Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Near-infrared emissive mononuclear lanthanide(III) complexes based on chiral Schiff base ligands: synthesis, crystal structure, luminescence, and magnetic properties

Yuri Jeonga, Ngoc Tram Anh Leb, Jeyun Jub, Iuliia Olshevskaiab, Daeheum Chob, Ryuya Tokunagac, Shinya Hayamic and Kil Sik Min*a
aDepartment of Chemistry Education and Science Education Research Institute, Kyungpook National University, Daegu 41566, Republic of Korea. E-mail: minks@knu.ac.kr
bDepartment of Chemistry, Kyungpook National University, Daegu 41566, Republic of Korea
cDepartment of Chemistry, Kumamoto University, Kumamoto 860-8555, Japan

Received 7th May 2025 , Accepted 30th August 2025

First published on 1st September 2025


Abstract

In this study, novel chiral mononuclear complexes (teaH)[Ln((R,R)-dnsalcd)2] (Ln = Nd3+ (1), Ho3+ (2), Er3+ (3), Yb3+ (4), and Gd3+ (5)), where teaH = triethylammonium and (R,R)-H2dnsalcd = (R,R)-N,N’-bis(3,5-dinitrosalicylidene)-1,2-cyclohexanediamine, were synthesized and characterized. Their structural features were analyzed via single-crystal diffraction, and their chiral and magnetic properties were examined through circular dichroism and magnetic susceptibility, respectively. Structural analysis revealed two distinct coordination modes: complexes 1 and 4 were isomorphous, each exhibiting a single intramolecular π–π stacking interaction, and complexes 3 and 5 were also isomorphous, each displaying two intramolecular π–π stacking interactions. Complexes 1 and 4 show strong near-infrared (NIR) emissions in both solid and solution states, attributed to efficient antenna effects. In contrast, complexes 2 and 3 exhibit weaker NIR emissions with Stark splitting features. Circular dichroism measurements confirmed the chiroptical activity of the complexes, and magnetic susceptibility data revealed typical lanthanide-type paramagnetic behavior. Density functional theory (DFT) calculations for complex 3 support a ligand-to-metal energy transfer mechanism, involving two sequential internal conversion processes.


Introduction

Lanthanide metal complexes exhibit unique properties, including sharp and long-lived emission peaks and large pseudo-Stokes shifts owing to the shielding of 4f electrons.1–3 Over the decades, extensive research has explored their multifunctional technological applications in bioimaging, sensing, light-emitting diodes, and single-molecule magnets.4–13 Among them, NIR-emitting lanthanide complexes can be used for applications in telecommunications, biosciences, and solar energy conversion.14–17 Generally, lanthanide complexes with Nd(III), Ho(III), Er(III), and Yb(III) ions can exhibit NIR emissions between 900 nm and 1600 nm.18,19 For instance, Pikramenou et al. reported Ln(tpOp)3 (Ln = Nd, Er, and Yb) complexes featuring a tetraphenyl imidodiphosphonate ligand (HtpOp).20 These complexes exhibit NIR luminescence with long lifetimes, ranging from 3.3 μs for Nd(tpOp)3 to 20 μs for Yb(tpOp)3. Spodine et al. reported three Er(III) hexaazamacrocylic complexes—Er-EDA (EDA = ethylenediamine), Er-OPDA (OPDA = ortho-phenylenediamine), and Er-DAP (DAP = 1,3-diaminopropane).21 Bluish-green and green emissions from the Er(III) ion upon ligand-centered excitation are observed only in complexes with macrocycles containing aliphatic spacers (Er-EDA and Er-DAP), whereas the complex with aromatic spacers (Er-OPDA) exhibits only ligand-based emission. This highlights the structural dependence of energy transfer from the ligand to the Er(III) emission levels.

Chiral metal complexes have been utilized as chiral catalysts and in chiral magnetism.22,23 Particularly, in the case of chiral lanthanide(III) complexes, they can be applied for circularly polarized luminescence (CPL), due to their strong emissions.24,25 Thus, it is quite important to design chiral ligands that can act as chiral luminophores. Recently, we have reported two mononuclear complexes (teaH)[Ln(dnsalcd)2] with chiral ligands (S,S/R,R)-N,N’-bis(3,5-dinitrosalicylidene)-1,2-cyclohexanediamine ((S,S/R,R)-H2dnsalcd), which exhibit unusual luminescence properties depending on the molecular geometry (Ln = Eu3+, Tb3+, teaH = triethylammonium).26 The Eu(III) complex shows highly efficient red emission, whereas the Tb(III) complex displays no visible emission owing to differences in π–π stacking interactions. We have found that the photoluminescence properties of the lanthanide(III) complexes can be effectively tuned by the coordination modes of the dnsalcd2− ligands.

In this context, our research has focused on the development of new lanthanide(III) complexes capable of exhibiting near-infrared (NIR) emissions. The chiral dnsalcd2− ligand demonstrates a significant antenna effect, influenced by its coordination geometry—specifically, the energy transfer from the ligand to the lanthanide(III) ion and potential self-quenching within the ligand. Based on our preliminary findings, we anticipate that NIR emissions can be both induced and controlled in lanthanide(III) complexes formed through the reaction of chiral dnsalcd2− ligands with lanthanide(III) ions, owing to the antenna effect of the coordinated ligands. Herein, the synthesis, crystal structure, circular dichroism, magnetic and luminescence properties of (teaH)[LnIII((R,R)-dnsalcd)2] (Ln = Nd (1), Ho (2), Er (3), Yb (4), and Gd (5)) are reported (Chart 1). Additionally, DFT calculations were performed to elucidate the emission mechanism.


image file: d5dt01071c-c1.tif
Chart 1 Structure of (R,R)-N,N’-bis(3,5-dinitrosalicylidene)-1,2-cyclohexanediamine in a tetradentate configuration.

Experimental

General

All chemicals used in this study were of reagent grade and used without further purification. (R,R)-N,N’-Bis(3,5-dinitrosalicylidene)-1,2-cyclohexanediamine ((R,R)-H2dnsalcd) was synthesized as previously described.26,27 Lanthanide salts containing nitrate or chloride anion were purchased from Sigma-Aldrich. Infrared (IR) spectra were measured using KBr pellets on a Thermo Fisher Scientific Nicolet iS5 spectrophotometer (±1 cm−1). UV-vis absorption spectra were recorded using a SCINCO S-3100 spectrophotometer (solution and diffuse reflectance mode). X-ray powder diffraction patterns were measured using a 4 kW Empyrean instrument (Panalytical, Netherlands) with degree increments and time steps of 0.02° and 0.2 s per step at 2θ and room temperature. Elemental analyses were performed using a Fisons/Carlo Erba EA1108 instrument in the air. 1H NMR spectra were measured with a Bruker AVANCE III 500 spectrometer, while circular dichroism (CD) spectra were recorded using a Jasco 1500 spectropolarimeter. Temperature-dependent magnetic susceptibilities were measured using a Quantum Design MPMS-5S superconducting quantum interference device (SQUID) magnetometer in a 5000 Oe applied field from 2 to 300 K at a 2 K min−1 sweep rate. Diamagnetic corrections were made using Pascal's constants.28 Luminescence spectra were recorded using a SCINCO FS-2 fluorescence spectrometer or a Horiba Fluorolog-3 spectrofluorometer equipped with a photomultiplier tube (PMT) detector. Luminescence spectra were recorded at multiple excitation wavelengths, which were selected based on the strongest UV-vis absorption peaks of each sample. Density functional theory (DFT) calculations were performed at the M062X theory level using the SARC2-ZORA-QZVP basis set for the Er(III) atom and the ZORA-def2-TZVP basis set for all other atoms.29–33

Syntheses of the compounds

(teaH)[NdIII((R,R)-dnsalcd)2] (1). A 5 mL dimethoxyethane solution of (R,R)-H2dnsalcd (69 mg, 0.136 mmol) was mixed with triethylamine (28 mg, 0.272 mmol) at room temperature and stirred for 15 min, yielding a yellow solution. A methanol solution (2 mL) of neodymium(III) nitrate hydrate (30 mg, 0.068 mmol) was then added dropwise to the yellow solution and stirred for 10 min. The resulting mixture was filtered, and yellow diamond-shaped crystals were obtained through diethyl ether diffusion, collected via filtration, washed with diethyl ether, and air-dried. Yield: 63 mg (68%). Anal. calcd for C46H48N13NdO20: C, 44.30; H, 3.88; N, 14.60. Found: C, 44.61; H, 4.12; N, 14.18. FT-IR (KBr, cm−1): 3084, 2936, 2861, 1632, 1596, 1561, 1529, 1328, 1099. UV/vis (in MeCN), λmax (ε): 234 nm (7.3 × 104 M−1 cm−1), 352 nm (7.7 × 104 M−1 cm−1). UV/Vis (diffuse reflectance spectrum), λmax: 200–550 nm (broad), 586 nm.
(teaH)[HoIII((R,R)-dnsalcd)2] (2). A 20 mL ethanol solution of (R,R)-H2dnsalcd (116 mg, 0.232 mmol) was mixed with triethylamine (47 mg, 0.46 mmol) at room temperature and stirred for 15 min, yielding a yellow solution. A 5 mL ethanol solution of holmium(III) chloride hydrate (44 mg, 0.12 mmol) was then added, and the mixture was stirred for 10 min. The resulting yellow precipitate was then filtered, washed with ethanol, and air-dried. Yield: 136 mg (84%). Yellow diamond-shaped crystals of 2 suitable for X-ray crystallography were obtained by diffusing diethyl ether into a solution containing (R,R)-H2dnsalcd and triethylamine in 1,2-dimethoxyethane and holmium(III) chloride hydrate in methanol for 3 days. Anal. calcd for C46H48HoN13O20: C, 43.58; H, 3.82; N, 13.36. Found: C, 43.46; H, 3.88; N, 13.63. IR (KBr, cm−1): 3078, 2933, 2856, 1632, 1596, 1567, 1528, 1330, 1099. UV/vis (in MeCN), λmax (ε): 234 nm (8.6 × 104 M−1 cm−1), 352 nm (9.0 × 104 M−1 cm−1). UV/vis (diffuse reflectance spectrum), λmax: broad absorption (200–500 nm), peaks at 537 nm and 650 nm.

Synthesis of (teaH)[MIII((R,R)-dnsalcd)2] (M = Er (3), Yb (4), and Gd (5))

A synthetic procedure similar to that of 2 was followed for the isolation of complexes 3–5, but holmium(III) chloride hydrate was replaced with the corresponding lanthanide precursors erbium(III) nitrate hydrate (51 mg, 0.12 mmol), ytterbium(III) nitrate hydrate (44 mg, 0.12 mmol), and gadolinium(III) nitrate hydrate (52 mg, 0.12 mmol). All crystals of 3–5 suitable for X-ray crystallography were obtained by diffusing diethyl ether into a mixed solution of (R,R)-H2dnsalcd and triethylamine in 1,2-dimethoxyethane and each metal nitrate hydrate in methanol for 3 days.

For 3: Yield: 116 mg (74%). Anal. calcd for C46H48ErN13O20: C, 43.50; H, 3.81; N, 14.34. Found: C, 43.65; H, 3.77; N, 14.19. IR (KBr, cm−1): 3085, 2935, 2860, 1634, 1596, 1588, 1529, 1332, 1101. UV/vis (in MeCN), λmax (ε): 234 nm (8.6 × 104 M−1 cm−1), 352 nm (8.8 × 104 M−1 cm−1). UV/vis (diffuse reflectance spectrum), λmax: 200–500 nm (broad), 519 nm, 659 nm.

For 4: Yield: 113 mg (71%). Anal. calcd for C46H48N13O20Yb: C, 43.30; H, 3.79; N, 14.27. Found: C, 43.58; H, 3.54; N, 13.90. IR (KBr, cm−1): 3080, 2935, 2857, 1633, 1596, 1567, 1529, 1331, 1100. UV/vis (in MeCN), λmax (ε): 234 nm (7.0 × 104 M−1 cm−1), 352 nm (7.2 × 104 M−1 cm−1). UV/vis (diffuse reflectance spectrum), λmax: 200–600 nm (broad).

For 5: Yield: 101 mg (70%). Anal. calcd for C46H48GdN13O20: C, 43.84; H, 3.84; N, 14.45. Found: C, 43.59; H, 3.90; N, 14.20. IR (KBr, cm−1): 3068, 2935, 2859, 1634, 1596, 1568, 1527, 1333, 1102. UV/vis (in MeCN), λmax (ε): 234 nm (8.6 × 104 M−1 cm−1), 351 nm (9.1 × 104 M−1 cm−1). UV/vis (diffuse reflectance spectrum), λmax: 200–600 nm (broad).

X-ray crystallographic data collection and refinement

Single crystals of 1–5 were mounted on a CryoLoop® with Paratone® oil. Intensity data were collected for all complexes at 223(2) K on the Bruker D8 VENTURE diffractometer equipped with a microfocus Mo-target X-ray tube (λ = 0.71073 Å) and a PHOTON 100 CMOS detector (Korea Basic Science Institute, Western Seoul Center). The data of 1–5 were processed using the Bruker APEX3 program,34 while data reduction and integration were performed using the Bruker SAINT program (v8.40A) and absorption-corrected with SADABS.35 Structure solutions were performed with direct methods using SHELXT-2014/536 and refined with a full-matrix least-squares approach in the SHELXL-2019/3 computer program.37 Unfortunately, the structure of 2 has been omitted from this paper because of a significant issue in the Fobs versus Fcalc plot, despite the fact that its crystal structure could be identified. The refinements of 3–5 were carried out using selected data cuts at 2θ = 46°, 43°, and 45°, respectively, to eliminate poor reflections and low-resolution data at higher 2θ angles. Due to the disorder observed in 1 and 3–5, the structures were refined using disorder models. To achieve satisfactory refinement of the disordered benzene rings and NO2 groups, a combination of constraints (EADP) and restraints (SAME, SADI, SIMU, DFIX, DANG, ISOR, and RIGU) was applied. In particular, refinement of the cation site in structure 4 was carried out using the PLATON squeeze option due to positional disorder.38 The positions of all non-hydrogen atoms were refined with anisotropic displacement factors. All hydrogen atom positions were constrained to their parent atoms using the appropriate HFIX command in SHELXL-2019/3 and a riding model. Additionally, we have added the plots of Rmerge and I/σ versus resolution of 1 and 3–5 (Fig. S1). Table 1 summarizes the crystallographic data and refinement results of the five complexes.
Table 1 Summary of the crystallographic data for 1 and 3–5
a R1 = ∑||Fo| − |Fc||/∑|Fo|.b wR2 = [∑w(Fo2Fc2)2/∑w(Fo2)2]1/2.
Compound 1 3 4 5
Empirical formula C51H62N13NdO22 C46H48ErN13O20 C46H48N13O20Yb C46H48GdN13O20
Formula weight 1353.37 1270.23 1276.01 1260.22
Crystal system Orthorhombic Hexagonal Tetragonal Hexagonal
Space group P212121 P61 P43212 P61
a (Å) 13.5169(12) 13.3689(16) 12.5403(9) 13.424(4)
b (Å) 19.006(2) 13.3689(16) 12.5403(9) 13.424(4)
c (Å) 22.585(3) 49.867(11) 35.322(4) 49.95(2)
α (°) 90 90 90 90
β (°) 90 90 90 90
γ (°) 90 120 90 120
V3) 5802.2(11) 7718(2) 5554.6(10) 7795(6)
Z 4 6 4 6
dcalc (g cm−3) 1.549 1.640 1.526 1.611
λ (Å) 0.71073 0.71073 0.71073 0.71073
T (K) 223(2) 223(2) 223(2) 223(2)
μ (mm−1) 0.985 1.723 1.769 1.367
F(000) 2780 3858 2580 3834
Reflections collected 42[thin space (1/6-em)]223 19[thin space (1/6-em)]248 11[thin space (1/6-em)]757 20[thin space (1/6-em)]845
Independent reflections 14[thin space (1/6-em)]285 6651 3188 6632
Reflections with I > 2σ(I) 12[thin space (1/6-em)]210 4837 2432 4646
Goodness-of-fit on F2 1.017 1.004 1.012 0.987
Flack's parameter −0.014(6) −0.006(14) −0.002(13) −0.019(18)
R1[thin space (1/6-em)]a [I > 2σ(I)] 0.0379 0.0596 0.0533 0.0590
wR2[thin space (1/6-em)]b [I > 2σ(I)] 0.0654 0.0987 0.1182 0.1063
CCDC 2448369 2448371 2448372 2448373


Results and discussion

Synthesis and characterization

Mononuclear chiral complexes (teaH)[MIII((R,R)-dnsalcd)2] (M = Nd (1); Ho (2); Er (3); Yb (4); Gd (5)) were obtained as yellow crystalline solids in good yields (68–84%) through the reaction of lanthanide salts with a tetradentate N2O2-type ligand. The synthesis involved mixing a lanthanide(III) salt, (R,R)-H2dnsalcd, and triethylamine in a 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]4 molar ratio in ethanol (dimethoxyethane/methanol for 1). For complex 1, single crystals of (teaH)[Nd((R,R)-dnsalcd)2] were obtained and utilized for all characterization processes. In contrast, solids 2–5 were prepared as precipitates, with their structures confirmed through powder X-ray diffraction (Fig. S2–S5). The IR spectrum of 1 exhibited the characteristic peaks of azomethine imine bonds at 1632 cm−1, which were downshifted by ca. 22 cm−1 for the free (R,R)-H2dnsalcd ligand (Fig. S6).39 Additionally, peaks corresponding to benzene double bonds were downshifted to ca. 1596 cm−1 owing to coordination with Nd(III) ion.26 The aromatic and aliphatic C–H bands of (R,R)-dnsalcd2− and teaH+ cations appear at ca. 3084, 2936, and 2861 cm−1, respectively. Moreover, the band at 1328 cm−1 corresponds to the symmetric stretching vibrations of the nitro groups in (R,R)-dnsalcd2−. Similarly, the IR spectra of 2–5 exhibited similar patterns, with slight wavenumber differences owing to coordination with different lanthanide(III) metal ions. Fig. S7 shows the UV-Vis spectra for the (R,R)-H2dnsalcd ligand and complexes 1–5 in the solid state at room temperature. The spectra display broad absorption bands between 200 and 600 nm. Additionally, complexes 1–3 showed characteristic absorption peaks of lanthanide metal ions consistent with a previous report.40 For complex 1, the sharp peak was observed at 586 nm owing to the transition from the ground state 4I9/2 to an emitting level 4G5/2. Complex 2 displayed two absorption lines at 537 and 650 nm, attributed to the 4f–4f transitions of 5I85F4 and 5F5, respectively. Similarly, complex 3 showed two absorption bands at 519 and 649 nm, corresponding to 4I15/22H11/2 and 4F9/2, respectively. In contrast, complexes 4 and 5 did not show any observable f–f transitions owing to strong ligand absorption.41,42 However, their UV-Vis absorption spectra in solution have been measured in acetonitrile (Fig. S8). Two sharp absorption peaks were observed at 233 and 352(1) nm, as previously reported.26 These peaks correspond to the π → π* transitions of the phenolate and azomethine groups, respectively.

To examine the structural characteristics of optically active chiral compounds, CD spectra for the (R,R)-H2dnsalcd and complexes 1–5 were obtained in an acetonitrile solution at room temperature (Fig. 1). The CD spectrum of (R,R)-H2dnsalcd exhibited two positive Cotton effects at 270 and 390 nm and two negative Cotton effects at 250 and 434 nm. Upon ligand complexation with Ln(III) ions, the CD absorption peak at 434 nm, corresponding to charge transfer from the benzene to the nitro group, disappeared, as confirmed via the UV-Vis spectra (Fig. S8). The negative band observed at 271 and 276 nm for complexes 1 and 5, respectively, indicates the π → π* transition of the ligand phenolic group. However, for complexes 2–4, it splits into 246 nm and 283(1) nm, owing to the Stark effect (Fig. 1). Additionally, the positive and negative Cotton effects at 318–327 and 362–382 nm, respectively, were attributed to the π → π* transition of the azomethine groups in the (R,R)-dnsalcd2− ligand.26


image file: d5dt01071c-f1.tif
Fig. 1 CD spectra of (a) (R,R)-H2dnsalcd, (b) Nd (1), (c) Ho (2), (d) Er (3), (e) Yb (4), and (f) Gd (5) in MeCN solutions at room temperature.

Description of the crystal structures

Structure of 1. Complex 1 crystallizes in the orthorhombic P212121 space group. Fig. 2 shows the oak ridge thermal ellipsoid plot (ORTEP) representation of 1, and Table 2 lists the selected bond lengths and angles. The asymmetric unit comprises a crystallographically independent Nd(III) complex anion with a teaH+ cation and lattice solvent molecules (i.e., one CH3OH and one CH3CH2OCH2CH3). The Nd(III) cation is coordinated by four phenoxide ions and four nitrogen atoms from two (R,R)-dnsalcd2− ligands, forming an N4O4 coordination environment with a distorted square-antiprismatic geometry (Table S1). The average Nd–N and Nd–O bond lengths were 2.627(2) and 2.377(2) Å, respectively. These differences reflect the different radii and strains of N and O.43 Complex 1 showed an intramolecular π–π stacking interaction between the two phenoxide groups (O2a/O2b) of two (R,R)-dnsalcd2− ligands. This is attributed to the severely distorted bonding of each (R,R)-dnsalcd2− ligand to the Nd(III) core.26 Furthermore, the centroid–centroid distance was 3.570 Å, with a dihedral angle of 17.5(2)° between the benzene rings, indicating slight distortion in the π–π stacking interaction (Fig. S9).44
image file: d5dt01071c-f2.tif
Fig. 2 ORTEP representation of complex 1, with atoms represented by 50% probable thermal ellipsoids. All hydrogen atoms and solvent molecules are omitted for clarity, except the hydrogen atom on the nitrogen of the triethylammonium cation.
Table 2 Selected bond distances (Å) and angles (°) for 1 and 3–5
Compound 1 3 4 5
         
Ln–N1a 2.599(4) 2.518(16) 2.525(13) 2.610(15)
Ln–N2a 2.649(3) 2.537(14) 2.471(13) 2.538(15)
Ln–N1b 2.582(4) 2.556(15) 2.525(13) 2.530(13)
Ln–N2b 2.662(4) 2.496(13) 2.471(13) 2.630(16)
Ln–O1a 2.367(3) 2.272(12) 2.254(10) 2.306(12)
Ln–O2a 2.390(3) 2.281(13) 2.225(11) 2.299(13)
Ln–O1b 2.395(3) 2.228(12) 2.254(10) 2.387(13)
Ln–O2b 2.356(3) 2.319(13) 2.225(11) 2.279(13)
         
O1a–Ln–O1b 116.53(11) 95.5(4) 104.4(6) 94.0(5)
O1a–Ln–O2b 81.12(12) 91.1(5) 84.7(4) 92.7(5)
O1a–Ln–O2a 152.66(11) 149.5(4) 152.3(4) 151.2(5)
O1a–Ln–N1a 68.51(11) 71.6(5) 70.0(4) 71.1(5)
O1a–Ln–N2a 126.06(11) 138.6(5) 133.8(4) 138.1(5)
O1a–Ln–N1b 82.79(12) 73.6(4) 73.0(4) 75.5(5)
O1a–Ln–N2b 81.19(11) 79.3(5) 76.0(4) 82.3(5)
O1b–Ln–O2a 87.40(11) 92.5(4) 84.7(4) 90.4(5)
O1b–Ln–O2b 154.99(13) 150.7(4) 152.3(4) 153.1(4)
O1b–Ln–N1a 84.94(12) 76.1(5) 73.0(4) 77.7(4)
O1b–Ln–N2a 80.79(11) 79.4(4) 76.0(4) 81.0(4)
O1b–Ln–N1b 69.83(12) 72.8(4) 70.0(4) 69.8(5)
O1b–Ln–N2b 128.06(11) 137.8(5) 133.8(4) 135.0(5)
N1a–Ln–N2a 62.50(12) 67.3(5) 66.1(4) 67.2(5)
N1a–Ln–N1b 127.81(12) 130.0(5) 117.6(6) 130.9(5)
N1a–Ln–N2b 143.52(12) 137.8(5) 141.7(4) 139.9(5)
N2a–Ln–N1b 146.48(12) 139.8(5) 141.7(4) 137.8(5)
N2a–Ln–N2b 129.81(11) 130.7(4) 136.9(7) 129.5(5)
N1b–Ln–N2b 64.45(12) 65.5(4) 66.1(4) 65.9(5)


Structure of 3. Complex 3 crystallized in the hexagonal P61 space group, and Fig. 3 shows its ORTEP representation, while Table 2 shows the selected bond lengths and angles. The asymmetric unit was composed of one Er(III) cation, two (R,R)-dnsalcd2− ligands, and one teaH+ cation, which served to balance the charge of the complex anion. That is, the central Er(III) ion was eight-coordinated, adopting a distorted square-antiprismatic geometry (Table S1). The average Er–N and Er–O bond distances were determined to be 2.525(7) and 2.273(6) Å, respectively. These differences were attributed to the different sizes and strains of N and O atoms.43 Complex 3 showed two intramolecular π–π stacking interactions between the two phenoxide groups (O1a/O1b and O2a/O2b) of the two (R,R)-dnsalcd2− ligands. This can be attributed to each (R,R)-dnsalcd2− ligand binding to the Er(III) center in an overlapping mode, adopting a staggered conformation with a dihedral angle of 88.7(3)° through the coordinated N and O atoms.26 Moreover, the centroid–centroid distances were found to be 3.486 and 3.508 Å, while the dihedral angles between the benzene rings were 7.6(5)° and 7.2(1.1)°, indicating slight distortions in the π–π stacking interactions (Fig. S9).44
image file: d5dt01071c-f3.tif
Fig. 3 ORTEP representation of complex 3. The atoms are represented by 30% probable thermal ellipsoids. All hydrogen atoms are omitted for clarity, except for the hydrogen atom on the nitrogen of the triethylammonium cation.
Structure of 4. Complex 4 crystallized in the tetragonal P43212 space group, and Fig. 4 shows its ORTEP representation, while Table 2 shows the selected bond lengths and angles. The asymmetric unit contained one crystallographically independent Yb(III) complex cation, present as a half complex, along with a half teaH+ cation. Thus, the complex adopted a distorted square-antiprismatic geometry with a coordination number of 8, imposed by the symmetry operation. Although the triethylammonium cation was identified in the crystal structure, it was treated using the squeezing option owing to disorder. In the structure, the Yb(III) cation was coordinated to four phenoxide ions and four nitrogen atoms from two (R,R)-dnsalcd2− ligands, resulting in an N4O4 coordination environment and a distorted square-antiprismatic geometry (Table S1). The average Yb–N and Yb–O bond distances were determined to be 2.498(7) and 2.241(5) Å, respectively. Likewise, these differences were related to the different atomic radii and strains of N and O atoms.43 Complex 4 showed one intramolecular π–π stacking interaction between the two phenoxide groups (O1a/O1b) of the two (R,R)-dnsalcd2− ligands, similar to what was observed in complex 1. This can be attributed to each (R,R)-dnsalcd2− ligand binding to the Yb(III) center in a highly distorted mode.26 Moreover, the centroid–centroid distance was found to be 3.723 Å, with a dihedral angle of 3.4(7)° between the benzene rings, indicating a slight distortion in the π–π stacking interaction (Fig. S9).44 Each (R,R)-dnsalcd2− ligand was also coordinated to the Yb(III) ion in a square planar mode, adopting a staggered conformation with a dihedral angle of 89.3(3)° through the coordinated N and O atoms.
image file: d5dt01071c-f4.tif
Fig. 4 ORTEP representation of complex anions of 4. The atoms are represented by 20% probable thermal ellipsoids. All hydrogen atoms are omitted for clarity.
Structure of 5. Complex 5 was isostructural with complex 3, implying that it shared the same crystal system, space group, and asymmetric unit. Complex 5 crystallized in the hexagonal P61 space group, and Fig. 5 shows its ORTEP representation, while Table 2 shows the selected bond lengths and angles. The asymmetric unit was composed of one Gd(III) cation, two (R,R)-dnsalcd2− ligands, and one teaH+ cation, which served to balance the charge of complex anions. That is, the central Gd(III) ion was eight-coordinated, adopting a distorted square-antiprismatic geometry (Table S1). The average Gd–N and Gd–O bond distances were determined to be 2.572(7) and 2.317(6) Å, respectively. These differences were attributed to the different sizes and stains of the corresponding N and O atoms.43 Complex 5 showed two intramolecular π–π stacking interactions between the two phenoxide groups (O1a/O1b and O2a/O2b) of the two (R,R)-dnsalcd2− ligands. This can be attributed to each (R,R)-dnsalcd2− ligand binding to the Gd(III) center in an overlapping mode, adopting a staggered conformation with a dihedral angle of 88.4(3)° through the coordinated N and O atoms.26 Moreover, the centroid–centroid distances were measured at 3.493 and 3.558 Å, while the dihedral angles between the benzene rings were 6.6(1.0)° and 6.9(0.8)°, indicating slight distortions in the π–π stacking interactions (Fig. S9).44
image file: d5dt01071c-f5.tif
Fig. 5 ORTEP representation of complex 5. The atoms are represented by 30% probable thermal ellipsoids. All hydrogen atoms are omitted for clarity, except for the hydrogen atom on the nitrogen of the triethylammonium cation.

Photophysical properties

The photoluminescence properties of the complexes, along with those of the (R,R)-H2dnsalcd ligand, were investigated in the solid and acetonitrile solution states at room temperature. Fig. 6 shows the emission spectra in the solid state, while Fig. 7 shows the observed f–f transitions of complexes 1–5. For the pure chiral ligand, broad emission was detected between 450 and 600 nm, with a maximum at approximately 500 nm under 390 nm excitation.26 The chiral ligand can generate chiral metal complexes, which may potentially induce interesting circularly polarized luminescence. Upon excitation at 380 nm, complex 1 exhibited characteristic NIR emissions, corresponding to the 4F3/24I9/2 (891 and 915 nm), 4F3/24I11/2 (1059 nm), and 4F3/24I13/2 (1330 and 1379 nm) transitions.45 The 4F3/24I9/2 and 4F3/24I13/2 emission bands in 1 split into two peaks, probably attributed to the Stark splitting effect.46 Similar splitting patterns have been previously reported,47 which are attributed to the splitting of the emitting levels induced by ligand field effects.48 For complex 2, the emission spectrum showed two transition bands, attributed to the 5F55I8 (659 nm) and 5F55I7 (985 and 1057 nm) transitions upon excitation at 380 nm. For the 5F55I7 transition, the peak exhibited a splitting pattern, which can again be attributed to the Stark splitting effect.47 In the spectrum of complex 3, excitation at 380 nm led to split emission peaks at 1497, 1527, and 1552 nm, along with a very weak band at 968 nm. These peaks corresponded to the 4I11/24I15/2 (968 nm) and 4I13/24I15/2 (1497, 1527, and 1552 nm) transitions. For complexes 2 and 3, the starred peaks were observed through the long pass filter at 400 nm owing to their low intensity. These complexes further exhibited weak ligand-centered emission in the 400–700 nm spectral range, indicating incomplete energy transfer from the ligand to the excited states of Ln(III) ions.49 The emission spectrum of complex 4, upon excitation at 380 nm, showed strong emission peaks at 977, 995, and 1017 nm, which can be attributed to the 2F5/22F7/2 transition, with Stark splitting effects.48 Unlike complexes 1–4, complex 5 exhibited only a weak, broad emission band at the 400–500 nm range, with a maximum at 465 nm, which was attributed to the ligand π → π* electronic transition. This phenomenon can be attributed to Gd(III) lacking visible photoluminescence, as its first excited state (6P7/2) lies ∼32[thin space (1/6-em)]000 cm−1 above its ground state.50
image file: d5dt01071c-f6.tif
Fig. 6 Emission spectra of (a) (R,R)-H2dnsalcd (λex = 390 nm) and (b) complexes 1 (λex = 380 nm), (c) 2 (λex = 380 nm), (d) 3 (λex = 380 nm), (e) 4 (λex = 380 nm), and (f) 5 (λex = 400 nm) in the solid state at room temperature. * indicates Rayleigh scattering peaks.

image file: d5dt01071c-f7.tif
Fig. 7 Energy illustration of the observed f–f transitions in the emission spectra of complexes 1–5 in the solid state.51

Fig. 8 shows the emission spectra of complexes 1–4 in the MeCN solution. No emission was observed for the chiral Schiff base ligand and complex 5, owing to solvent-induced quenching, which resulted from ligand–solvent interactions.52 For complex 1, the emission spectrum showed the same pattern as the solid state, exhibiting three characteristic NIR emission peaks: 4F3/24I9/2 (887 and 918 nm), 4F3/24I11/2 (1060 and 1083 nm), and 4F3/24I13/2 (1331 nm) transitions. For complex 2, three sets of emission bands were observed again, though with significantly reduced intensities compared to those observed in the solid state. The Stark splitting emission peaks at 612, 657, and 701 nm were attributed to 5F55I8 transitions, while the broad splitting peaks at 862 and 980 nm corresponded to the 5F55I7 transition.47b A weak band was also observed at 1186 nm, corresponding to the 5I65I8 transition. Complex 3 exhibited a characteristic 4I13/24I15/2 transition band of the Er(III) ion at 1527 nm, with a shoulder at ca. 1561 nm with a lower intensity than that in the solid state. The weak 4I11/24I15/2 transition band disappeared in the solution state. Conversely, the emission spectrum of complex 4 exhibited an intense peak in the 900–1100 nm range, with a maximum at 995 nm, corresponding to the 2F5/22F7/2 transition.48 This behavior demonstrates that the (R,R)-dnsalcd2− ligand effectively sensitizes Nd(III) and Yb(III) ions through an efficient antenna effect in the solid state and MeCN solution. These complexes are comparable to those of Eu(III) complexes reported by our group.26 Moreover, NIR emissions were observed in the Ho(III) and Er(III) complexes, which are quite unusual compared to those with the Tb(III) complexes.26 In the case of the Tb(III) complexes, the presence of two intramolecular π–π stacking interactions appears to quench or hinder the energy transfer process, leading to suppressed lanthanide-centered emission. Therefore, it was initially expected that the Ho(III) and Er(III) complexes would not exhibit energy transfer from the ligand to the Ln(III) ions due to the presence of two intramolecular π–π stacking interactions. Unexpectedly, however, NIR emissions were observed in both complexes, albeit not intense. These results suggest that, in contrast to the Tb(III) complexes, the two intramolecular π–π stacking interactions in the Ho(III) and Er(III) complexes allow for a limited but measurable ligand-to-metal energy transfer. Additionally, since the NIR emissions of complexes 1–4 are observed in both the solid and solution states, intermolecular interactions can be excluded as the primary contributors to these emissions. In the case of complexes 1 and 4, each of which exhibits a single intramolecular π–π stacking interaction, strong metal-centered emissions are observed, attributed to efficient ligand-to-metal energy transfer (i.e., the antenna effect) without significant quenching. These findings indicate that the NIR emissions in complexes 1–4 are highly dependent on the number of intramolecular π–π stacking interactions.


image file: d5dt01071c-f8.tif
Fig. 8 Emission spectra of (a) complexes 1 (λex = 380 nm), (b) 2 (λex = 380 nm), (c) 3 (λex = 380 nm), and (d) 4 (λex = 380 nm) in MeCN solution at room temperature.

Magnetic properties

The magnetic susceptibility data of 1–5 were obtained on solid-state samples under an applied field of 5000 Oe over a temperature range of 2–300 K using a SQUID magnetometer. Fig. 9 shows the resulting χMT versus T plot. At room temperature, complex 1 exhibited a χMT value of 1.53 emu K mol−1, slightly lower than the theoretical value expected for a single, non-interacting Nd3+ ion (gJ = 8/11, calculated χMT = 1.64 emu K mol−1). This difference can be attributed to the ground-state term symbol for the Nd(III) ion, 4I, which splits into four different J states (4I9/2–15/2) owing to spin–orbit coupling.53 For complex 2, the χMT value at 300 K was 13.9 emu K mol−1, slightly lower than that of the expected value for the 5I8 ground state of the Ho3+ ion (gJ = 5/4, calculated χMT = 14.07 emu K mol−1). Upon cooling, χMT remained relatively stable until reaching 50 K, after which it sharply reduced to 3.69 emu K mol−1 at 2 K. This reduction can be attributed to the progressive depopulation of excited Stark sublevels influenced by ligand field effects, indicating the presence of significant magnetic anisotropy, a phenomenon frequently observed in lanthanide compounds.54 For complex 3, the χMT value at 300 K was 11.3 emu K mol−1, slightly lower than the expected value for the 4I15/2 ground state of the Er3+ ion (gJ = 6/5, calculated χMT = 11.48 emu K mol−1). This difference can be attributed to the ground-state term symbol for the Er(III) ion, 4I, which undergoes spin–orbit coupling, splitting into four different J states (4I9/2–15/2). Upon cooling, χMT remained relatively constant until reaching 80 K, after which it sharply reduced to 5.24 emu K mol−1 at 2 K. For complex 4, the χMT value at 300 K was 2.52 emu K mol−1, slightly lower than the expected value for the 2F7/2 ground state of the Yb3+ ion (gJ = 8/7, calculated χMT = 2.57 emu K mol−1). This deviation can be attributed to the ground-state term symbol for the Yb(III) ion, 2F, which undergoes spin–orbit coupling splitting into two different J states (2F5/2–7/2).53 For complex 5, the χMT value at 300 K was 8.06 emu K mol−1, slightly higher than that of the calculated value for a single non-interacting Gd3+ ion (gJ = 7/2, calculated χMT = 7.87 emu K mol−1). The Gd3+ ion exhibited an 8S7/2 ground state (4f,7 J = 7/2, L = 0, and S = 7/2) with no contribution from orbital angular momentum and was located approximately 104 cm−1 below the first excited state.55 Upon cooling, χMT remained constant until reaching 20 K, after which it slightly reduced to 7.04 emu K mol−1 at 2 K. For the entire temperature range, the χM−1(T) versus T plot adhered to the Curie–Weiss law, with θ = −0.28 K (C = 8.06 emu K mol−1) for complex 5. This is most probably attributed to antiferromagnetic interactions between the mononuclear species.
image file: d5dt01071c-f9.tif
Fig. 9 Magnetic susceptibility data for complexes 1–5, displayed as the χMT product.

Density functional theory calculations

The emission band of complex 3 was observed at 1497, 1527, and 1552 nm upon excitation at λex = 380 nm (Fig. 7), corresponding to the 4I13/24I15/2 transition. Thus, the DFT calculations were performed to investigate the electronic structure of the Er(III) complex. All calculations were performed using the ORCA program employing the M06-2X functional with the SARC2-ZORA-QZVP basis set for the Er atom and the ZORA-def2-TZVP basis set for all other atoms.32,33 The calculation result showed that energy transfer occurs from the ligand to the Er(III) ion. The transition densities between the ground and excited states confirm that ligand-to-Er(III) metal energy transfer facilitates the Er(III)-centered emission. These findings suggest that the 1527 nm emission (4I13/24I15/2 transition) probably occurs from a series of internal conversions from the populated 2P3/2 level to the long-lived 4I13/2 level (Fig. 10).56 This mechanism differs significantly from the emission behavior observed in the Tb(III) complex.26 In that case, the presence of two intramolecular π–π stacking interactions results in strong confinement of the π-electrons in the L+1 orbital (where L refers to the lowest unoccupied molecular orbital), thereby suppressing ligand-to-metal charge transfer upon excitation. In contrast, for complex 3, density functional theory (DFT) calculations revealed that the transition densities of the (R,R)-dnsalcd2− ligands are effectively transferred to the Er(III) ion within the complex anion, which facilitates NIR emissions.
image file: d5dt01071c-f10.tif
Fig. 10 Schematic representation of the energy transfer from the ligands to the metal ion in the Er(III) complex. The energies relative to the ground 4I15/2 state are given in parentheses.

Conclusions

The synthesis and characterization of five enantiopure mononuclear complexes (teaH) [Ln((R,R)-dnsalcd)2] are reported (Ln = Nd, Ho, Er, Yb, and Gd). Within this series, the chiral tetradentate ligand is coordinated to lanthanide ions in a bis-chelating mode while exhibiting two different coordination modes related to π–π interactions between benzene groups. The Nd(III) and Yb(III) complexes feature a single intramolecular π–π interaction between benzene rings, whereas the Ho(III), Er(III), and Gd(III) complexes exhibit two intramolecular π–π interactions within their complex anions. The Nd(III) and Yb(III) complexes in the solid and solution states exhibited strong NIR emissions corresponding to each metal ion within the near-infrared range, facilitated by an efficient antenna effect. In contrast, the Ho(III) and Er(III) complexes showed weak NIR emissions corresponding to each metal ion. From these results, we confirmed that the two intramolecular π–π interactions weaken the energy transfer more effectively than a single π–π interaction. Nevertheless, despite the presence of strong π–π interactions in the Ho(III) and Er(III) complexes, weak NIR emissions were still observed, contrary to previous findings. The Gd(III) complex exhibited only ligand-based emission in the solid state, whereas no emission was observed in the solution. DFT calculations confirmed that the NIR emissions in the Er(III) complex were driven by the antenna effect, which was associated with internal conversion during energy transfer from the ligands to the metal ion. Moreover, all lanthanide complexes showed paramagnetic behavior at room temperature. Upon cooling, their magnetic moments were influenced by the thermal depopulation of their respective sublevels.

Author contributions

Yuri Jeong: data curation, analysis, and writing the manuscript draft. Ngoc Tram Anh Le: data curation, analysis, and writing the manuscript draft. Jeyun Ju: DFT calculations. Iuliia Olshevskaia: DFT calculations. Daeheum Cho: DFT calculations and writing the manuscript draft. Ryuya Tokunaga: data curation and analysis. Shinya Hayami: data curation and analysis. Kil Sik Min: supervision, conceptualization, project design, and writing the manuscript draft.

Conflicts of interest

There are no conflicts to declare.

Data availability

The data supporting this article have been included as part of the SI: compound spectroscopic characterisation (XRPD, IR, UV-Vis) and SC-XRD collection and refinement details. See DOI: https://doi.org/10.1039/d5dt01071c.

CCDC 2448369 (1) and 2448371–2448373 (3–5) contain the supplementary crystallographic data for this paper.57a–d

Acknowledgements

This research was supported by the Basic Science Research Program (No. 2021R1F1A1062108) and the framework of the International Cooperation Program (No. 2021K2A9A2A08000136) through the National Research Foundation of Korea (NRF). We thank Dr Jong Won Shin and Dr Huiyeong Ju for analyzing the X-ray crystal data.

References

  1. F. Wang and X. Liu, Acc. Chem. Res., 2014, 47, 1378–1385 CrossRef CAS PubMed.
  2. N. Kiseleva, P. Nazari, C. Dee, D. Busko, B. S. Richards, M. Seitz and A. Turshatov, J. Phys. Chem. Lett., 2020, 11, 2477–2481 CrossRef CAS PubMed.
  3. T. Ahmed, A. Chakraborty, A. Paul and S. Baitalik, Dalton Trans., 2023, 52, 14027–14038 RSC.
  4. A. J. Amoroso and S. J. Pope, Chem. Soc. Rev., 2015, 44, 4723–4742 RSC.
  5. M. Sy, A. Nonat, N. Hildebrandt and L. J. Charbonnière, Chem. Commun., 2016, 52, 5080–5095 RSC.
  6. X. Qiu, J. Xu, M. Cardoso Dos Santos and N. Hildebrandt, Acc. Chem. Res., 2022, 55, 551–564 CrossRef CAS.
  7. D. Parker, Coord. Chem. Rev., 2000, 205, 109–130 CrossRef CAS.
  8. D. Parker, J. D. Fradgley and K. L. Wong, Chem. Soc. Rev., 2021, 50, 8193–8213 RSC.
  9. A. de Bettencourt-Dias, Dalton Trans., 2007, 22, 2229–2241 RSC.
  10. L. Wang, Z. Zhao, C. Wei, H. Wei, Z. Liu, B. Bian and C. Huang, Adv. Opt. Mater., 2019, 7, 1801256 CrossRef.
  11. A. Dey, P. Kalita and V. Chandrasekhar, ACS Omega, 2018, 3, 9462–9475 CrossRef CAS PubMed.
  12. D. N. Woodruff, R. E. P. Winpenny and R. A. Layfield, Chem. Rev., 2013, 113, 5110–5148 CrossRef CAS.
  13. R. J. Blagg, L. Ungur, F. Tuna, J. Speak, P. Comar, D. Collison, W. Wernsdorfer, E. J. L. McInnes, L. F. Chibotaru and R. E. P. Winpenny, Nat. Chem., 2013, 5, 673–678 CrossRef CAS PubMed.
  14. J. C. G. Bünzli, S. Comby, A. S. Chauvin and C. D. Vandevyver, J. Rare Earths, 2007, 25, 257–274 CrossRef.
  15. H. Suzuki, J. Photochem. Photobiol., A, 2004, 166, 155–161 CrossRef CAS.
  16. J. C. G. Bünzli, Chem. Rev., 2010, 110, 2729–2755 CrossRef.
  17. B. M. Van Der Ende, L. Aarts and A. Meijerink, Phys. Chem. Chem. Phys., 2009, 11, 11081–11095 RSC.
  18. Y. Ning, Y.-W. Liu, Y.-S. Meng and J.-L. Zhang, Inorg. Chem., 2018, 57, 1332–1341 CrossRef CAS PubMed.
  19. T. N. Nguyen, G. Capano, A. Gładysiak, F. M. Ebrahim, S. V. Eliseeva, A. Chidambaram, B. Valizadeh, S. Petoud, B. Smit and K. C. Stylianou, Chem. Commun., 2018, 54, 6816–6819 RSC.
  20. D. Davis, A. J. Carrod, Z. Guo, B. M. Kariuki, Y.-Z. Zhang and Z. Pikramenou, Inorg. Chem., 2019, 58, 13268–13275 CrossRef CAS PubMed.
  21. Y. Gil, R. C. de Santana, A. S. S. de Camargo, L. G. Merízio, P. F. Carreño, P. Fuentealba, J. Manzur and E. Spodine, Dalton Trans., 2023, 52, 3158–3168 RSC.
  22. A. Fanourakis and R. J. Phipps, Chem. Sci., 2023, 14, 12447–12476 RSC.
  23. (a) F.-X. Xu, Y.-L. Li, X.-Q. Wei, D. Shao, L. Shi, H.-Y. Wei and X.-Y. Wang, Dalton Trans., 2023, 52, 5575–5586 RSC; (b) Z. Zhu, C. Zhao, T. Feng, X. Liu, X. Ying, X.-L. Li, Y.-Q. Zhang and J. Tang, J. Am. Chem. Soc., 2021, 143, 10077–10082 CrossRef CAS.
  24. M. Tsurui, R. Takizawa, Y. Kitagawa, M. Wang, M. Kobayashi, T. Taketsugu and Y. Hasegawa, Angew. Chem., Int. Ed., 2024, 63, e202405584 CrossRef CAS PubMed.
  25. (a) J. Liu, W. Song, H. Niu, Y. Lu, H. Yang, W. Li, Y.-z. Zhao and Z. Miao, Inorg. Chem., 2024, 63, 18429–18437 CrossRef CAS PubMed; (b) Y. Tang, M. Jian, B. Tang, Z. Zhu, Z. Wang and Y. Liu, Inorg. Chem. Front., 2024, 11, 2039–2048 RSC; (c) T. Feng, R. Cai, Z. Zhu, Q. Zhou, A. Sickinger, O. Maury, Y. Guyot, A. Bensalah-Ledoux, S. Guy, B. Baguenard, B. Le Guennic and J. Tang, Chem. – Eur. J., 2025, 31, e202500910 CrossRef CAS.
  26. S. Jeon, Y. Jeong, L. N. T. Anh, J. Ju, D. Cho, Y. J. Jang, R. Tokunaga, S. Hayami and K. S. Min, J. Ind. Eng. Chem., 2023, 128, 275–285 CrossRef CAS.
  27. G. Shen, F. Gou, J. Cheng, X. Zhang, X. Zhou and H. Xiang, RSC Adv., 2017, 7, 40640–40649 RSC.
  28. G. A. Bain and J. F. Berry, J. Chem. Educ., 2008, 85, 532–536 CrossRef CAS.
  29. A. Verma, M. K. Tiwari, N. Subba and S. Saha, J. Photochem. Photobiol., A, 2022, 433, 114130 CrossRef.
  30. E. G. Hohenstein, S. T. Chill and C. D. Sherrill, J. Chem. Theory Comput., 2008, 4, 1996–2000 CrossRef CAS PubMed.
  31. Z. Zhou, J. McNeely, J. Greenough, Z. Wei, H. Han, M. Rouzières, A. Y. Rogachev, R. Clérac and M. A. Petrukhina, Chem. Sci., 2022, 13, 3864–3874 RSC.
  32. D. Aravena, F. Neese and D. A. Pantazis, J. Chem. Theory Comput., 2016, 12, 1148–1156 CrossRef CAS PubMed.
  33. F. Weigend and R. Ahlrichs, Phys. Chem. Chem. Phys., 2005, 7, 3297–3305 RSC.
  34. APEX3, v2019.1-0, Bruker, Madison, WI, USA, 2019 Search PubMed.
  35. (a) SADABS, version 2016/2, Bruker-AXS, Madison, WI, USA, 2016 Search PubMed; (b) L. Krause, R. H. Irmer, G. M. Sheldrick and D. Stalke, J. Appl. Crystallogr., 2015, 48, 3–10 CrossRef CAS PubMed.
  36. G. M. Sheldrick, SHELXT, Universität Göttingen, Göttingen, Germany, 2014 Search PubMed.
  37. G. M. Sheldrick, Acta Crystallogr., Sect. A:Found. Adv., 2015, C71, 3–8 CrossRef.
  38. (a) P. van der Sluis and A. L. Spek, Acta Crystallogr., Sect. A:Found. Crystallogr., 1990, A46, 194–201 CrossRef CAS; (b) A. L. Spek, Acta Crystallogr., Sect. A:Found. Adv., 2015, C71, 9–18 Search PubMed.
  39. A. S. Smirnov, A. S. Kritchenkov, N. A. Bokach, M. L. Kuznetsov, S. I. Selivanov, V. V. Gurzhiy, A. Roodt and V. Y. Kukushkin, Inorg. Chem., 2015, 54, 11018–11030 CrossRef CAS.
  40. S. K. Sharma, T. Behm, T. Köhler, J. Beyer, R. Gloaguen and J. Heitmann, Crystals, 2020, 10, 593 CrossRef CAS.
  41. K. Binnemans, R. V. Deun, C. Görller-Walrand, S. R. Collinson, F. Martin, D. W. Bruce and C. Wickleder, Phys. Chem. Chem. Phys., 2000, 2, 3753–3757 RSC.
  42. I. Pospieszna-Markiewicz, M. A. Fik-Jaskółka, Z. Hnatejko, V. Patroniak and M. Kubicki, Molecules, 2022, 27, 8390 CrossRef CAS PubMed.
  43. J. C. Slater, J. Chem. Phys., 1964, 41, 3199–3204 CrossRef CAS.
  44. (a) M. Mantina, A. C. Chamberlin, R. Valero, C. J. Cramer and D. G. Truhlar, J. Phys. Chem. A, 2009, 113, 5806 CrossRef CAS PubMed; (b) A. R. Jeong, J. W. Shin, J. H. Jeong, S. Jeoung, H. R. Moon, S. Kang and K. S. Min, Inorg. Chem., 2020, 59, 15987 CrossRef CAS.
  45. Y. Hasegawa, Y. Wada and S. Yanagida, J. Photochem. Photobiol., C, 2004, 5, 183–202 CrossRef CAS.
  46. (a) L. Zhang, T. Xue, D. He, M. Guzik and G. Boulon, Opt. Express, 2015, 23, 1505–1511 CrossRef CAS PubMed; (b) B. Yang, X. Liu, X. Wang, J. Zhang, L. Hu and L. Zhang, Opt. Lett., 2014, 39, 1772–1774 CrossRef CAS PubMed.
  47. Z. Li, J. Yu, L. Zhou, R. Deng and H. Zhang, Inorg. Chem. Commun., 2009, 12, 151–153 CrossRef CAS.
  48. (a) J. A. Jiménez, R. Amesimenu and M. Thomas, J. Phys. Chem. B, 2024, 128, 2995–3003 CrossRef; (b) M. A. Katkova, A. P. Pushkarev, T. V. Balashova, A. N. Konev, G. K. Fukin, S. Y. Ketkov and M. N. Bochkarev, J. Mater. Chem., 2011, 21, 16611–16620 RSC.
  49. (a) I. Pospieszna-Markiewicz, M. A. Fik-Jaskółka, Z. Hnatejko, V. Patroniak and M. Kubicki, Molecules, 2022, 27, 8390 CrossRef CAS PubMed; (b) J. Choo, A. R. Jeong, H. Yeo, S. Hayami and K. S. Min, J. Mol. Struct., 2020, 1206, 127726 CrossRef CAS.
  50. J. F. Greisch, M. E. Harding, B. Schäfer, M. Ruben, W. Klopper, M. M. Kappes and D. Schooss, J. Phys. Chem. Lett., 2014, 5, 1727–1731 CrossRef CAS PubMed.
  51. J.-C. G. Bünzli and C. Piguet, Chem. Soc. Rev., 2005, 34, 1048–1077 RSC.
  52. T. J. Penfold, S. Karlsson, G. Capano, F. A. Lima, J. Rittmann, M. Reinhard, M. H. Rittmann-Frank, O. Braem, E. Baranoff, R. Abela, I. Tavernelli, U. Rothlisberger, C. J. Milne and M. Chergui, J. Phys. Chem. A, 2013, 117, 4591–4601 CrossRef CAS PubMed.
  53. E. R. Martí, A. B. Canaj, T. Sharma, A. Celmina, C. Wilson, G. Rajaraman and M. Murrie, Inorg. Chem., 2022, 61, 9906–9917 CrossRef.
  54. P. Wytrych, J. Utko, M. Stefanski, J. Kłak, T. Lis and Ł. John, Inorg. Chem., 2023, 62, 2913–2923 CrossRef CAS PubMed.
  55. R. C. A. Vaz, I. O. Esteves, W. X. C. Oliveira, J. Honorato, F. T. Martins, L. F. Marques, G. L. dos Santos, R. O. Freire, L. T. Jesus, E. F. Pedroso, W. C. Nunes, M. Julve and C. L. M. Pereira, Dalton Trans., 2020, 49, 16106–16124 RSC.
  56. A. Upadhyay, K. R. Vignesh, C. Das, S. K. Singh, G. Rajaraman and M. Shanmugam, Inorg. Chem., 2017, 56, 14260–14274 CrossRef CAS PubMed.
  57. (a) Y. Jeong, N. T. A. Le, J. Ju, I. Olshevskaia, D. Cho, R. Tokunaga, S. Hayami and K. S. Min, CCDC 2448369: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n5qn7; (b) Y. Jeong, N. T. A. Le, J. Ju, I. Olshevskaia, D. Cho, R. Tokunaga, S. Hayami and K. S. Min, CCDC 2448371: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n5qq9; (c) Y. Jeong, N. T. A. Le, J. Ju, I. Olshevskaia, D. Cho, R. Tokunaga, S. Hayami and K. S. Min, CCDC 2448372: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n5qrb; (d) Y. Jeong, N. T. A. Le, J. Ju, I. Olshevskaia, D. Cho, R. Tokunaga, S. Hayami and K. S. Min, CCDC 2448373: Experimental Crystal Structure Determination, 2025,  DOI:10.5517/ccdc.csd.cc2n5qsc.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.