Shunichiro Itoab,
Kazuo Tanaka
*ab and
Yoshiki Chujoa
aDepartment of Polymer Chemistry, Graduate School of Engineering, Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan
bDepartment of Technology and Ecology, Graduate School of Global Environmental Studies, Kyoto University, Katsura, Nishikyo-ku, Kyoto 615-8510, Japan
First published on 19th May 2025
A solid-state luminescent lithium β-diketiminate complex was synthesized by the reaction of the corresponding proligand and n-butyllithium in hexane. The complex showed a low-coordinated structure in the solution and crystalline states. On the other hand, the lithium complex reacted with diethyl ether, tetrahydrofuran, and 1,4-dioxane to form three-coordinated complexes, which showed lower emission efficiency. Remarkably, the photoluminescence of the crystal of the low-coordinate species was quenched by the treatment with the diethyl ether vapor. NMR spectroscopy and X-ray diffraction analyses revealed that diethyl ether reacts with the complex to afford the three-coordinated ether adduct even in the crystalline state. To the best of our knowledge, this is the first report on the luminescence switching in crystalline materials based on the modification of the coordination number on a lithium atom.
In this context, π-extended β-diketiminate ligands are intriguing candidates for constructing luminescent lithium complexes because their group 13 metal complexes exhibit aggregation-induced emission (AIE) and crystallization-induced emission (CIE) properties.29–39 These solid-state luminescent properties and environment sensitiveness are advantageous for constructing film-type optical sensors and detection devices for tiny environmental changes and weak external stimuli. Therefore, we envisioned that stimuli-responsive materials might be obtained based on the strong Lewis acidity of base-free lithium complexes with sterically encumbered β-diketiminates. Herein, we report the preparation and characterization of a luminescent base-free lithium β-diketiminate complex and the switching of solid-state luminescence properties by coordination with various ethers (Chart 1, bottom). The base-free complex exhibited efficient luminescence in the crystalline state, while the ether adducts showed lower emission efficiencies. Interestingly, the photoluminescence in the crystalline state was dramatically quenched by the vapor fuming with diethyl ether. From the results of single-crystal X-ray diffraction analyses, it was found that diethyl ether coordinates to the lithium atom to give the three-coordinated species even in the crystalline state.
![]() | ||
Scheme 1 Synthetic scheme of the base-free lithium complex LLi and the ether adduct LLi·OEt2 (Dipp = 2,6-diisopropylphenyl). |
LLi, LLi·OEt2, LLi·THF, and LLi·DOX formed crystals belonging to the monoclinic P21/n, triclinic P, triclinic P
, and tetragonal P42/n space groups, respectively (Table S1†). The crystallographic parameters of LLi·OEt2 agreed with the reported data.40 In all crystals, each asymmetric unit included a single crystallographically independent complex molecule. Only for the case of LLi·DOX, its asymmetric unit involved an additional free DOX molecule. The 1H NMR spectrum of LLi·DOX after drying in vacuo showed a 1
:
1 molar ratio of the complex and DOX, indicating that the free DOX molecules in the crystal can be removed. The crystal structures of the complexes are shown in Fig. 1, and the selected structural parameters are listed in Table 1. Notably, in the crystalline state of LLi, the lithium cation was not coordinated to any solvent molecules and instead interacted with the carbon atoms of the phenyl ring in a neighboring molecule with the formation of the polymeric packing structure (Li⋯C7 = 2.670(4), Li⋯C8 = 2.572(4) Å). These Li⋯C interactions are weaker than the reported data in [Li(Dipp)NCH2CH2N(Dipp)Li]2 (2.153(10) Å)41 and stronger than those in the related β-diketiminate complex, Dipp2nacnacLi (2.834(7) Å).15 The six-membered LiN2C3 ring of the complexes was nearly planar with the slight deviation of one of the nitrogen atoms from the least-squares plane of the other five atoms (LLi, 0.2657(26) Å; LLi·OEt2, 0.1550(34) Å; LLi·THF, 0.1296(26) Å; LLi·DOX, 0.1004(32) Å). In addition, no apparent bond alteration was observed within the N2C3 moiety, indicating the delocalization of π-electrons. Interestingly, the average Li–N bond length increased from 1.901(3) and 1.881(3) Å for LLi to 1.919(4) and 1.929(4) Å for LLi·OEt2, 1.903(3) and 1.910(3) Å for LLi·THF, and 1.902(4) and 1.918(4) Å for LLi·DOX. This result suggests that the coordination of diethyl ether weakens the Li–N bonds. This weakening effect results in a more acute coordination angle for the ether adducts (98.4(2)°) than for LLi (99.4(1)°), LLi·THF (98.8(1)°), and LLi·DOX (98.5(2)°).
LLi | LLi·OEt2 | LLi·THF | LLi·DOX | ||||
---|---|---|---|---|---|---|---|
Bond | |||||||
d/Å | d/Å | d/Å | d/Å | ||||
Li–N1 | 1.901(3) | Li–N1 | 1.919(4) | Li–N1 | 1.903(3) | Li–N1 | 1.902(4) |
Li–N2 | 1.881(3) | Li–N2 | 1.929(4) | Li–N2 | 1.910(3) | Li–N2 | 1.918(4) |
N1–C1 | 1.329(3) | N1–C1 | 1.329(3) | N1–C1 | 1.326(2) | N1–C1 | 1.324(2) |
N2–C3 | 1.321(3) | N2–C3 | 1.330(3) | N2–C3 | 1.326(2) | N2–C3 | 1.324(2) |
C1–C2 | 1.411(2) | C1–C2 | 1.413(3) | C1–C2 | 1.409(2) | C1–C2 | 1.411(3) |
C2–C3 | 1.413(2) | C2–C3 | 1.419(3) | C2–C3 | 1.415(2) | C2–C3 | 1.415(3) |
O–Li | 1.929(5) | O–Li | 1.899(4) | O–Li | 1.918(4) |
Angle | |||||||
---|---|---|---|---|---|---|---|
θ/° | θ/° | θ/° | θ/° | ||||
N1–Li–N2 | 99.4(1) | N1–Li–N2 | 98.4(2) | N1–Li–N2 | 98.8(1) | N1–Li–N2 | 98.5(2) |
O–Li–N1 | 129.7(2) | O–Li–N1 | 132.1(2) | O–Li–N1 | 128.8(2) | ||
O–Li–N2 | 129.8(2) | O–Li–N2 | 127.1(2) | O–Li–N2 | 130.7(2) |
Photophysical properties of the synthesized complexes were evaluated with UV–vis absorption and photoluminescence (PL) measurements (Fig. 2a–d and Table 2). All measurements were carried out under a N2 atmosphere because of the instability of the complexes under aerobic conditions. All complexes showed similar absorption spectra peaked at around 365 nm. Additionally, emission from their dilute solution (2-methylpentane (2MP)/toluene = 99/1, 1 × 10−5 M) was not detectable at room temperature and even in the frozen solution at 77 K. The absolute luminescence quantum yields were not determined due to small values below the detection limit (<0.01). On the other hand, intense green PL was observed from crystalline LLi even at room temperature (ΦPL = 0.65, Fig. 2e). This result clearly indicates that LLi possesses the CIE property, similarly to the related β-diketiminate complexes containing the group 13 elements.29–39 The molecular motions in the excited states, like intramolecular vibrations34 and structural deformations,37 could open the channel for the nonradiative decay of the excited molecules in solution states, while the rigid crystalline packing could annihilate such nonradiative decay processes. Remarkably, in contrast to the base-free complex, the ether adducts exhibited only slight enhancement of the emission by crystallization (ΦPL = 0.03 for LLi·OEt2, 0.03 for LLi·THF, and 0.08 for LLi·DOX, Fig. 2e).
Temperature | Solutionb | Crystal | ||
---|---|---|---|---|
λabs/nm | λPL/nm | ΦPL![]() |
||
a PL spectra and absolute quantum yields were recorded with the photoexcitation at 360 nm.b 2MP/toluene (99/1, v/v) solution (1 × 10−5 M).c Not determined.d Absolute PL quantum yield determined with the integration sphere method. | ||||
LLi | r.t. | 364 | 527 | 0.65 |
77 K | —c | 521 | 0.90 | |
LLi·OEt2 | r.t. | 363 | 527 | 0.03 |
77 K | —c | 543 | 0.14 | |
LLi·THF | r.t. | 366 | 572 | 0.03 |
77K | —c | 594 | 0.06 | |
LLi·DOX | r.t. | 364 | 539 | 0.08 |
77K | —c | 535 | 0.32 |
In order to gain further insights into the photophysical processes of the complexes, luminescence lifetimes of the crystals of LLi and LLi·OEt2 were measured by the time-correlated single photon counting (TCSPC) measurement (Fig. 2f and Table 3). The nanosecond-order lifetimes strongly suggested that the luminescence from both crystals should be assigned to fluorescence. As shown in Table 3, the emission decay of LLi includes the relatively long components (1.71 and 3.81 ns), whereas the luminescence from LLi·OEt2 decays with the sub-nanosecond lifetimes (0.26 and 0.65 ns). The nonradiative decay rate constant (knr) of LLi·OEt2 was 17 times larger than that of LLi, while the radiative decay rate constant (kr) of the ether adduct was less than one-third of that of the base-free complex (Fig. 2g and Table 3). These results indicate that significantly fast nonradiative decay processes hamper the efficient luminescence of LLi·OEt2, and such processes are absent in the crystals of LLi. The molecular motions of the coordinated ether molecules could be responsible for such quenching paths.
τ1/ns (f1![]() |
τ2/ns (f2![]() |
χ2 | 〈τ〉c/ns | kr![]() |
knr![]() |
|
---|---|---|---|---|---|---|
a Excited at 375 nm with an LED laser. Decay curves were fitted with the multi-component exponential function: ![]() ![]() ![]() |
||||||
LLi | 1.71 (8.26%) | 3.81 (91.74%) | 1.17 | 3.63 | 1.8 | 1.0 |
LLi·OEt2 | 0.26 (20.70%) | 0.65 (79.30%) | 1.22 | 0.57 | 0.5 | 17 |
To investigate the electronic nature of the complexes, density functional theory (DFT) calculations were carried out. Geometries of the compounds were optimized at the CAM-B3LYP/6-31G(d,p) level of theory. The optimized structures were confirmed as local minima by performing frequency calculations at the same level. The calculated energy diagrams and the Kohn–Sham (KS) frontier orbital distributions of the complexes are shown in Fig. 3. The energies of the KS highest occupied molecular orbitals (HOMOs) were hardly affected by the ether. The KS lowest unoccupied molecular orbital (LUMO) of LLi, the next LUMO (LUMO+1) of LLi·OEt2 and LLi·THF, and the LUMO of LLi·DOX assigned to the LUMO of the N2C3 moiety were located in almost the same energy region. On the other hand, the LUMO+1 of LLi, which was composed mainly of the vacant 2p orbital of the lithium atom, was significantly stabilized by the complexation owing to the bonding interaction between the vacant orbital and the lone pair of ether. Additionally, the LUMO of LLi·OEt2 and LLi·THF and LUMO+1 of LLi·DOX, which were assigned to the antibonding orbital in the ether moiety, seemed to be degenerated with the LUMO+1 because there is almost no interaction between the π* orbital of the ligand and the σ* orbital of the ether.
The S0–S1 electronic transitions were calculated with time-dependent DFT (TD-DFT) at the CAM-B3LYP/6-311++G(d,p) level of theory and the results are shown in Fig. 3 and Table 4. The S1 states of all complexes possessed the same character of the π–π* transition mainly located at the N2C3 moiety. Consequently, the transition energy was not disturbed independently of whether the lithium atom was coordinated by the ethers or not. These results are consistent with the experimental observation of the UV–vis absorption spectra in which the complexes exhibit almost identical absorption bands. It is assumed that an increase in the degree of molecular motions might cause the luminescence quenching in the ether adducts. Indeed, the crystal structures of the complexes imply that the Li–N coordination could be weakened by the ether coordination. On the other hand, TD-DFT calculations for the monomeric and dimeric structures in the LLi crystal suggested that the relatively weak Li-π interactions in the crystal should not significantly affect the electronic transition properties (Table S12 and Fig. S1†).
Energyb/eV | λ/nm | fc | Composition | Coefficientd | |
---|---|---|---|---|---|
a Calculated at the TD-CAM-B3LYP/6-311++G(d,p) level of theory.b Excitation energies between S0 and S1 states.c Oscillator strength.d Configuration interaction coefficient of the component for the S0–S1 transition. | |||||
LLi | 3.70 | 336 | 0.6707 | HOMO → LUMO | 0.68589 |
LLi·OEt2 | 3.77 | 329 | 0.6281 | HOMO → LUMO+1 | 0.67493 |
LLi·THF | 3.74 | 332 | 0.6171 | HOMO → LUMO+1 | 0.67896 |
LLi·DOX | 3.77 | 329 | 0.6266 | HOMO → LUMO | 0.68099 |
To test whether the luminescence of the crystals of LLi was switched to that of LLi·OEt2 even in the solid state, the crystalline powder of LLi was exposed to the vapor of diethyl ether in a glovebox. In a 50 mL sealed vial containing 3 mL of diethyl ether, a 5 mL open vial equipped with the crystalline powder of LLi was placed and subsequently stored at room temperature for 12 h. After the treatment, we confirmed that the emission from the crystals was critically weakened. The PL spectrum and the absolute PL quantum yield of the treated sample (0.02) were almost identical to those of the crystals of LLi·OEt2 (Fig. 4a). Powder X-ray diffraction (PXRD) patterns for the samples before and after treatment were identical with the simulated patterns of LLi and LLi·OEt2, respectively, indicating that a crystal–crystal transition occurred by the ether vapor treatment (Fig. 4b). The 1H NMR spectrum (Chart S5†) of the treated crystal after drying under vacuum showed the same peaks as those of LLi·OEt2, and there was no residual signal assigned to LLi, indicating that the complexation between LLi and Et2O is not reversed on simple vacuum drying. It can be said that the base-free lithium complex could be able to be utilized not only for the detection of Lewis bases but also for constructing other types of stimuli-responsive materials.
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2391300, 2391301, 2440658 and 2440659. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d5dt00954e |
This journal is © The Royal Society of Chemistry 2025 |