A cationic radical lanthanide organic tetrahedron with remarkable coordination enhanced radical stability

Zhengzhong Zhu Shaojun Hu Zhi Liu Lipeng Zhou Chongbin Tian Qingfu Sun

Citation:  Zhengzhong Zhu, Shaojun Hu, Zhi Liu, Lipeng Zhou, Chongbin Tian, Qingfu Sun. A cationic radical lanthanide organic tetrahedron with remarkable coordination enhanced radical stability[J]. Chinese Chemical Letters, 2025, 36(2): 109641. doi: 10.1016/j.cclet.2024.109641 shu

A cationic radical lanthanide organic tetrahedron with remarkable coordination enhanced radical stability

English

  • Coordination–driven self–assembly, a fascinating synthetic strategy in supramolecular chemistry, is widely used to prepare a variety of discrete metallosupramolecular architectures, including helicates, macrocycles/wheels and cages/capsules [16]. So far, the majority of the coordination–supramolecular architectures are based on transition metal ions [722], such as Pd, Pt, Fe, Au, Ag and Zn, while architectures containing lanthanide [2327] are less explored. This is mainly due to the difficulty in controlling the coordination number and geometry of Ln ions, which have variable and labile coordination habits. However, lanthanide supramolecular compounds have attractive optical and magnetic properties derived from the 4f electrons, which make them valuable for functional exploration. Our research group has been devoted to the synthesis and study of lanthanide organic polyhedrons (LOPs) since 2015 [24]. Along with other groups, we have successfully prepared several types of Ln2L3 helicates, Ln4L4 tetrahedrons and Ln8L6 cubes [23,24] with particular interests toward their luminescence, cell imaging, chirality, and catalysis [23,2833]. However, all reported LOPs are synthesized by diamagnetic closed–shell ligands. Incorporating radical–containing ligands into the multi–component assemblies can provide a new platform to examine multiple spin–spin magnetic interactions and access novel functions that are unavailable in single radical systems [3437]. To the best of our knowledge, no examples of such radical lanthanide polyhedrons exist.

    Organic radical species are prone to disproportion or intermolecular dimerization. It had been demonstrated that covalent approach was an effectively method to improve the stability of organic radicals by using bulky substituents for steric protection or introducing conjugated groups [38,39], polar substituents [40] or heteroatoms [41] to delocalize the spin density. To prevent the dimerization, non–covalent strategy, also known as supramolecular approach, mainly relies on host–guest chemistry [4244] and non–covalent interactions, such as hydrogen bonds [45], electrostatic interaction [46], ππ interaction [47]. Apart from these, coordination bonds were found to enhance the organic radicals' stability. For example, Wang and co–works demonstrated that metal–coordination can synergistically stabilize the radical cations of the main group elements [48]. Moreover, recent reports have also revealed that the coordination of radical ligands to transition metal ions could generate impressive radical stability in metallacycles [4951]. However, combining radical ligand with metal organic polyhedrons aimed at improving the radical stability was rarely investigated [5254], in particular by coordination with lanthanide ions.

    Triphenylamine (TPA) radical cations and its derivatives, obtained by one–electron oxidation of triphenylamine, have been extensively applied in a wide range of areas, including organic and hybrid solar cells, organic redox catalysis, magnetic materials, etc. [5466]. Herein, we report the first Ln4(L•+)4–type (Ln = La, Eu and Tb) radical–bridged lanthanide organic tetrahedrons synthesized from self–assembly of face–capping TPA–cored radical ligand with Ln ions (Scheme 1). TPA–cored cationic radical ligand L1•+ features three peripheral electron–donating pyridine diamide chelating sites at para position. For an facilitate–intuitive comparison, L2 with a triazolyl group replacing inner amide group on L1, and its Ln4(L2)4 assemblies were also synthesized. However, it was observed that L2 failed to oxidize into its radical form. All compounds have been well characterized by 1H NMR, DOSY and ESI–TOF–MS. Moreover, the structure of Eu4(L1)4, Eu4(L1•+)4 and Eu4(L2)4 have been unambiguously confirmed by X–ray crystallographic analyses, wherein Eu4(L1•+)4 represents the first example of radical LOPs. The generation of radical cations, their stability and lanthanide–radical magnetic interaction were further studied by UV−vis−NIR, CV, EPR spectra, magnetic susceptibilities measurements, together with density functional theory (DFT) calculations. After coordination with lanthanide ions, the enhanced radical half–life time of L1•+ in Ln4(L1•+)4 assemblies was observed. Remarkably, Eu4(L1•+)4, Tb4(L1•+)4, and Tb4(L1•+)4 exhibit much longer radical half–life time than that of La4(L1•+)4 and Lu4(L1•+)4, likely due to the magnetic interactions between paramagnetic Ln ions and L1•+.

    Scheme 1

    Scheme 1.  Synthetic route of TPA–bridging ligands L1, L2, L1•+ and self–assembly of Ln4(L)4 tetrahedral cages. (ⅰ) 6–(Isopropylcarbamoyl)picolinic acid, HATU, Et3N, 70% yield, (ⅱ) NaNO2/HCl, NaN3, (ⅲ) 6–ethynyl–N–isopropylpicolinamide, SA, CuSO4·5H2O, 75% yield, (ⅳ) NOBF4, 95% yield, (ⅴ) Ln(OTf)3, 99% yield.

    Ligand L1 was readily synthesized by a one–step amide condensation reaction between 6–(isopropylcarbamoyl)picolinic acid and a C3–symmetric tri(4–amino)triphenylamine (3NH2–TPA) core. As a control, ligand L2 was prepared by the Click reaction (Scheme 1, Scheme S1 and S2, Figs. S1–S4, S9 and S10 in Supporting information).

    When ligand was treated with 1 equiv. of Ln(OTf)3 in acetonitrile solution at room temperature for serval minutes, the turbid suspension of the ligand quickly turned into a clear solution, indicating the generation of new species. Taking L1 as an example, NMR and ESI–TOF–MS spectra show the formation of a high symmetric Eu4(L1)4 species (Fig. 1, Figs. S5 and S6 in Supporting information). Compared to free L1, characteristic upfield shifts for Hc-g on the chelating moieties of L1 are clearly observed (Fig. 1b), while Ha and Hb on TPA core are shifting downfield, consistent with the coordination of chelating moieties with Eu ions. Moreover, the signal of Hi on the terminal isopropyl splits into two doublets with equal integration, implying a rigid coordination configuration. The 1H diffusion–ordered NMR spectroscopy (DOSY, Fig. 1c) displays the presence of sole product with a hydrodynamic diameter of 3.2 nm. As shown in Fig. 1d, high resolution ESI–TOF–MS analyses confirm the formation of the assembly with a chemical formula of Eu4(L1)4(OTf)12. A series of prominent peaks with charging states from 3+ to 6+ derived by the successive loss of OTf counterions and active hydrogen on the amide of Eu4(L1)4(OTf)12 are observed. For instance, peak with m/z = 898.8013 corresponds to the [Eu4(L1)4(OTf)12–H4–(OTf)9]5+ charged species (Fig. 1d, inset). After replacing Eu ion with La or Tb ions, 1H NMR spectrum and/or the ESI–TOF–MS also confirmed the formation of La4(L1)4 and Tb4(L1)4 (Figs. S7, S14 and S17 in Supporting information). The analogous Eu4(L2)4 complex was also synthesized in a similar way for comparison purpose (Fig. S11 in Supporting information).

    Figure 1

    Figure 1.  1H NMR of (a) ligand L1 in DMSO–d6 and (b) tetrahedral cage Eu4(L1)4(OTf)12 in CD3CN. (c) 1H DOSY spectrum of Eu4(L1)4(OTf)12. (d) ESI–TOF–MS of Eu4(L1)4(OTf)12 with insets showing the observed (Obs.) and simulated (Sim.) isotopic patterns for the 5+ peaks. Single crystal X–ray structures of (e) Eu4(L1)4 and (f) Eu4(L2)4. Eu: Cyan, C: green, N: blue, O: red, H: white.

    Single crystals of the Eu4(L1)4 and Eu4(L2)4 assemblies were grown by slow vapor diffusion of diethyl ether into their acetonitrile solution, confirming their tetrahedral geometry (Figs. 1e and f), which are isomorphic to our previous reported Eu4L4–type tetrahedral cages [67,68]. The four metal centers on the tetrahedron display cooperative ΔΔΔΔ or ΛΛΛΛ chirality, and both enantiomers coexist in crystals (Figs. S20–S22 in Supporting information), as they all crystallize in the achiral Fm-3 space group. In the crystal structures of Eu4(L1)4 and Eu4(L2)4, each Eu centers are nine–coordinated and ligated by three chelating moieties from three different ligands.

    To investigate the redox properties of the free ligands and their assemblies, we initially oxidized the neutral ligand with nitrosonium tetrafluoroborate (NOBF4), a common one–electron oxidant. Upon addition of NOBF4, the solution color of L1 in acetonitrile rapidly changed from yellow to green, which is a characteristic color of TPA–cored radical cation, indicating the generation of L1•+. The formation of radical cation was also unambiguously confirmed by the UV–vis and EPR spectra (vide infra). It should be noted that the 1H NMR spectra of L1•+ were significantly broadened, due to the paramagnetic nature of the radical species on the TPA units (Fig. S4). Unfortunately, we failed to produce the radical cation L2•+ using several reported efficient oxidation reagents, including NOBF4, Cu(ClO4)2 and AgBF4, suggesting its high oxidation potential.

    To elucidate the distinction between L1 and L2, cyclic voltammetry (CV) was measured for both ligands in DMSO containing tetrabutylammonium hexafluorophosphate (nBu4NPF6) at a scan rate of 100 mV/s (vs. Ag/AgCl) under an argon atmosphere. Free ligand L1 exhibited one reversible oxidation wave around +0.82 V and an irreversible oxidation peak around +1.22 V, corresponding to the formation of L1•+ and L12+, respectively (Fig. 2a). Similar CV waves were observed for L1•+ (Fig. S48 in Supporting information), which further confirmed the assignment of the oxidation peaks and the stability of the radical cation form. However, only one reversible oxidation wave was observed around +1.23 V for ligand L2 (Fig. 2b), which was even higher than the second oxidation potential of L1. We also calculated the frontier orbitals of ligands L1 and L2 using DFT methods at the UB3LYP/6–31G(d) level of theory to explain the different behavior under electrochemical condition. Fig. 2c shows that the energy of the HOMO (HOMO = highest occupied molecular orbital) of L2 (–5.54 eV) is lower than L1 (–4.65 eV) by 0.89 eV [69]. This lower delocalized HOMO indicates weaker electron loss ability, which is in good agreement with the higher first reversible oxidation potential (Eox) of L2, consistent with the CV results (Eox (L1) = 0.82 V vs. Eox (L2) = 1.23 V). Moreover, we performed the CV experiments for Eu4(L1)4 and Eu4(L2)4 in acetonitrile. Compared with L1•+, only one broadened oxidation wave was observed for the cage Eu4(L1)4 with a positive shift of the oxidation potential around +1.44 V, likely due to the complexation with Eu and the polycationic nature (12+) of the Ln–linked cage structure [52,70]. This behavior demonstrates the considerable stability of the Eu4(L1•+)4 radical cage as well as the absence of disassembly during the redox process. In sharp contrast, only one irreversible oxidation wave was observed around 1.37 V in the Eu4(L2)4 acetonitrile solution.

    Figure 2

    Figure 2.  (a, b) Cyclic voltammetry of ligand L1, Eu4(L1)4 and L2, Eu4(L2)4 (1 mmol/L for ligands, 0.25 mmol/L for cages, 0.1 mol/L nBu4NPF6, scan rate 100 mV/s) at room temperature. (c) Frontier orbitals of L1 and L2 calculated using DFT methods (B3LYP/6–31G(d)).

    Next, we examined the assembly of L1•+ with Ln ions. We obtained a clear assembly solution by mixing the radical cation L1•+ with Eu(OTf)3 in acetonitrile. The UV–vis spectrum of this solution showed three absorption peaks, two of which (under 500 nm) were similar to the absorption peaks of Eu4(L1)4, indicating the formation of Eu4(L1•+)4 (Fig. S34 in Supporting information). It should be noted that a new shoulder peak at about 735 nm in Eu4(L1•+)4 was observed compared to Eu4(L1)4. To verify this difference, we measured the UV–vis absorption of L1 and L1•+ in the same solution (Fig. 3a). Unlike L1, a typical radical absorption peak around 784 nm was found in L1•+, in agreement with the reported TPA radical cation [71]. Therefore, the new peak observed in Eu4(L1•+)4 could be attributed to the absorption band of the radical ligands L1•+ (Fig. S34), and its slight blue shift in Eu4(L1•+)4 may be caused by the coordination with Eu ions. Similarly, the successful preparation of La4(L1•+)4 and Tb4(L1•+)4 was also confirmed by the UV–vis spectra (Fig. 3a and Fig. S34). We failed to obtain ESI–TOF–MS data for the radical assemblies Ln4(L1•+)4 (Ln = La, Eu and Tb), which may be due to their easy reduction under the ionization conditions.

    Figure 3

    Figure 3.  (a) Absorption spectra of L1 (0.1 mmol/L, blue dot line), L1•+ (0.1 mmol/L, blue line), Tb4(L1)4 (0.025 mmol/L, pink dot line), Tb4(L1•+)4 (0.025 mmol/L, pink line) in CH3CN, Inset: the color view of L1 and L1•+. (b) Excitation and emission spectra of Tb4(L1)4 (blue line) and Tb4(L1•+)4 (pink line) in CH3CN.

    Interestingly, the greenish cubic single crystals (Fig. S21) of the radical assembly were grown overnight by the same vapor diffusion method as for the Eu4(L1)4 crystal growth process. Like Eu4(L1)4, this radical cage also crystalizes in the cubic space group Fm-3. The main difference lies in their C–N bond length on the TPA panels. This bond length in the radical cage (1.404 Å, Fig. S28 in Supporting information) is shorter by about 0.015 Å than that in cage Eu4(L1)4 (1.419 Å), similar to the trends observed in the mono–radical cation of thiophene–bridged bis(TPA) [72], indicative of the radical feature of TPA panels in radical cage. The single crystal structure provides reliable evidence for the existence of radical centers in a tetrahedral cage. However, the crystal structure cannot confirm that each tetrahedral cage contains four radicals, due to the easy recovery feature of TPA-type radical during the crystal growth process. According to the calculated half–life time from UV–vis tracking experiment (vide infra) coupled with the crystal growth time, we speculate that about 2 radicals may be remained in a cage after the crystallization process.

    The spin density distribution of L1•+ was estimated by DFT calculations at the UB3LYP/6–31G(d) level of theory. Spin density of L1•+ was found to be mainly delocalized on the TPA core and three inner amido groups, with the highest Mulliken atomic spin density (M–ASD) of ~0.3 single electron on the central N atom (Fig. S33 and Table S4 in Supporting information for the spin density map and the M–ASD values of all atoms in L1•+). The delocalization effects resulted in partial double bond character for the central three C–N single bonds, which accounted for the shortened C–N bond length observed in the radical Eu4(L1•+)4 cage. Most of the organometallic assemblies reported in the literature were prepared by the reaction of a neutral or an anionic ligand with a cationic metal precursor, while those fabricated by a cationic ligand and metal ions were rare [52,73]. In some cases, transferring the neutral ligand into cationic state even leads to the disassembly of the assemblies [74]. The stability of our assembly may be attributed to the low positive charge density of the radical cation on the tridentate chelating claws, which reduced the cation–cation repulsion between the radical ligand and the Ln ion.

    Luminescence measurements indicated that both L1 and L2 can sensitize Tb (Fig. 3b and Fig. S46 in Supporting information). Taking Tb4(L1)4 as an example, upon excitation at 315 nm, a cascade of characteristic emission bands at 618, 583, 543, 488 nm assigning to 5D47FJ (J = 3 − 6) transitions is observed. The longest wavelength absorption maximum of Tb4(L1•+)4 appears at 740 nm, with a tail that extends beyond 800 nm. It is predictable that L1•+ cannot sensitize Tb4(L1•+)4 under the same test condition due to low energy levels of the radical species. As shown in Fig. 3b, a broadened band emission centered at 506 nm takes the place of the typical line−like Tb green luminescence, which probably originated from those recovered neutral ligands in Tb's cage.

    The radical stability of L1•+, La4(L1•+)4 and Tb4(L1•+)4 was evaluated by monitoring the time–dependent UV–vis absorption spectra under ambient circumstance (Figs. 4a-c). The half–life times (t1/2) of L1•+, La4(L1•+)4 and Tb4(L1•+)4 were determined to be 53 min, 482 min and 822 min, respectively. The assembly with Ln ions significantly improved the stability of L1•+, with approximately 9−fold and 15−fold enhancements for La and Tb assemblies, respectively (Fig. 4d). The stability of the radical ligands L1•+ could be attributed to the covalent effect, as the incorporation of the ONO chelating moiety to the TPA core enabled the spin density to delocalize to the oxygen atom of inner amide (Fig. S33). Moreover, the coordination of the radical ligands with Ln ions anchored the radical ligands on the four faces of the tetrahedron, resulting in reduced close stacking of the TPA core and thus enhancing the radical L1•+ stability in the cage. Furthermore, OTf and solvent molecules occupied the cavity of the cage and the space between the cages, which prevented the interactions with other reactive molecules, especially O2, and slowed down decay reactions.

    Figure 4

    Figure 4.  Time–dependent UV–vis absorption spectra of (a) L1•+, (b) La4(L1•+)4 and (c) Tb4(L1•+)4 in CH3CN recorded at ambient air and light conditions. (d) Plot of the absorbance at 784 nm (L1•+), 760 nm (La4(L1•+)4), 740 nm (Tb4(L1•+)4) with time, showing the half–life times were increased by 9–fold and 15–fold for La4(L1•+)4, Tb4(L1•+)4, respectively.

    In addition, we found a 1.8-fold increase in t1/2 from La4(L1•+)4 to Tb4(L1•+)4. To understand this difference, we first consider the influence of lanthanide paramagnetism, thereby EPR spectra under ambient conditions were tested in acetonitrile solution. As shown in Fig. 5a, the EPR spectrum of free ligand L1•+ showed a signal at g = 2.003 with a hyperfine structure originating from the nitrogen nuclei (I = 1, triplet), indicating that the higher spin density is mainly localized on the central nitrogen atom. Upon complexation with Ln (Ln = La and Tb) ions, distinct EPR signals were observed, which confirmed the formation of radical L1•+, since the diamagnetic behavior of La ions and the fast spin–lattice relaxation feature of Tb ions make both them EPR silent at room temperature [75]. This was further verified by the absence EPR signals for their non–radical counterparts La4(L1)4 and Tb4(L1)4. These results demonstrated that the radical ligands and assemblies were successfully prepared. Moreover, La4(L1•+)4 exhibited the same EPR hyperfine signals as the radical ligand L1•+, implying that there was no magnetic interaction between the radical and La ions or among the four radical ligands themselves. This is reasonable because of the intrinsic diamagnetism of La ions and the large separation of the central N atoms with higher spin density in the La4(L1•+)4 tetrahedron (the N–N distance in the Eu analogue was 6.669 Å). In sharp contrast, six EPR peaks centered at g = 2.007 were observed for Tb4(L1•+)4, which differed from the radical ligand L1•+, suggesting that there was magnetic interaction between Tb and the organic radical.

    Figure 5

    Figure 5.  (a) EPR spectra of ligand L1 (light yellow line), L1•+ (yellow line, g = 2.003), La4(L1)4 (light blue line), La4(L1•+)4 (blue line, g = 2.003), Tb4(L1)4 (light pink line) and Tb4(L1•+)4 (pink line, g = 2.007) in acetonitrile solution at 293 K. (b) Temperature dependence of χmT for L1•+ (hexagon), Tb4(L1)4 (triangle) and Tb4(L1•+)4 (square) measured at 5000 Oe.

    To gain a deeper insight into the radical characteristic of L1•+ and the magnetic interaction between Tb and the radical spins in Tb4(L1•+)4, direct current (DC) magnetic susceptibilities for the freshly prepared samples of L1•+, Tb4(L1)4 and Tb4(L1•+)4 were recorded in the temperature range of 2 − 300 K. As shown in Fig. 5b, the χmT value of L1•+ at 300 K is 0.37 cm3 K/mol, very close to the expected value for a magnetically isolated system of one S = 1/2 with g = 2.0 (0.375 cm3 K/mol), confirming the open shell character of L1•+. The χmT values decrease continuously with lowing temperature, indicative of the presence of antiferromagnetic (AF) interaction between adjacent L1•+. For Tb4(L1•+)4, the experimental χmT value at room temperature is 47.13 cm3 K/mol, quite close to the theoretical value of 48.75 cm3 K/mol for four non−interacting Tb ions (J = 6, g = 3/2) and four magnetically isolated radical spin units. Upon cooling, the χmT values undergo a smooth decrease until about 50 K, below which it decreases rapidly and reaches a minimum value of 16.30 cm3 K/mol at 2 K. The decrease in low temperature for Tb4(L1•+)4 is most likely caused by the thermal depopulation of Stark sublevels, but the contribution of AF exchange interactions cannot be ignored. In contrast, the χmT product of Tb4(L1)4 maintains roughly constant with lowing temperature down to about 100 K, and then drops to a value of 21.26 cm3 K/mol at 2 K. This different decrease behavior in χmT vs. T curves unambiguously demonstrates the presence of AF coupling between Tb ion and the radical spin carrier. To further prove the AF interactions, the χmT of Tb4(L1)4 is subtracted from the χmT of Tb4(L1•+)4 to offset the contribution stemming from thermal depopulation of the Stark sublevels of Tb ions, and the temperature dependence of ΔχmTχmT = χmT(Tb4(L1•+)4)χmT(Tb4(L1)4)) is obtained (Fig. S51 in Supporting information). The ΔχmT falls progressively with lowering temperature, confirming again the AF correlations between Tb ion and radicals. The inherent unquenched orbital angular momentum of Tb ion together with the delocalization behavior of radical electron make the quantitative estimation of the AF interaction magnitude tricky, which may deserve separate investigation in the future.

    Apart from the influence of lanthanide paramagnetism, the Ln radius may also exert effect on the stability of radical L1•+, as the Ln size can affect the rigidity of the ligands and the planarity of the central N atoms. To study this, the t1/2 values of Eu4(L1•+)4, Gd4(L1•+)4 and Lu4(L1•+)4 were estimated, giving the t1/2 values of 624 min, 1248 min and 347 min, respectively. By comparing the t1/2 values with Ln radii (Tables S5 in Supporting information), it was found that the stability of the radical is not dependent on the Ln radius. However, we observed that the assemblies containing diamagnetic Ln ions (Ln = La and Lu) show lower t1/2 values than those assemblies constructed by paramagnetic Ln assemblies (Ln = Eu, Gd, Tb). Based on this observation, we propose that the existence of magnetic interaction between the radical ligands and paramagnetic Ln ions in Ln4(L1•+)4 (Ln = Eu, Gd and Dy) increases the degree of radical electronic delocalization compared to that in Ln4(L1•+)4 (Ln = La and Lu) fabricated by diamagnetic Ln ions, and hence resulting in enhanced radical stability in Ln4(L1•+)4 (Ln = Eu, Gd and Dy).

    In summary, we report the first example of a radical−bridged lanthanide organic tetrahedron constructed via coordination−driven self−assembly. The progressive stability enhancement of the radical panels was observed from the tetrahedron fabricated by diamagnetic Ln ions to that by paramagnetic Ln ions, as certificated by the UV−vis, EPR and magnetic susceptibility measurements. We infer that the enhancement in paramagnetic Ln ions should be ascribed to the Ln−radical magnetic interaction, which enables the radical electron more delocalization compared to the Ln tetrahedron containing diamagnetic Ln ions. Our findings not only offer some guidance for the preparation of stable radical systems, but also provide new candidates for smart lanthanide 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

    This work was supported by National Key Research and Development Program of China (Nos. 2021YFA1500400 and 2022YFA1503300), the National Natural Science Foundation of China (Nos. 21825107, 21971237, 22171264 and 22301301) and the Science Foundation of the Fujian Province (No. 2021J02016).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.109641.


    1. [1]

      Y. Sun, C. Chen, J. Liu, et al., Chem. Soc. Rev. 49 (2020) 3889–3919. doi: 10.1039/D0CS00038H

    2. [2]

      E.G. Percastegui, T.K. Ronson, J.R. Nitschke, Chem. Rev. 120 (2020) 13480–13544. doi: 10.1021/acs.chemrev.0c00672

    3. [3]

      C.M. Hong, R.G. Bergman, K.N. Raymond, et al., Acc. Chem. Res. 51 (2018) 2447–2455. doi: 10.1021/acs.accounts.8b00328

    4. [4]

      M.D. Ward, C.A. Hunter, N.H. Williams, Acc. Chem. Res. 51 (2018) 2073–2082. doi: 10.1021/acs.accounts.8b00261

    5. [5]

      D.L. Caulder, K.N. Raymond, Acc. Chem. Res. 32 (1999) 975–982. doi: 10.1021/ar970224v

    6. [6]

      M. Fujita, M. Tominaga, A. Hori, et al., Acc. Chem. Res. 38 (2005) 369–378. doi: 10.1021/ar040153h

    7. [7]

      S. Bai, Y.F. Han, Acc. Chem. Res. 56 (2023) 1213–1227. doi: 10.1021/acs.accounts.3c00102

    8. [8]

      C.T. McTernan, J.A. Davies, J.R. Nitschke, Chem. Rev. 122 (2022) 10393–10437. doi: 10.1021/acs.chemrev.1c00763

    9. [9]

      C.B. Tian, Q.F. Sun, Chem. Eur. J. 29 (2023) e202300195. doi: 10.1002/chem.202300195

    10. [10]

      Y. Domoto, M. Fujita, Coord. Chem. Rev. 466 (2022) 214605–214617. doi: 10.1016/j.ccr.2022.214605

    11. [11]

      F. Wang, S. Bai, Q.W. Zhu, et al., CCS Chem. 5 (2023) 633–640. doi: 10.31635/ccschem.022.202201919

    12. [12]

      K. Acharyya, S. Bhattacharyya, H. Sepehrpour, et al., J. Am. Chem. Soc. 141 (2019) 14565–14569. doi: 10.1021/jacs.9b08403

    13. [13]

      P. Howlader, E. Zangrando, P.S. Mukherjee, J. Am. Chem. Soc. 142 (2020) 9070–9078. doi: 10.1021/jacs.0c03551

    14. [14]

      M. Pan, K. Wu, J.H. Zhang, et al., Coord. Chem. Rev. 378 (2019) 333–349. doi: 10.1016/j.ccr.2017.10.031

    15. [15]

      K. Wu, K. Li, S. Chen, et al., Angew. Chem. Int. Ed. 59 (2019) 2639–2643.

    16. [16]

      J. Zhu, X. Chen, X. Jin, et al., Chin. Chem. Lett. 34 (2023) 108002–108005. doi: 10.1016/j.cclet.2022.108002

    17. [17]

      Y.L. Lai, H.J. Zhang, J. Su, et al., Chin. Chem. Lett. 34 (2023) 107686–107690. doi: 10.1016/j.cclet.2022.07.029

    18. [18]

      C. Li, Y. Wang, Y. Lu, et al., Chin. Chem. Lett. 31 (2020) 1183–1187. doi: 10.1016/j.cclet.2019.09.034

    19. [19]

      D. Luo, B. Pan, J. Zhang, et al., Chin. Chem. Lett. 32 (2021) 1397–1399. doi: 10.1016/j.cclet.2020.11.002

    20. [20]

      R.Y. Chen, Y.P. He, J. Zhang, Polyoxometalates 1 (2022) 9140002–9140007. doi: 10.26599/POM.2022.9140002

    21. [21]

      R.Y. Chen, G.H. Chen, Y.P. He, et al., Chin. J. Struct. Chem. 41 (2022) 2201001–2201006.

    22. [22]

      X. Liu, X. Feng, K.R. Meihaus, et al., Angew. Chem. Int. Ed. 59 (2020) 10610–10618. doi: 10.1002/anie.202002673

    23. [23]

      D.J. Bell, L.S. Natrajan, I.A. Riddell, Coord. Chem. Rev. 472 (2022) 214786–214810. doi: 10.1016/j.ccr.2022.214786

    24. [24]

      X.Z. Li, C.B. Tian, Q.F. Sun, Chem. Rev. 122 (2022) 6374–6458. doi: 10.1021/acs.chemrev.1c00602

    25. [25]

      M.H. Du, D.H. Wang, L.W. Wu, et al., Angew. Chem. Int. Ed. 61 (2022) e202200537. doi: 10.1002/anie.202200537

    26. [26]

      X.L. Li, J. Wu, L. Zhao, et al., Chem. Commun. 53 (2017) 3026–3029. doi: 10.1039/C7CC00048K

    27. [27]

      X.L. Li, J. Wu, J. Tang, et al., Chem. Commun. 52 (2016) 9570–9573. doi: 10.1039/C6CC05326B

    28. [28]

      J. Wang, C. He, P. Wu, et al., J. Am. Chem. Soc. 133 (2011) 12402–12405. doi: 10.1021/ja2048489

    29. [29]

      D.E. Barry, D.F. Caffrey, T. Gunnlaugsson, Chem. Soc. Rev. 45 (2016) 3244–3274. doi: 10.1039/C6CS00116E

    30. [30]

      J. Hamacek, A. Vuillamy, Eur. J. Inorg. Chem. 2018 (2017) 1155–1166.

    31. [31]

      H.Y. Wong, W.S. Lo, K.H. Yim, et al., Chem 5 (2019) 3058–3095. doi: 10.1016/j.chempr.2019.08.006

    32. [32]

      S.J. Hu, X.Q. Guo, L.P. Zhou, et al., J. Am. Chem. Soc. 144 (2022) 4244–4253. doi: 10.1021/jacs.2c00760

    33. [33]

      S.J. Hu, X.Q. Guo, L.P. Zhou, et al., Chin. J. Chem. 41 (2023) 797–804. doi: 10.1002/cjoc.202200649

    34. [34]

      B.S. Dolinar, D.I. Alexandropoulos, K.R. Vignesh, et al., J. Am. Chem. Soc. 140 (2018) 908–911. doi: 10.1021/jacs.7b12495

    35. [35]

      D.I. Alexandropoulos, B.S. Dolinar, K.R. Vignesh, et al., J. Am. Chem. Soc. 139 (2017) 11040–11043. doi: 10.1021/jacs.7b06925

    36. [36]

      X. Liu, Y. Zhang, W. Shi, et al., Inorg. Chem. 57 (2018) 13409–13414. doi: 10.1021/acs.inorgchem.8b01981

    37. [37]

      X. Meng, W. Shi, P. Cheng, Coord. Chem. Rev. 378 (2019) 134–150. doi: 10.1016/j.ccr.2018.02.002

    38. [38]

      E.T. Seo, R.F. Nelson, J.M. Fritsch, et al., J. Am. Chem. Soc. 88 (1966) 3498–3503. doi: 10.1021/ja00967a006

    39. [39]

      D.H. Reid, Tetrahedron 3 (1958) 339–352. doi: 10.1016/0040-4020(58)80039-3

    40. [40]

      S. Kumar, M.R. Ajayakumar, G. Hundal, et al., J. Am. Chem. Soc. 136 (2014) 12004–12010. doi: 10.1021/ja504903j

    41. [41]

      P.P. Power, Chem. Rev. 103 (2003) 789–810. doi: 10.1021/cr020406p

    42. [42]

      H. Garcia, H.D. Roth, Chem. Rev. 102 (2002) 3947–4007. doi: 10.1021/cr980026x

    43. [43]

      V. Pushkara Rao, M.B. Zimmt, N.J. Turro, J. Photochem. Photobio. A: Chem. 60 (1991) 355–360. doi: 10.1016/1010-6030(91)90037-T

    44. [44]

      H. Hu, Y.Y. Zhang, H. Ma, et al., Angew. Chem. Int. Ed. 62 (2023) e202308513. doi: 10.1002/anie.202308513

    45. [45]

      E. Breinlinger, A. Niemz, V.M. Rotello, J. Am. Chem. Soc. 117 (1995) 5379–5380. doi: 10.1021/ja00124a029

    46. [46]

      Q. Song, F. Li, Z. Wang, et al., Chem. Sci. 6 (2015) 3342–3346. doi: 10.1039/C5SC00862J

    47. [47]

      E. Moulin, F. Niess, M. Maaloum, et al., Angew. Chem. Int. Ed. 49 (2010) 6974–6978. doi: 10.1002/anie.201001833

    48. [48]

      Z. Feng, S. Tang, Y. Su, et al., Chem. Soc. Rev. 51 (2022) 5930–5973. doi: 10.1039/D2CS00288D

    49. [49]

      G. Huo, X. Shi, Q. Tu, et al., J. Am. Chem. Soc. 141 (2019) 16014–16023. doi: 10.1021/jacs.9b08149

    50. [50]

      Q. Tu, G.F. Huo, X.L. Zhao, et al., Mat. Chem. Front. 5 (2021) 1863–1871. doi: 10.1039/D0QM00992J

    51. [51]

      S.K. Zhang, L.Z. Ma, W.Q. Ma, et al., Angew. Chem. Int. Ed. 61 (2022) e202209054. doi: 10.1002/anie.202209054

    52. [52]

      K. Yazaki, S. Noda, Y. Tanaka, et al., Angew. Chem. Int. Ed. 55 (2016) 15031–15034. doi: 10.1002/anie.201608350

    53. [53]

      Z. Lu, R. Lavendomme, O. Burghaus, et al., Angew. Chem. Int. Ed. 58 (2019) 9073–9077. doi: 10.1002/anie.201903286

    54. [54]

      G.F. Jin, Y.Z. Zhang, L. Yu, et al., Nano Res. 16 (2023) 10678–10683. doi: 10.1007/s12274-023-5690-2

    55. [55]

      J. Wang, K. Liu, L. Ma, et al., Chem. Rev. 116 (2016) 14675–14725. doi: 10.1021/acs.chemrev.6b00432

    56. [56]

      G. Tan, X. Wang, Acc. Chem. Res. 50 (2017) 1997–2006. doi: 10.1021/acs.accounts.7b00229

    57. [57]

      E. Moulin, J.J.T. Armao, N. Giuseppone, Acc. Chem. Res. 52 (2019) 975–983. doi: 10.1021/acs.accounts.8b00536

    58. [58]

      L. Mao, M. Zhou, X. Shi, et al., Chin. Chem. Lett. 32 (2021) 3331–3341. doi: 10.1016/j.cclet.2021.05.004

    59. [59]

      B. Huang, L. Mao, X. Shi, et al., Chem. Sci. 12 (2021) 13648–13663. doi: 10.1039/D1SC01618K

    60. [60]

      W.L. Jiang, Z. Peng, B. Huang, et al., J. Am. Chem. Soc. 143 (2021) 433–441. doi: 10.1021/jacs.0c11738

    61. [61]

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

    62. [62]

      Ł. Skórka, J.M. Mouesca, J.B. Gosk, et al., J. Mater. Chem. C 5 (2017) 6563–6569. doi: 10.1039/C7TC01932G

    63. [63]

      M. Uebe, T. Kazama, R. Kurata, et al., Angew. Chem. Int. Ed. 56 (2017) 15712–15717. doi: 10.1002/anie.201709874

    64. [64]

      Y. Yokoyama, D. Sakamaki, A. Ito, et al., Angew. Chem. Int. Ed. 51 (2012) 9403–9406. doi: 10.1002/anie.201204106

    65. [65]

      X. Li, Y.L. Wang, C. Chen, et al., Chemistry 29 (2023) e202203242. doi: 10.1002/chem.202203242

    66. [66]

      X. Li, Y.L. Wang, C. Chen, et al., Nat. Commun. 13 (2022) 5367–5374. doi: 10.1038/s41467-022-33130-1

    67. [67]

      X.Q. Guo, L.P. Zhou, L.X. Cai, et al., Chem. Eur. J. 24 (2018) 6936–6940. doi: 10.1002/chem.201801132

    68. [68]

      L.L. Yan, C.H. Tan, L.P. Zhou, et al., J. Am. Chem. Soc. 137 (2015) 8550–8555. doi: 10.1021/jacs.5b03972

    69. [69]

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

    70. [70]

      W.Y. Sun, T. Kusukawa, M. Fujita, J. Am. Chem. Soc. 124 (2002) 11570–11571. doi: 10.1021/ja0206285

    71. [71]

      K. Sreenath, C.V. Suneesh, V.K. Kumar, et al., J. Org. Chem. 73 (2008) 3245–3251. doi: 10.1021/jo800349n

    72. [72]

      S. Zheng, S. Barlow, C. Risko, et al., J. Am. Chem. Soc. 128 (2006) 1812–1817. doi: 10.1021/ja0541534

    73. [73]

      Y. Satoh, L. Catti, M. Akita, et al., J. Am. Chem. Soc. 141 (2019) 12268–12273. doi: 10.1021/jacs.9b03419

    74. [74]

      V. Croue, S. Goeb, G. Szaloki, et al., Angew. Chem. Int. Ed. 55 (2016) 1746–1750. doi: 10.1002/anie.201509265

    75. [75]

      J.E. McPeak, S.S. Eaton, G.R. Eaton, Methods Enzymol. 651 (2021) 63–101.

  • Scheme 1  Synthetic route of TPA–bridging ligands L1, L2, L1•+ and self–assembly of Ln4(L)4 tetrahedral cages. (ⅰ) 6–(Isopropylcarbamoyl)picolinic acid, HATU, Et3N, 70% yield, (ⅱ) NaNO2/HCl, NaN3, (ⅲ) 6–ethynyl–N–isopropylpicolinamide, SA, CuSO4·5H2O, 75% yield, (ⅳ) NOBF4, 95% yield, (ⅴ) Ln(OTf)3, 99% yield.

    Figure 1  1H NMR of (a) ligand L1 in DMSO–d6 and (b) tetrahedral cage Eu4(L1)4(OTf)12 in CD3CN. (c) 1H DOSY spectrum of Eu4(L1)4(OTf)12. (d) ESI–TOF–MS of Eu4(L1)4(OTf)12 with insets showing the observed (Obs.) and simulated (Sim.) isotopic patterns for the 5+ peaks. Single crystal X–ray structures of (e) Eu4(L1)4 and (f) Eu4(L2)4. Eu: Cyan, C: green, N: blue, O: red, H: white.

    Figure 2  (a, b) Cyclic voltammetry of ligand L1, Eu4(L1)4 and L2, Eu4(L2)4 (1 mmol/L for ligands, 0.25 mmol/L for cages, 0.1 mol/L nBu4NPF6, scan rate 100 mV/s) at room temperature. (c) Frontier orbitals of L1 and L2 calculated using DFT methods (B3LYP/6–31G(d)).

    Figure 3  (a) Absorption spectra of L1 (0.1 mmol/L, blue dot line), L1•+ (0.1 mmol/L, blue line), Tb4(L1)4 (0.025 mmol/L, pink dot line), Tb4(L1•+)4 (0.025 mmol/L, pink line) in CH3CN, Inset: the color view of L1 and L1•+. (b) Excitation and emission spectra of Tb4(L1)4 (blue line) and Tb4(L1•+)4 (pink line) in CH3CN.

    Figure 4  Time–dependent UV–vis absorption spectra of (a) L1•+, (b) La4(L1•+)4 and (c) Tb4(L1•+)4 in CH3CN recorded at ambient air and light conditions. (d) Plot of the absorbance at 784 nm (L1•+), 760 nm (La4(L1•+)4), 740 nm (Tb4(L1•+)4) with time, showing the half–life times were increased by 9–fold and 15–fold for La4(L1•+)4, Tb4(L1•+)4, respectively.

    Figure 5  (a) EPR spectra of ligand L1 (light yellow line), L1•+ (yellow line, g = 2.003), La4(L1)4 (light blue line), La4(L1•+)4 (blue line, g = 2.003), Tb4(L1)4 (light pink line) and Tb4(L1•+)4 (pink line, g = 2.007) in acetonitrile solution at 293 K. (b) Temperature dependence of χmT for L1•+ (hexagon), Tb4(L1)4 (triangle) and Tb4(L1•+)4 (square) measured at 5000 Oe.

  • 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:  2024-01-15
  • 接受日期:  2024-02-07
  • 修回日期:  2024-01-30
  • 网络出版日期:  2024-02-13
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