Nardis, M.L. Naitana, et al., Inorg. Chem. 52 (2013) 4061–4070. doi: 10.1021/ic400162y

    40. [40]

      S. Ito, T. Murashima, N. Ono, et al., Chem. Commun. (1998) 1661–1662. doi: 10.1039/A803656J

    41. [41]

      Z. Gross, N. Galili, L. Simkhovich, et al., Org. Lett. 1 (1999) 599–602. doi: 10.1021/ol990739h

    42. [42]

      T. Ding, J.D. Harvey, C.J. Ziegler, J. Porphyr. Phthalocyanines 09 (2005) 22–27. doi: 10.1142/S1088424605000058

    43. [43]

      J. Capar, J. Conradie, C.M. Beavers, et al., J. Phys. Chem. 119 (2015) 3452–3457. doi: 10.1021/jp511188c

    44. [44]

      Y. Fang, Z. Ou, K.M. Kadish, Chem. Rev. 117 (2017) 3377–3419. doi: 10.1021/acs.chemrev.6b00546

    45. [45]

      Y. Fang, M.O. Senge, E. Van Caemelbecke, et al., Inorg. Chem. 53 (2014) 10772–10778. doi: 10.1021/ic502162p

    46. [46]

      J. Gaŕate-Morales, F. Tham, C. Reed, et al., Inorg. Chem. 46 (2007) 1514–1516. doi: 10.1021/ic062213g

    47. [47]

      S. Fukuzumi, T. Honda, T. Kojima, Coord. Chem. Rev. 256 (2012) 2488–2502. doi: 10.1016/j.ccr.2012.01.011

    48. [48]

      C.M. Lemon, R.L. Halbach, M. Huynh, et al., Inorg. Chem. 54 (2015) 2713–2725. doi: 10.1021/ic502860g

    49. [49]

      T. Naito, A. saito, M. Ueda, et al., Heterocycles 65 (2005) 1857–1869. doi: 10.3987/COM-05-10426

    50. [50]

      D.Y. Jeong, D.S. Lee, H.L. Lee, et al., ACS Catal. 12 (2022) 6047–6059. doi: 10.1021/acscatal.2c00763

    51. [51]

      S. Xu, F. Yang, H. Fan, et al., New J. Chem. 46 (2022) 9994–9998. doi: 10.1039/d2nj00341d

    52. [52]

      D. Shen, T. Ren, Z. Luo, et al., Org. Biomol. Chem. 21 (2023) 4955–4961. doi: 10.1039/d3ob00663h

    53. [53]

      J. Mack, M.J. Stillman, N. Kobayashi, Coord. Chem. Rev. 251 (2007) 429–453. doi: 10.1016/j.ccr.2006.05.011

    54. [54]

      J. Helaja, M. Stapelbroek-Mo¨llmann, I. Kilpela¨inen, et al., J. Org. Chem. 65 (2000) 3700–3707. doi: 10.1021/jo9918955

    55. [55]

      A. Mahammed, J.J. Weaver, H.B. Gray, et al., Tetrahedron Lett. 44 (2003) 2077–2079. doi: 10.1016/S0040-4039(03)00174-6

    56. [56]

      A. Ghosh, T. Wondimagegn, A.B.J. Parusel, J. Am. Chem. Soc. 122 (2000) 5100–5104. doi: 10.1021/ja9943243

    57. [57]

      Z. Liu, T. Lu, Q. Chen, Carbon 165 (2020) 461–467. doi: 10.1016/j.carbon.2020.05.023

    58. [58]

      S. Klod, E. Kleinpeter, J. Chem. Soc., Perkin Trans. 2. 10 (2001) 1893–1898. doi: 10.1039/B009809O

    59. [59]

      T. Lu, F. Chen, J. Comput. Chem. 33 (2012) 580–592. doi: 10.1002/jcc.22885

    60. [60]

      A. Mahammed, J.J. Weaver, H.B. Gray, et al., Tetrahedron Lett. 44 (2003) 2077–2079. doi: 10.1016/S0040-4039(03)00174-6

    61. [61]

      Z. Chen, C. Wannere, C. Corminboeuf, et al., Chem. Rev. 105 (2005) 3842–3888. doi: 10.1021/cr030088+

    62. [62]

      D. Geuenich, K. Hess, F. Köhler, et al., Chem. Rev. 105 (2005) 3758–3772. doi: 10.1021/cr0300901

    63. [63]

      I. Aviv-Harel, Z. Gross, Coord. Chem. Rev. 255 (2011) 717–736. doi: 10.1016/j.ccr.2010.09.013

    64. [64]

      M. Pribisko, J. Palmer, R.H. Grubbs, et al., Proc. Natl. Acad. Sci. U. S. A. 113 (2016) E2258–E2266. doi: 10.1073/pnas.1525564113

    65. [65]

      L.G. Liu, Y.M. Sun, Z.Y. Liu, et al., Inorg. Chem. 60 (2021) 2234–2245. doi: 10.1021/acs.inorgchem.0c03016

    66. [66]

      J. Bendix, I.J. Dmochowski, H.B. Gray, et al., Angew. Chem. Int. Ed. 39 (2000) 4048–4051. doi: 10.1002/1521-3773(20001117)39:22<4048::AID-ANIE4048>3.0.CO;2-7

    67. [67]

      B. Basumatary, J. Rai, R.V.R. Reddy, et al., Chem. Eur. J. 23 (2017) 17458–17462. doi: 10.1002/chem.201704457

    68. [68]

      Q.C. Chen, N. Fridman, Y. Diskin-Posner, et al., Chem. Eur. J. 26 (2020) 9481–9485. doi: 10.1002/chem.202002682

    69. [69]

      F. Neese, WIREs Comp. Mol. Sci. 2 (2011) 73–78. doi: 10.1002/wcms.81

    70. [70]

      F. Neese, WIREs Comp. Mol. Sci. 12 (2022) e1606. doi: 10.1002/wcms.1606

    71. [71]

      W. Zhao, Z. He, B.Z. Tang, Nat. Rev. Mater. 5 (2020) 869–885. doi: 10.1038/s41578-020-0223-z

  • Figure 1  (a) Structures of free-base aromatic porphyrin, corrole, and phthalocyanine. (b) Structure of the free-base benzocorrole radical and schematic of its coordination performance and reduction-induced luminescence.

    Figure 2  (a) Synthesis of BC: (i) CH3OH, H2O, HCl, r.t.; (Ⅱ) CHCl3, p-chloranil, reflux; (Ⅲ) 250 ℃, in vacuo. (b) ESR spectrum of BC in CH2Cl2. (c) Top and (d) side view of the single-crystal structure of BC with thermal ellipsoids at 30% probability. (e) Packing view of BC with space-fill representation.

    Figure 3  (a) Electrochemical measurements of BC in CH2Cl2. (b) Absorption and fluorescence spectra of 1 in CH2Cl2ex = 415 nm), and absorption spectra of BC in various solvents. (c) Images of BC solutions displaying the color change upon protonation and the fluorescence after reduction. (d) Absorption spectral change of BC upon adding HCl in the CH2Cl2 solution. (e) Absorption spectral changes of BC during oxidation and reduction recorded by spectroelectrochemistry. (f) Absorption, fluorescence (λex = 458 nm), and MCD spectra of BC in DMF (10 µmol/L, 3 mL) upon adding DBU (0.5 µL).

    Figure 4  (a, b) Frontier molecular orbital diagrams of BC, TPC, TPP, and BC, with the spin density map of BC. (c) TD-DFT simulated emission spectra of BC. (d) The electron (blue) and hole (yellow) distribution of the S1 excited state on BC. (e) Molecular electrostatic potential and ICSS maps (1 Å above the XY plane) of BC, TPC, and TPP.

    Figure 5  (a) Synthesis and chemical structures of the radical complexes. (i) Zn-BC: Zn(OAc)2·2H2O, pyridine, reflux; (Ⅱ) Ga-BC: GaCl3, pyridine, reflux; (Ⅲ) Pd-BC: Pd(OAc)2, pyridine, reflux. Single-crystal structures of Ga-BC and Pd-BC are presented (thermal ellipsoids at 30% probability) with central coordination structures highlighted. (b) Fluorescence spectra of Ga-BC in toluene as a function of the volume proportion of the added pyridine (λex = 467 nm). (c) Absorption and emission spectra, and time-resolved photoluminescence (PL) decay curve of Pd-BC in deaerated DMF upon adding DBU (λex = 624 nm).

    Figure 6  Schematic diagram of the calculated energy gaps, spin-orbital coupling constants between S1 and Tn (n = 1–4), and proposed energy transfer processes for fluorescence (Fluo.) and phosphorescence (Phos.).

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  8
  • HTML全文浏览量:  1
文章相关
  • 发布日期:  2025-02-15
  • 收稿日期:  2024-05-23
  • 接受日期:  2024-07-23
  • 修回日期:  2024-07-14
  • 网络出版日期:  2024-07-23
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章