Open Access Article
Qiong-Yan
Hong
,
Bin
Huang
,
Yanfei
Niu
,
Cuihong
Wang
,
Xiao-Li
Zhao
,
Hai-Bo
Yang
and
Xueliang
Shi
*
State Key Laboratory of Petroleum Molecular & Process Engineering, Shanghai Key Laboratory of Green Chemistry and Chemical Processes, School of Chemistry and Molecular Engineering, East China Normal University, Shanghai 200062, China. E-mail: xlshi@chem.ecnu.edu.cn
First published on 27th November 2025
Phenoxyl radicals have attracted considerable attention due to their unique electronic structures and wide-ranging applications in physics, chemistry, biology, and materials science. Herein, we report the synthesis and in-depth characterization of a previously unreported 4,4′,4″-nitrilotriphenoxyl radical 3, structurally derived from Yang's biradical scaffold. Interestingly, electronic structure analyses reveal that 3 is a monoradical with an open-shell doublet ground state, wherein the unpaired electron is delocalized over the three peripheral 2,6-di-tert-butylphenoxyl units. Notably, radical 3 is metastable in solution, reverting to closed-shell hydroxyl precursor 2 with a half-life of ∼116 minutes. The hydroxyl 2 can be deprotonated to form anionic 4, which exhibits a closed-shell singlet ground state. Their unique electronic structures are further elucidated by single-crystal X-ray diffraction and assisted by computational methods, in which 3 is fully conjugated and adopts a C3-symmetric geometry, while both 2 and 4 exhibit distinct quinoidal character with apparent C2-symmetry. These findings offer fundamental insights into the electronic structures of nitrogen-bridged polyphenoxyl radicals and establish a new design platform for stable organic open-shell systems.
The unpaired electron in a phenoxyl radical is primarily localized on the oxygen atom and the ortho and para positions of the aromatic ring (Fig. 1a). Consequently, introducing bulky tert-butyl groups at these high-spin-density sites serves as an effective strategy for stabilizing the radical center.37,38 Beyond steric protection, the electronic delocalization of phenoxyl radicals can be finely tuned through para-substitution, allowing for systematic structural modifications.39–48 Notably, para-bridging enables the construction of poly(phenoxyl) radicals by linking multiple radical units, offering a valuable platform for investigating their collective electronic properties and reactivities.1,39–54 A diverse array of poly(phenoxyl) radicals featuring different bridging motifs has been synthesized, many of which exhibit distinctive electronic structures, pronounced near-infrared (NIR) absorption,55 and remarkable thermal and redox stability.56–58 The bridging units in these systems can generally be classified into two categories (Fig. 1b): (1) π-conjugated frameworks, including naphthalene diimide (NDI),59 quinoidal bisanthene,60 thiophene-based heterophenoquinones,61 perylene diimide (PDI),62 helicenes,63 boron dipyrromethene (BODIPY),64 and carbazole,65 and (2) substituted single atoms, such as methine carbon or nitrogen. The nature of the bridging motif plays a pivotal role in determining the ground-state electronic configuration. For instance, bridging via a methine carbon leads to the formation of the classic Galvinoxyl monoradical,66,67 while further adding a 2,6-di-tert-butylphenoxyl unit yields Yang's biradical (Fig. 1b).68,69 Of particular interest are nitrogen-bridged polyphenoxyl systems, which continue to attract attention due to their structural diversity and unique electronic behavior. Imine-bridged derivatives typically yield monoradicals,70 whereas amine linkages favor closed-shell configurations.71,72 Intriguingly, theoretical studies have suggested that nitrogen-centered analogs of Yang's biradical may exist as either open-shell monoradicals or high-spin quartet species. However, previously reported analogs appear unstable, and their detailed electronic structures remain poorly understood.73,74
In this study, we report the successful synthesis and comprehensive characterization of a previously unknown 4,4′,4″-nitrilotriphenoxyl radical 3, structurally derived from Yang's biradical framework (Fig. 1b). Single-crystal X-ray diffraction reveals that 3 adopts a planar, C3-symmetric geometry. Notably, 3 exhibits high reactivity and limited persistence in solution, with a measured half-life of ∼116 minutes in toluene under ambient conditions, gradually converting to its closed-shell hydroxyl precursor 2. Electronic structure analysis via DFT calculations shows that the unpaired electron in 3 is primarily delocalized over the three phenoxyl moieties, while the central nitrogen atom remains spin-inactive, consistent with the absence of 14N hyperfine splitting in the EPR spectrum. Variable-temperature EPR spectroscopy further confirms that 3 adopts an open-shell doublet ground state. In contrast, both 2 and its deprotonated anionic form 4, exhibit C2-symmetric geometries and are EPR silent with well-resolved 1H NMR signals, indicating closed-shell singlet ground states. These findings collectively reveal distinct ground-state electronic structures across this series and provide rare insights into the structure–property relationships of nitrogen-bridged polyphenoxyl radicals, offering a valuable platform for designing new open-shell organic materials.
O stretching band, consistent with its more delocalized electronic structure and elongated C–O bonds, which impart greater single-bond character and thus a reduced stretching force constant compared with 2 and 4 (Fig. S14). Crucially, their molecular structures were unambiguously confirmed by single-crystal X-ray diffraction (see below), providing solid structural evidence for their identities.
The photophysical properties of compounds 1–4 reveal their distinct electronic structures (Fig. 3a). The colorless compound 1 shows a strong absorption band between 260 and 400 nm, with a maximum at 300 nm. Upon oxidation to form 2, the solution turns blue, and the absorption spectrum undergoes a bathochromic shift, displaying a new peak at 690 nm. Further oxidation to the radical 3 deepens the solution color to dark blue and causes a hypsochromic shift in the absorption maximum to 610 nm, approximately 80 nm blue-shifted compared to 2. This shift is attributed to a larger SOMO–SUMO gap in 3 (2.14 eV, Fig. S33) relative to the HOMO–LUMO gap in 2 (1.93 eV, Fig. S31), as revealed by DFT calculations. Subsequent deprotonation of 2 to the anionic species 4 results in a light green solution, with a significantly red-shifted absorption maximum at 815 nm. This pronounced bathochromic shift reflects a narrower HOMO–LUMO gap in 4 (1.52 eV, Fig. S34). The calculated absorption spectra exhibit good agreement with the experimental results, particularly in the main absorption regions. Specifically, the major absorption bands of compound 2 (600–800 nm) and compound 3 (500–800 nm) mainly originate from the HOMO → LUMO transitions (Tables S3–S5 and Fig. S37). Differently, the main absorption band of compound 4 (600–900 nm) is attributed to both HOMO → LUMO and HOMO−2 → LUMO transitions (Table S6). Altogether, the clear spectral differences among 1–4 underscore their distinct electronic configurations, spanning closed-shell neutral, radical, and anionic states. These results highlight the tunability of the optical and electronic properties via redox and deprotonation modulation within this nitrogen-bridged triphenoxyl system.
The electrochemical properties of 2 and 3 were investigated using cyclic voltammetry (CV) and differential pulse voltammetry (DPV) in a dry DCM/toluene (1/4, v/v) solution. Due to the presence of a quinoidal conjugation electronic structure, 2 exhibited two reversible one-electron reduction waves with E1/2 at −1.86 and −1.56 V (vs. Fc/Fc+) (Fig. 3b and S25). These processes generated radical anions and dianions (Fig. 3c), while a reversible oxidation (E1/2 = 0.37 V vs. Fc/Fc+) yielded a radical cation (Fig. 3b, c and S25). Such redox behavior of 2 was consistent with that of the amine-bridged derivative of the Galvinoxyl radical 2-OMe (Fig. S24).71 Compared to 2, 3 displayed three reversible one-electron reduction processes with half-wave potentials (E1/2) at −2.01, −1.62, and −1.30 V (vs. Fc/Fc+) (Fig. 2b), corresponding to the formation of anions, radical dianions and trianions, respectively (Fig. 2d). No oxidation peak was observed for 3, indicating that it is a fully oxidized end state (Fig. S27). The divergent redox behaviors observed in compounds 2 and 3 reflected their fundamentally distinct electronic structures.
Under ambient conditions in dry toluene, 3 gradually decayed, as monitored by time-dependent UV-vis-NIR spectroscopy (Fig. 4a). Its characteristic absorption bands (500–900 nm, λmax = 610 nm) diminished over time, while a new band (400–800 nm, λmax = 690 nm) appeared, with well-defined isosbestic points at 325, 370, 420, 522, 645, and 745 nm, indicating a clean single-step conversion. Kinetic analysis gave a half-life of 116 minutes, confirming the meta-stable nature of 3 (Fig. 3b). In addition, we extended the investigation to other solvents, including mesitylene, dichloromethane and methanol. The results show that 3 exhibits comparable stability in dry mesitylene and dichloromethane, with half-lives of 53 minutes and 46 minutes, respectively (Fig. S17 and S18). In contrast, when dissolved in methanol, the recorded spectrum of the solution was identical to that of 2 in methanol (Fig. S19). Interestingly, the final product of 3 in different solvents all exhibited an absorption spectrum nearly identical to that of 2, confirming that 3 converts into 2 upon decay. This transformation is consistent with a protonation-coupled reduction pathway, analogous to the known phenoxyl radical to phenol conversion.75–79 To verify this, acetic acid was added to the toluene solution of 3, and the characteristic absorption band at 610 nm disappeared completely, leaving only the spectrum of 2 (Fig. 4c). In contrast, adding acetic acid to 2 or a reported phenoxyl radical derivative 2-OMe produced no change. These results demonstrate that acid accelerates the decay of 3. Cyclic voltammetry indicates that the reduction potential of 3 alone is insufficient to drive this process; however, protonation likely lowers the LUMO energy level, thereby facilitating reduction and promoting the conversion from 3 to 2 (Fig. 4d).
![]() | ||
| Fig. 4 The decay process of 3 and its mechanism. (a) Change in the UV-vis-NIR absorption spectra of 3 (∼4.4 × 10−4 M in toluene) over time. (b) Stability plot of 3. All spectra were recorded in toluene under ambient conditions (air, room temperature, and normal light). (c) Normalized UV-vis-NIR absorption spectra (∼4.4 × 10−4 M in toluene) of 2, 3, and the reported compound 2-OMe71 before and after adding AcOH. (d) Proposed mechanism for the decay of 3 into 2 involving a protonation-coupled reduction process. | ||
Single crystals of compounds 1–4 suitable for X-ray diffraction were obtained by slow diffusion of hexane into dichloromethane under inert conditions.80 Structural analysis revealed two distinct conformational types (Fig. 5). Compounds 1 and 3 exhibit C3 symmetry, with 3 showing a fully conjugated, nearly planar framework. In contrast, compounds 2 and 4 display C2 symmetry and a pronounced quinoidal character. These conformational differences arise from variations in electronic delocalization around the central nitrogen atom. In 1 and 3, the three bond angles (θ1–θ3) around N1 are nearly equivalent, consistent with their symmetrical geometries. For 2 and 4, the angles are markedly unequal (Table 1). The N1 atom in 1 lies 0.058 Å out of the C4–C7–C13 plane, indicative of a flattened tetrahedral geometry due to a localized nitrogen lone pair. By contrast, in 2–4, especially 3, N1 is essentially coplanar with adjacent atoms, reflecting delocalization of the lone pair over the extended π-system and favoring a fully planar structure. Bond length analysis further supports these findings (Table 1). 3 features highly delocalized phenoxyl units with uniform C–O bonds (1.269 Å), significantly shorter than the C–OH bonds in 1 and 2, yet slightly longer than a conventional C
O double bond. The C–N bonds (C4–N1, C7–N1, and C13–N1) in 3 and 4, as well as C7–N1 and C13–N1 in 2, are shorter than the single C–N bond in 1 (1.440 Å) (Fig. 5 and Table 1), highlighting enhanced conjugation. Bond length alternation (BLA) analysis of 3 reveals values intermediate between 1 (localized) and 2/4 (quinoidal), consistent with its fully delocalized electronic structure. To corroborate these interpretations, variable-temperature 1H NMR (VT-NMR) of 2 was performed. Upon heating, significant signal broadening in both aromatic and aliphatic regions (Fig. S12) was observed, confirming restricted C–N bond rotation and thus its partially conjugated, quinoidal nature.71
| 1 | 2 | 3 | 4 | |
|---|---|---|---|---|
| a Bond lengths, dihedral angles and BLA parameters of 1, 2, 3 and 4 were determined from single crystals. | ||||
| O1–C1 (Å) | 1.319 | 1.344 | 1.269 | 1.261 |
| O3–C16 (Å) | 1.319 | 1.248 | 1.269 | 1.251 |
| O2–C10 (Å) | 1.319 | 1.248 | 1.269 | 1.263 |
| C4–N1 (Å) | 1.440 | 1.453 | 1.408 | 1.399 |
| C7–N1 (Å) | 1.440 | 1.388 | 1.408 | 1.405 |
| C13–N1 (Å) | 1.440 | 1.388 | 1.408 | 1.400 |
| BLA(A) (Å) | 0.0538 | 0.0085 | 0.0655 | 0.0638 |
| BLA(B) (Å) | 0.0538 | 0.0688 | 0.0655 | 0.0675 |
| BLA(C) (Å) | 0.0538 | 0.0688 | 0.0655 | 0.0738 |
| θ1 (°) | 119.84 | 118.60 | 120.00 | 120.89 |
| θ2 (°) | 119.84 | 118.60 | 120.00 | 120.23 |
| θ3 (°) | 119.84 | 122.19 | 120.00 | 118.87 |
| d (Å) | 0.058 | 0.000 | 0.000 | 0.008 |
The electronic structures of 2, 3 and 4 were further investigated using EPR and NMR spectroscopy. Both 2 and 4 were EPR silent (Fig. 6a), but showed sharp 1H NMR peaks (Fig. 6b), indicating that both 2 and 4 adopted closed-shell singlet ground states (Fig. 6a and b). This result was consistent with previous studies, indicating that such structures exhibited a distinct quinoidal character with a very small diradical contribution.71,72 In contrast, 3 showed a broad and intense EPR signal (Fig. 6a), while its 1H NMR spectrum at room temperature was completely broadened (Fig. 6b), indicating its paramagnetic behavior. The absence of well-resolved hyperfine splitting in the EPR spectra of 3 can be attributed to the delocalization of the unpaired electron over multiple aromatic rings leading to overlapping hyperfine features from several equivalent hydrogen atoms. The g-value of 3 is determined to be 2.0045 (Fig. S28). This g-value is characteristic of delocalized organic oxygen-centered radicals; the slightly stronger spin–orbit coupling of the oxygen atom results in a marginally higher g-value than that of typical carbon-centered radicals (≈2.0030). This observation supports that the unpaired electron is predominantly delocalized over the conjugated C–O π-system (phenoxyl moieties) rather than localized on a single heavy-atom center. Subsequently, the ground state of compound 3 was investigated by variable-temperature EPR (VT-EPR) spectroscopy in toluene over the temperature range of 105–295 K (Fig. 6c). Upon cooling, the EPR signal intensity of 3 gradually increased, a common behavior for paramagnetic monoradicals,81 attributable to the temperature-dependent Boltzmann distribution of spin populations. The corresponding I versus 1/T plot (where I represents the EPR signal intensity at each temperature and T represents the temperature) is shown in Fig. S30. The I value exhibited a linear correlation with 1/T, consistent with the Curie law for paramagnetic species. This behavior is characteristic of monoradicals with a doublet ground state, further confirming that compound 3 exists predominantly in a thermally stable open-shell state without significant thermal population of high spin states.
To gain deeper insight into the electronic structure of 3, we performed DFT calculations (UB3LYP/6-31G(d)) of its ground state, spin density, and frontier molecular orbitals, alongside a comparative study of Yang's biradical at the same level. The electronic structure was analyzed using Multiwfn.82 The results show that 3 adopts a doublet ground state with a large doublet-quartet energy gap (ΔED–Q ≈ −6.27 kcal mol−1) (Fig. S32), whereas Yang's biradical favors a triplet ground state (ΔES–T ≈ 5.42 kcal mol−1) (Fig. S35). The unpaired electron in 3 is extensively delocalized over the three phenoxyl units, mainly on the oxygen atoms and the ortho/para positions of the aromatic rings, with no spin density on the central nitrogen atom (Fig. 7a). This accounts for the absence of 14N hyperfine splitting in its EPR spectrum (Fig. 6a). In contrast, the singlet spin density of Yang's biradical is localized on just two phenoxyl moieties (Fig. 7a), though their triplet states exhibit similar distributions (Fig. 7b). The marked difference in ground-state multiplicity arises from the replacement of the central carbon in Yang's biradical with a π-conjugated sp2 nitrogen atom in 3. The nitrogen's lone pair engages in π-conjugation with adjacent phenyl rings, stabilizing the single unpaired electron and favoring a doublet state. In Yang's biradical, the carbon bridge lacks this conjugative stabilization, leading to two unpaired electrons and a triplet ground state, consistent with its trimethylenemethane analogue.83 Resonance analysis of 3 reveals contributions from ionic structures (e.g., 3a) and delocalized quinoidal monoradical forms (3b–3d) (Fig. 7a). Frontier molecular orbital (FMO) analysis further supports these findings. 3 shows significant SOMO–SUMO overlaps with a large energy gap of 2.14 eV (Fig. 7a), matching the observed absorption maximum at 610 nm and confirming the high electronic stability of this delocalized open-shell species. Overall, nitrogen bridging in 3 fundamentally alters its electronic configuration compared to carbon-bridged Yang's biradical, conferring doublet character, enhanced planarity, and global π-delocalization.
CCDC 2454706 (1), 2454730 (2), 2454731 (3) and 2454733 (4) contain the supplementary crystallographic data for this paper.84a–d
| This journal is © The Royal Society of Chemistry 2026 |