Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Apr 1.
Published in final edited form as: J Lumin. 2020 Dec 15;232:117835. doi: 10.1016/j.jlumin.2020.117835

Lanthanide nitrato complexes bridged by the bis-tridentate ligand 2,3,5,6-tetra(2-pyridyl)pyrazine: Syntheses, crystal structures, Hirshfeld surface analyses, luminescence properties, DFT calculations, and magnetic behavior

Li-Ling Cai a, Sheng-Mei Zhang a, Yan Li a,*, Kai Wang a, Xue-Ming Li a, Gilles Muller b,**, Fu-Pei Liang a,c, Ya-Tao Hu a, Gui-Xia Wang a
PMCID: PMC8460135  NIHMSID: NIHMS1675641  PMID: 34565833

Abstract

Six dinuclear lanthanide(III) nitrato complexes [Ln(NO3)3(H2O)]2(μ-tppz) (where tppz = 2,3,5,6-tetra(2-pyridyl) pyrazine and Ln(III) = Nd (1), Sm (2), Eu (3), Gd (4), Tb (5), and Dy (6)) with bis-tridentate N-heterocyclic 2,3,5,6-tetra(2-pyridyl)pyrazine as bridging ligand have been solvothermally synthesized and characterized via elemental analysis, infrared spectroscopy, thermogravimetric analysis, single-crystal X-ray diffraction, and powder X-ray diffraction. The 3-D Hirshfeld surface and 2-D fingerprint plots show that the main interactions in 16 are the O⋯H/H⋯O intermolecular interactions with relative contributions of about 62%. Although the poor lanthanide(III)-centered luminescence properties clearly point to the efficiency of nonradiative quenching processes (presence of water molecules in the coordination sphere of the lanthanide(III) ions), the ligand tppz is better suited to sensitize the lanthanide(III)’s emissions of EuIII and NdIII than SmIII, TbIII, and DyIII. Finally, the magnetic data of DyIII comple×6 reveals antiferromagnetic coupling between DyIII ions.

Keywords: Lanthanide complex; 2,3,5,6-tetra(2-pyridyl)pyrazine; Hirshfeld surface; Luminescence; DFT

1. Introduction

On account of their excellent luminescent properties such as sharp line-like and high color-pure emissions, long lifetimes, and large Stokes/anti-Stoke shifts, lanthanide(III) complexes with luminescence arising from ff transitions possesses diverse applications in the fields of lighting, telecommunications, road marking, laser materials, sensors for analytes, upconversion materials, bioimaging, immunoassays, medical diagnosis and therapy, luminescent lanthanide-based Single-Molecule Magnets (SMM), and so on [19]. To sensitize the Laporte-forbidden ff transitions, various organic chromophores are selected to coordinate to the lanthanide(III) ions by utilizing the “antenna effect” [1,5,8,10]. Among them, multi-dentate N-heterocyclic ligands have been shown excellent antenna-ligands to sensitize the luminescence of lanthanide (III) complexes, due to their fascinating electrochemical, photophysical and photochemical properties stemming from the conjugated aromatic cores [1114]. The π-conjugated N-heterocyclic ligand 2,3,5,6-tetra (2-pyridyl) pyrazine (tppz) possessing two trans-terpyridine coordination sites is one of the most fascinating molecules [1521]. Although the tppz ligand was first synthesized by Goodwin and co-workers in 1959 [22], little work on tppz-containing complexes appeared until 1993 [23]. Since then, numerous studies on the synthesis and application of the tppz-based complexes have been reported [2331]. Although most of them focused on complexes with transition metals, there has been a few cases with lanthanide(III) ions [15,17,3235]. In particular, [Cp*2Yb]2(μ-tppz) bridged by tppz was reported by Morris and John in 2003 and 2006 to explore electron transfer molecular wire complexes [32,33]. [(Cp*2Ln)2(μ-tppz)]+ and [(Cp*2Ln)2(μ-tppz)] (Ln(III) = Gd, Tb, and Dy) were reported by Long in 2014 to demonstrate how tppz can be used to synthesize a series of dinuclear lanthanide(III) complexes for which the redox-active ligand tppz existed in the monoanionic and tri-anionic forms [34]. Fukuzumi used tppz as a probe molecule and a host ligand for metal ions, enabling “OFF-OFF-ON” switch ability for metal ion-fluorescence sensors and metal ion-promoted electron transfer reactions [35]. More recently, tppz was used to construct highly luminescent mononuclear lanthanide(III) complexes [15]. Owing the capability of tppz to possess different modes of coordination towards transition metal ions and lanthanide(III) ions, the challenge work, consisting of construction of d-f hetero-dinuclear luminescent complexes with tppz as π-conjugated bridge, is of great interest to us. To fulfill this goal, a series of lanthanide(III) tppz-containing complexes were firstly studied.

Herein, a series of dinuclear lanthanide(III) nitrato complexes, [Ln (NO3)3(H2O)]2(μ-tppz) (Ln(III) = Nd (1), Sm (2), Eu (3), Gd (4), Tb (5), and Dy (6)), bridged by the bis-tridentate N-heterocyclic ligand 2,3,5,6-tetra(2-pyridyl) pyrazine ligand, are reported. Complexes 1–6 were characterized by elemental analysis, infrared spectroscopy, thermogravimetric analysis, single-crystal X-ray diffraction, and powder X-ray diffraction. Of special importance, we investigated the near-infrared (1–2) and visible luminescent (2–6) properties for these solid-state compounds, whereas the Hirshfeld surface analysis was used to explore the interactions within the crystal structures of 1–6 for a deeper understanding of the chemical binding in these luminescent lanthanide (III) compounds. Furthermore, the HOMO–LUMO energy gaps of the structures of tppz, EuIII complex 3, GdIII complex 4, and TbIII complex 5 were revealed by employing the density functional theory (DFT) calculations. Finally, the magnetic behavior of the DyIII comple×6 was also studied.

2. Experimental

2.1. Materials and methods

Samarium nitrate hexahydrate (99.9%) and gadolinium nitrate hexahydrate (99.9%) were purchased from Amethyst Chemicals, whereas neodymium nitrate hexahydrate (99.9%), europium nitrate hexahydrate (99.9%), and terbium nitrate hexahydrate (99.9%) were purchased from Strem Chemicals. Dysprosium nitrate hexahydrate (99.9%) was purchased from Beijing HWRK Chem Ltd and 2,3,5,6-tetra (2-pyridyl)pyrazine (tppz) was synthesized according to the literature method [22].

[Nd(NO3)3(H2O)]2(μ-tppz) (1). 2,3,5,6-tetra(2-pyridyl)pyrazine (19.4 mg, 0.05 mmol) in 6 mL dichloromethane was added dropwise to a 3 mL acetonitrile solution containing Nd(NO3)3·6H2O (43.8 mg, 0.1 mmol). The mixture was added to a 25 mL teflon-lined stainless auto-clave and heated at 120 °C for 2 days. Pale yellow crystals of 1 suitable for X-ray crystallographic measurements were collected by filtration, washed twice with dichloromethane, and then dried in air. Yield: 50.6 mg, 93.3% based on Nd(NO3)3·6H2O. Anal. Calcd. for C24H20N12O20Nd2: C, 26.55; H, 1.84; N, 15.49%; Found: C, 26.44; H, 1.85; N, 15.48%. IR (KBr pellets, cm−1): 3350, 3095, 1634, 1598, 1503, 1385, 1281, 1157, 1024, 811, 785, 739.

[Sm(NO3)3(H2O)]2(μ-tppz) (2). For the synthesis of 2 the same procedure as that for 1 was employed, using the 2,3,5,6-tetra(2-pyridyl)pyrazine (19.4 mg, 0.05 mmol) and Sm(NO3)3·6H2O (44.4 mg, 0.1 mmol). Yield 50.9 mg, 92.8% based on Sm(NO3)3·6H2O. Calcd. for C24H20N12O20Sm2: C, 26.25; H, 1.82; N, 15.31%; Found: C, 26.35; H, 1.83; N,15.37%. IR (KBr pellets, cm−1): 3350, 3093, 1634, 1598, 1503, 1385, 1283, 1157, 1024, 812, 785, 738.

[Eu(NO3)3(H2O)]2(μ-tppz) (3). For the synthesis of 3 the same procedure as that for 1 was employed, using the 2,3,5,6-tetra(2-pyridyl) pyrazine (19.4 mg, 0.05 mmol) and Eu(NO3)3·6H2O (44.6 mg, 0.1 mmol). Yield 50.1 mg, 91.1% based on Eu(NO3)3·6H2O. Calcd. for C24H20N12O20Eu2: C, 26.17; H, 1.82; N, 15.27%; Found: C, 25.95; H, 1.83; N,15.10%. IR (KBr pellets, cm−1): 3356, 3095, 1635, 1598, 1505, 1386, 1286, 1158, 1025, 811, 785, 739.

[Gd(NO3)3(H2O)]2(μ-tppz) (4). For the synthesis of 4 the same procedure as that for 1 was employed, using the 2,3,5,6-tetra(2-pyridyl)pyrazine (19.4 mg, 0.05 mmol) and Gd(NO3)3·6H2O (45.1 mg, 0.1 mmol). Yield 50.8 mg, 91.4% based on Gd(NO3)3·6H2O. Calcd. for C24H20N12O20Gd2: C, 25.92; H, 1.80; N, 15.12%; Found: C, 26.05; H, 1.81; N,15.08%. IR (KBr pellets, cm−1): 3354, 3095, 1635, 1598, 1505, 1386, 1287, 1158, 1025, 812, 785, 741.

[Tb(NO3)3(H2O)]2(μ-tppz) (5). For the synthesis of 5 the same procedure as that for 1 was employed, using the 2,3,5,6-tetra(2-pyridyl)pyrazine (19.4 mg, 0.05 mmol) and Tb(NO3)3·6H2O (45.3 mg, 0.1 mmol). Yield 51.1 mg, 91.7% based on Tb(NO3)3·6H2O. Calcd. for C24H20N12O20Tb2: C, 25.84; H, 1.79; N, 15.08%; Found: C, 25.76; H, 1.80; N, 15.05%. IR (KBr pellets, cm−1): 3361, 3095, 1634, 1598, 1506, 1388, 1288, 1157, 1026, 811, 784, 742.

[Dy(NO3)3(H2O)]2(μ-tppz) (6). For the synthesis of 6 the same procedure as that for 1 was employed, using the 2,3,5,6-tetra(2-pyridyl)pyrazine (19.4 mg, 0.05 mmol) and Dy(NO3)3·6H2O (45.6 mg, 0.1 mmol). Yield 51.3 mg, 91.5% based on Dy(NO3)3·6H2O. Calcd. for C24H20N12O20Dy2: C, 25.68; H, 1.78; N, 14.98%; Found: C, 25.59; H, 1.79; N, 14.95%. IR (KBr pellets, cm−1): 3377, 3097, 1634, 1599, 1506, 1388, 1291, 1158, 1027, 811, 785, 745.

2.2. X-ray diffraction analysis

The diffraction data was collected at 293(2) K with an Agilent G8910A CCD diffractometer with graphite monochromated Mo-Kα radiation (λ = 0.71073 Å), using the ω-θ scan mode in the range 3.34° ≤ θ ≤ 25.01°. The structures were solved by direct methods and refined by full matrix least-squares on F2. An empirical absorption correction was applied with the program SADABS. All non-hydrogen atoms were refined anisotropically. All hydrogen atoms were added geometrically. Calculations and graphics were performed with SHELXL [36,37]. The crystallographic data and the refinement of structure parameters for 16 are summarized in Table S1. Selected bond lengths and angles for 16 are given in Table S2.

2.3. Other measurements

Elemental analyses (C, H, N) were performed with a vario MICRO cube elemental analyzer. FT–IR spectra were recorded from KBr pellets with a Bio-Rad FTS-7 spectrophotometer in the range of 4000–400 cm−1. Thermogravimetric analyses (TGA) were run on a SDT-Q600 comprehensive thermal analyzer under N2 atmosphere with a flow rate of 100 mL min−1, a heating rate of 10 °C·min−1, and a temperature range of 30–900 °C. Powder X-ray diffraction (PXRD) data were obtained on a PANalytical X’Pert3 power diffractometer with a PIXcel detector. The X-ray generator was operated at 40 kV and 40 mA. The Cu Kα line at 1.54056 Å was used as the radiation source. Samples were scanned from 5° to 50° (2θ) in stage sizes of 0.02626° with a scanning speed of 0.1347°/s. Hirshfeld surface analysis was performed using the CrystalExplorer program [38,39]. Solid-state luminescence properties of tppz and complexes 16 were recorded in the solid state at room temperature on an Edinburgh FLS920 lifetime and steady-state spectrometer. The absolute quantum yield of EuIII complex 3 was obtained by following the standard procedures using the same Edinburgh FLS920 instrument, which is equipped with an integrating sphere.

2.4. Computational details

Geometries were optimized at the B3LYP level of theory [40]. The geometries obtained from X-ray diffraction data were used as input structures for the geometry optimizations, which were performed under the freezing of no-hydrogen atoms. The SDD basis set with large-core quasi-relativistic Pseudopotential ECP52MWB was used for Eu atoms, ECP53MWB for Gd atoms, and ECP54MWB for Tb atoms [4143], whereas the 6–31G(d) basis set was employed for all other atoms [44, 45]. The spin state of all the optimized neutral structures was considered as a singlet. Frequencies were analytically computed at the same level of theory to obtain the gas phase energies, and to confirm whether the structures are minima (no imaginary frequency). All of the calculations were performed using the Gaussian 16 program package, and the structures and molecule orbitals (MOs) were generated by GaussView 6.0. And the Cartesian coordinates of the ligand tppz and the complexes 35 used in the DFT calculation are provided in Table S8.

3. Results and discussion

3.1. Crystal structures

Complexes 16 are isostructural and crystallize in the monoclinic space group P21/n. Therefore, EuIII complex 3 was selected as a representative of the six compounds to discuss in detail their X-ray structures. The molecular structure of 3 is shown in Fig. 1 (a). Each ten-coordinated EuIII center is surrounded by three nitrogen atoms located on one side of the bis-tridentate ligand tppz (N2 comes from the pyrazine, whereas N1 and N3 originate from the pyridine), six oxygen atoms (O1 and O3, O4 and O6, O7 and O9) of three bidentate nitrates, and one oxygen atom (O10) from the coordinated water. The analysis of the coordination geometries around the LnIII center using SHAPE 2.1 reveals that the LnN3O7 lanthanide cores adopt a Sphenocorona J87 geometry (Fig. 1b) [4648]. The different figures of merits calculated for 16 are listed in Table S3.

Fig. 1.

Fig. 1.

(a) Molecular structure of [Eu(NO3)3(H2O)]2(μ-tppz) (3) showing 50% thermal ellipsoids (hydrogen atoms are not shown for clarity, symmertry code a: 1-x, 1-y, 2-z); (b) Sphenocorona J87 geometry of EuN3O7 in 3; (c) and (d) Perspective view of the molecular structure of 3 showing the pyrazine and pyridyl rings numbering, symmertry code a: 1-x,1-y, 2-z.

In order to describe the structural characteristics of the coordinated tppz ligand, the label of the pyrazine and pyridyl rings was done in accordance with the reported notation found in the literature [19]. The pyrazine ring is labeled as E and the four pyridyl rings are labeled as A, B, C and D corresponding to the rings on the 2, 3, 5, and 6 positions of the pyrazine, respectively (Fig. 1c and d). There are fourteen conformations of tppz, which can be divided into five families based on the relative positions of the four pyridyl nitrogen atoms. According to the torsion angles (N2C7C8N1 = 33.577°, N2C5C6N3 = −31.577°, N2aC7aC8aN1a = −33.577°, and N2aC5aC6aN3a = 31.577°), the conformation of tppz in 16 could be designated as 5XXXX. The number 5 indicates that the nitrogen atoms in the A and D rings are “up” and the nitrogen atoms in the B and C rings are “down”. The letter “X” following the number 5 indicates an “exo” conformation, where the pyridyl nitrogen is pointed towards the pyrazine nitrogens (|Npz–C–C–Npy| < 90°) [19]. The dihedral angles between the ring planes are listed in Table 1. The dihedral angles between the A-C rings and B-D rings are 0.311° and 0.689°, respectively. This suggests that the two rings are roughly coplanar. On the other hand, the dihedral angles between the A-B rings and A-D rings are 49.104° and 49.304°, which are consistent with the reported value of 50° [19]. Finally, it must be mentioned that the pyridyl rings A and D twist from coplanarity with the pyrazine E by 32.734° and 36.908°, respectively.

Table 1.

Dihedral angles of the pyridyl rings (A-D) with respect to the pyrazine ring (E) in 3.

Rings Angle (°) Rings Angle (°)
A-E 32.734 A-B 49.104
A-D 49.304 D-E 36.908
A-C 0.311 B-D 0.689

The Eu–N and Eu–O bond distances in the 2.583–2.609 Å and 2.391–2.604 Å ranges, are comparable to those reported for lanthanide complexes (2.556–2.610 and 2.309–2.620 Å), respectively [4951]. The average Ln-O and Ln-N bond lengths (Table S2 of ESI) and the distances between the LnIII⋯LnIII centers (Table 2) decrease gradually from 1 to 6, which are consistent with the effect of the lanthanide contraction [52, 53]. The cross-bridge LnIII⋯LnIII distances (7.694–7.871 Å) are within those reported for the [Cp*2Yb]2(tppz) (7.57 Å) [33] and [(Cp*2Dy)2(μ-tppz)](BPh4) (7.71 Å) complexes [34], but longer than the ones observed for the transition metal tppz complexes (6.497–7.523 Å) [27,30,5459].

Table 2.

LnIII⋯LnIII distances in complexes 1–6.

LnIII⋯LnIII Distances (Å) LnIII⋯LnIII Distances (Å)
NdIII⋯dIII 7.871 GdIII⋯GdIII 7.759
SmIII⋯SmIII 7.815 TbIII⋯TbIII 7.718
EuIII⋯EuIII 7.753 DyIII⋯DyIII 7.694

As shown in Fig. 2, the dinuclear lanthanide units show a parallel arrangement to each other through layered π-π type interactions originated from the two coordinated pyridyl rings (perpendicular distance of 3.469 Å, centroid-to-centroid distance of 4.303 Å, and offset angle of 60.53°). Furthermore, they are linked together by the hydrogen bonds O10–H10A⋯O3, O10–H10A⋯O2, and O10–H10B⋯O2, which involve the coordinated water molecule (O10) and two coordinated nitrate oxygen atoms (O2 and O3) and, thus, giving rise to a two-dimensional frame network (Table S4 of ESI).

Fig. 2.

Fig. 2.

(a) 2D structure of 3 interconnected by hydrogen bonds and π-π stacking interactions; (b) A view of the hydrogen bonds and π-π stacking interactions (yellow dotted lines) linking the double LnIII units.

3.2. IR spectroscopy

The IR spectral data of the tppz ligand (Fig. S1) and complexes 16 (Fig. S2) are listed in Table S5. The broad absorption bands around 3400 cm−1 shown in the spectra are assigned to ʋ(O–H), indicating the presence of water molecules. The aromatic C–H stretching vibrations at 3093 – 3097 cm−1 observed in the complexes shift towards higher wave numbers with respect to the free ligand (3049 cm−1). After coordination to the lanthanide centers the ʋ(C=C) and ʋ(C=N) stretching vibrations of the free ligand at 1585 and 1564 cm−1 are blue-shifted to 1634 and 1598 cm−1 for 16, respectively [27,54]. In addition, the ʋ(C–C) and ʋ (C–N) of the free ligand at 1391, 1125, and 1035 cm−1, and the aromatic C–H deformation vibrations of adjacent hydrogen from the benzene ring at around 806, 784 and 753 cm−1 are somewhat shifted upon coordination to the lanthanide ions. These observations are consistent with those reported in the literature and suggesting the coordination of the tppz ligand to the lanthanide ions [31,57]. The vibration of nitrate νas (NO3) appears at 1503–1506 cm−1 (ν4) and 1281–1291 cm−1 (ν1). The differences of 215–222 cm−1 between ν4 and ν1 suggest a bidentate chelating mode of the nitrato ligand [60]. All of the spectroscopic observations have been confirmed by the X-ray structures of 16.

3.3. Thermogravimetric analysis and PXRD

The thermal properties from crystalline samples of 16 were investigated. They all show a similar thermal decomposition behavior. The TGA/DTG curves of 16 all show a two-step weight loss pattern, as seen in Fig. S3 and from the data summarized in Table S6. Thus, the NdIII complex 1 was selected as a representative of the six compounds to discuss in detail their thermogravimetric analyses. The first decomposition step, which is attributed to the loss of the coordinated water molecule, occurs in the temperature range of 187–256 °C with a weight loss of 3.37 % (calcd. 3.32 %). The second step from 322 to 616 °C is assigned to the removal of the tppz ligand and nitrates with a weight loss of 69.66 % (calcd. 70.18 %). As a result, the final residue is attributed to Ln2O3, as suggested by the weight loss calculations.The powder XRD patterns of 16 were measured at room temperature. It must be noted that the simulated X-ray powder curves obtained from the single-crystal diffraction data and the experimental curves from bulk samples were also calculated and compared (Fig. S4). The simulated intensity and diffraction angles of the main diffraction peaks of 16 are in agreement with the experimental ones, which confirmed the phase purity of the complexes [52,61].

3.4. Hirshfeld surface analysis

To get a deeper insight into the intermolecular interactions within the crystal structures, CrystalExplorer was used to analyze the crystal structures of 16 [62,63]. The 3D Hirshfeld surface and the associated two-dimensional (2D) fingerprint plots of 3 (as an example since the six compounds showed similar patterns) are plotted in Fig. 3. The Hirshfeld surface is mapped with dnorm (the normalized contact distance), which is normalized from de (the nearest external distance), di (the nearest internal distance), and the van der Waals (vdW) radii of the two atoms to the surface. The fingerprint plots are derived from the 3D Hirshfeld surface and provide a visual summary of the frequency of each combination of de and di across the surface of a molecule [39,52]. Table S7 summarizes the relative contributions of various intermolecular interactions to the Hirshfeld surfaces in the 16 crystal structures.

Fig. 3.

Fig. 3.

(a) Hirshfeld surface mapped with dnorm, (b) shape index, (c) curvedness, and (d–f) fingerprint plots for EuIII complex 3.

In the dnorm Hirshfeld surface, the regions of shortest intermolecular contacts are visualized by red circle areas [64]. The main partial contributions of the different intermolecular contacts are H⋯O and O⋯H contacts in 3 with a relative contribution of 62.3% (Fig. 3a). In fact, the H⋯O/O⋯H contacts vary from 57.5% in 1–62.4% in 5 (Table S7). Thus, the Hirshfeld surface analysis for the X-ray structure of the six complexes reveals that the crystal packing is primarily determined by H⋯O and O⋯H intermolecular contacts.62 The red triangles in the shape index plot represent concave regions, which result from the carbon atoms of the π-stacked molecule above it. On the other hand, the blue triangles in the curvedness plot refer to convex regions because of the ring carbon atoms of the molecule inside the surface. The outline of alternating red and blue triangles is an indicator of π⋯π stacking interactions [65]. The π⋯π stacking information conveyed by the shape index (Fig. 3b) and the curvedness plot (Fig. 3c) are consistent with the crystal structure analyses.

3.5. UV–Vis absorption and photoluminescence

Ambient-temperature UV–Vis absorption spectra (Fig. S5) of tppz ligand and complexes 16 were recorded in DMF solution (1 × 10−6 mol/L). The spectrum of tppz displays two absorption bands positioned at around 278 and 310 nm, which can be ascribed to the n→π* or π→π* transitions centered on the π-conjugated unit, respectively. Upon complexation to the LnIII center, the bands observed in the UV–Vis absorption spectrum of tppz are slightly blue-shifted to about 276 and 308 nm. Similar results were obtained for all complexes.

The solid-state luminescent properties of tppz and 16 were investigated at 77 and 298 K. Under excitation at 342 nm, the emission spectrum of tppz exhibits one broad maximum emission centered at 385 nm at room temperature, which originates from the ligand-centered (LC) singlet excited state energy level 1ππ* (Fig. S6). Knowing that tppz is known to form two polymorphs (monoclinic P21/n and tetragonal I41/a), the observed solid-state luminescent features are consistent with those reported for the monoclinic polymorph. At 77 K, the emission band maxima at about 392 and 541 nm are assigned to the singlet and triplet excited state energy levels 1ππ* and 3ππ*, respectively (Fig. S7). On the other hand, the time-resolved emission spectrum recorded with a time delay of 1.0 ms only shows the triplet excited state energy level 3ππ* with a maximum around 542 nm.

Upon complexation to the nonluminescent GdIII center, the maximum of the emission band of the 1ππ* is red-shifted to about 476 nm with a much weaker and broad emission at 627 nm tentatively ascribed to the ligand-to-ligand charge transfer transitions (LLCT) [54, 55,66] (Fig. S8, room temperature). At 77 K, the steady-state emission spectrum (Fig. S9, left) showed that the LLCT band at around 633 nm becomes stronger than the 1ππ* band at 474 nm.52, 64 Since the nonluminescent GdIII ion is commonly used to determine the position of the excited triplet state energy level of coordinated ligands, the time-resolved luminescence spectrum of 4, measured at 77 K with a time delay of 1 ms (Fig. S9, right), showed a more structure emission band with two maxima at about 538 and 571 nm, and thus corresponding to a triplet energy level located at about 538 nm (or 18,587 cm−1). It is worth noting that this value is very close to the one observed for the triplet energy level of 4′-phenyl-2,2′:6′,2″-terpyridine (18,500 cm−1) [52]. Knowing that the corresponding singlet energy level is located at about 474 nm (21,097 cm−1), the energy difference ΔE(1ππ*−3ππ*) is determined to be 2510 cm−1, suggesting the presence of an inefficient intersystem crossing (ISC) process. This is in agreement with the fact that one expects that an efficient ISC process takes place when ΔE (1ππ*−3ππ*) is greater than 5000 cm−1 [2,67].

Upon excitation of the ligand-centered absorption band, the emission spectra of 1, 2, 3, and 5 are dominated by narrow bands corresponding to the typical NdIII, SmIII, EuIII, and TbIII lanthanide-centered transitions, respectively. The NdIII complex 1 shows typical NIR emissions upon excitation at 449 nm. As shown in Fig. 4a, 1 displays the strong characteristic emission band of the NdIII ion centered at 1054 nm and assigned to the 4F3/24I11/2 transition. It can also be seen the two splitting peaks of the 4F3/24I9/2 (881 and 901 nm) and 4F3/24I13/2 (1321 and 1333 nm) transitions. The energy of the ligand triplet state (18,587 cm−1), which is much higher than the corresponding emitting levels of NdIII (11,460 cm−1), thus facilitating their population through ligand-to-metal energy transfer [68]. Indeed, the observation of the characteristic emission bands of the NdIII confirms that the tppz moieties can act as effective sensors for the luminescence of NdIII ions [5,67].

Fig. 4.

Fig. 4.

(a–d) Solid-state excitation (λem = 1056, 494, 616 and 541 nm) and emission (λex = 449, 375, 439 and 337 nm) and emission spectra of 1, 2, 3 and 5 at room temperature; (e) CIE chromaticity coordinates.

As shown in Fig. 4b, SmIII complex 2 exhibits weak characteristic emission bands of the SmIII ion upon excitation at 375 nm at room temperature. The bands observed at 561, 599, 642, and 705 nm correspond to 4G5/26H5/2 (forbidden), 4G5/26H7/2 (magnetic dipole), 4G5/26H9/2 (electric dipole), and 4G5/26H11/2 transitions, respectively [69]. The observation of the ligand-centered emission around 450–470 nm in the emission spectrum of the SmIII complex at 77 K suggests an inefficient ISC, thus leading to unfavorable conditions to observe significantly the characteristic SmIII-centered emission (Fig. S10) [70].

The luminescent bands of EuIII in complex 3 (Fig. 4c) appear at 579, 592, 616, 648 and 689 nm, and are assigned to the 5D07FJ (J = 0, 1, 2, 3, and 4) transitions of the EuIII center with the CIE value of (0.6654, 0.3344) (Fig. 4e), respectively. The hypersensitive electric dipole transition 5D07F2 at 616 nm is the strongest emission band in the luminescence spectrum of 3 measured at 298 K and upon excitation of 449 nm. These observations suggest that the EuIII ion occupied a site of low symmetry without an inversion center. Furthermore, the appearance of the symmetry-forbidden emission 5D07F0 at 579 nm indicates that the EuIII cation possesses a non-centrosymmetric environment [52]. The absolute quantum yield of 3 is 1.31%, which is much less than the reported value (16.8%) of the terpyridine analogue complex [Eu (NO3)3(ptpy)(H2O)] (where ptpy is 4′-phenyl-2,2′:6′,2′′-terpyridine) [52]. In addition, the 5D07F2 transition was used to measure the luminescence lifetime. The decay curve of 3 was fitted using a mono-exponential function, which gave a lifetime of 0.035 ms that is much shorter than the one observed for [Eu(NO3)3(ptpy)(H2O)] (0.545 ms) [52]. It can be concluded from the low absolute quantum yield and much shorter luminescence lifetime observed for complex 3 that the tppz ligand does not have efficient sensitization abilities towards the EuIII cation unlike its analog ptpy ligand. However, it is noteworthy that the ligand-centered emissions are not observed in the steady-state and time-resolved emission spectra of 3 (Fig. S11) at 77 K, indicating efficient energy transfer processes taking place [71]. Additionally, the efficiency of the luminescence sensitization by the ligand, Φsen, amounted to 85.6% (see the Supporting Information for a detailed determination of this value).

Upon excitation of the ligand-centered absorption band at room temperature, the TbIII complex 5 displays the emission bands of 5D47F6 (488 nm), 5D47F5 (541 nm), 5D4 7F4 (581 nm), 5D47F3 (622 nm), 5D47F2 (647 nm), 5D47F1 (668 nm), and 5D47F0 (680 nm) of the TbIII cation (Fig. 4d), which are responsible for the characteristic green color of TbIII with the CIE value of (0.2985,0.5620) (Fig. 4e) [72]. The observation of the emission of the excited singlet state energy level 1ππ* in the luminescence spectra of 5 at room temperature and 77 K confirms that there is an inefficient ISC process taking place (Fig. S11).

At room temperature, the DyIII complex 6 only exhibits one broad emission centered at 483 nm (Fig. S12), which is assigned to the ligand-centered (LC) 1ππ* transition. At 77 K, this emission band appears more structured with a maximum around 463 nm (Fig. S12). Although the emission spectrum does not show the characteristic emission bands of the DyIII ion, the red-shift of the emission band of the 1ππ* state in the spectrum of 6 compared to its position for the free ligand confirms the coordination of the tppz ligand to the DyIII cation (a difference of about 94 nm). Like for the SmIII ion, it must be concluded that the tppz ligand is not a good sensor for the luminescence of the DyIII cation.

Normally, it necessitates the energy of the lowest excited triplet state level of the sensitizer to be nearly equal or above the resonance level. The energy-level match between the triplet state of the antenna (aka coordinated ligand) and the resonance level of LnIII ions is a vital factor that defines the luminescence efficiency of LnIII complexes. Evidently, the effectiveness of the energy transfer from the antenna moiety to the LnIII cation is somewhat proportional to the extent of overlap between the phosphorescence spectrum of the ligand and the absorption spectrum of the LnIII ion. An effective ligand to LnIII ion energy transfer is apparent from the quenching of the emission of the coordinated ligands [2]. The determined triplet energy level of the ligand (18,587 cm−1) is 1316 and 762 cm−1 higher than the excited state of EuIII (17,271 cm−1) and SmIII (17825 cm−1), but less than the excited states of TbIII (20,400 cm−1), and DyIII (21,000 cm−1). As a result, one expects that the energy transfer from tppz to the SmIII, TbIII, and DyIII centers is not favorable. It is worth noting that the 0-phonon of the 3ππ* state of the complexed ligand in the GdIII-containing compound 4 at 18,587 cm−1 is better suited to sensitize the EuIII (17,271 cm−1) ions than the SmIII (17,825 cm−1), TbIII (20,400 cm−1) and DyIII (21,000 cm−1) cations. It must be pointed out that it was not possible to obtain a measurable lifetime and quantum yield for the TbIII complex 5. This is also in line with the fact that the EuIII complex 3 has a low quantum yield (1.31%) and short lifetime (0.035 ms). These poor luminescence properties may be explained by the presence of O⋯H/H⋯O intermolecular interactions [73], as evidenced by the 3-D Hirshfeld surface and 2-D fingerprint plots. The O⋯H/H⋯O intermolecular interactions are the main’s interactions in 16 with a contribution of 62% as determined for 3. As a result, the presence of O–H oscillators from water molecules in the coordination sphere of the 16 compounds effectively quench the characteristic luminescence of the lanthanide(III) ions.50, 66, 68

3.6. DFT calculations

Density functional theory (DFT) calculations were conducted to unravel the HOMO–LUMO energy gaps of the structures of tppz and complexes 3, 4 and 5. The HOMO–LUMO energy gaps are listed in Table 3. As shown in Fig. 5, the energy gaps between HOMO and LUMO for tppz, 3, 4 and 5 are 4.80, 3.38, 3.30, and 3.37 eV, respectively. It is worth mentioning that the HOMO and LUMO profiles are all located on the pyrazine and pyridyl groups for the tppz ligand. After coordination with the lanthanide(III) center in 3, 4 and 5, the HOMO profiles are more located on one of the nitrates, whereas the LUMOs are located on the tppz ligand. As a result, the energy gaps between the HOMOs and LUMOs decrease by 1.42–1.50 eV (11,453–12,098 cm−1), which corroborate the formation of the complexes 3, 4 and 5. The structure calculated for the complexes 3, 4 and 5 presents average Eu–N, Gd–N and Tb–N distances of 2.601, 2.596 and 2.580 Å, respectively, which compare well with the average experimental values obtained from X-ray diffraction measurements (Eu–N: 2.601, Gd–N: 2.597 and Tb–N: 2.581 Å).

Table 3.

Theoretically energy values for tppz and complexes 3–5.

Compd. Energy
LUMO (eV) HOMO (eV) ΔE (eV) ΔE (cm−1)
Tppz −1.77 −6.57 4.80 38,714
3-Eu −3.96 −7.34 3.38 27,261
4-Gd −3.98 −7.28 3.30 26,616
5-Tb −3.95 −7.32 3.37 27,181

Fig. 5.

Fig. 5.

HOMOs, LUMOs and HOMO-LUMO gaps of tppz, complexes 3, 4 and 5.

3.7. Magnetic properties

The temperature dependence of the magnetic susceptibilities of the DyIII comple×6 was recorded from 2 to 300 K at 1000 Oe, as depicted in Fig. 6. It shows that the χm value decreases with T in a monotonous form (going 0.09 and 9.46 cm3 mol−1 from 300 to 2 K). The χmT values (where χm being the molar magnetic susceptibility) of 6 at 300 K is 28.01 cm3 K mol−1, which is slightly lower than the theoretical value of 28.34 cm3 K mol−1 for two magnetic uncoupled DyIII ions (6H15/2, g = 4/3 and χmT = 14.17 cm3 K mol−1). The χmT value decreases gradually and reaches a minimum of 18.95 cm3 K mol−1 at 2 K. The decrease of χmT profiles may arise from the progressive excited depopulation of stark sublevel and/or non-negligible intra- or intermolecular antiferromagnetic interactions between the DyIII ions [74,75]. From Fig. 6 inset, it can be seen that the temperature dependence of the reciprocal susceptibility (χm1) above 50 K follows the Curie–Weiss law [χm = C/(Tθ)] with a Weiss constant of θ = − 9.87 K and Curie constant of C = 28.99 mol cm−3. Both the overall profile of χmT-T curve and the negative θ value might suggest antiferromagnetic interactions between DyIII ions in the system [76].

Fig. 6.

Fig. 6.

Plots of χmT-T, χm-T and Curie-Weiss fitting for the χm1T data of 6.

4. Conclusions

The characterizations, Hirshfeld surface analyses and solid-state luminescence properties of six dinuclear bis-tridentate N-heterocyclic ligand bridging lanthanide complexes [Ln(NO3)3(H2O)]2(μ-tppz) (Ln (III) = Nd (1), Sm (2), Eu (3), Gd (4), Tb (5), and Dy (6)) were reported. Complexes 1 and 3 exhibit characteristic emissions of the central NdIII and EuIII ions under the UV excitation, which suggest that the tppz ligand could be a good organic chromophore to absorb energy and transfer it to these lanthanide(III) ions. Unlike for the NdIII and EuIII ions, tppz is not a good sensor for the luminescence of the TbIII, SmIII, and DyIII cations (the emission spectra of 2, 5 and 6 mainly displayed the ligand-centered (LC) 1ππ* emissions). Finally, the DFT calculations were conducted to unravel the HOMO–LUMO energy gaps of the structures of tppz, 3, 4, and 5. After coordination of tppz to the lanthanide(III) center, it showed that the HOMO–LUMO energy gaps decreased to 1.42–1.50 eV (11,453–12,098 cm−1). This observed decrease was a confirmation of the formation of the complexes studied. Additionally, the magnetic data of 6 confirmed the antiferromagnetic coupling phenomenon between the DyIII ions. It is worth noting that the ligand tppz is better suited to sensitize EuIII and NdIII, than the other lanthanide(III) ions investigated (SmIII, TbIII, and DyIII). Additionally, the overall poor luminescence properties of all lanthanide(III)-containing compounds are most likely the consequence of the presence of nonradiative quenching processes such as the presence of water molecules in the coordination sphere of the lanthanide(III) ions (important contribution of O⋯H/H⋯O intermolecular interactions observed for the EuIII complex 3). To remedy that, ongoing work involves the formation of related lanthanide(III) complexes by using co-ligands to act as additional sensitizers to replace the OH quenchers in order to improve the luminescent properties of the tppz-bridging systems. Thus, the study of such hetero-binuclear complexes with this bis-tridentate ligand is currently under investigation.

Supplementary Material

Table S1-S8-FS1-FS16-

Acknowledgements

We gratefully thank the financial support by the National Natural Science Foundation of China (grant no. 61765005), the Guangxi Key Laboratory of Electrochemical and Magneto-chemical Functional Materials (grant no. EMFM20161106), and the Doctoral Scientific Research Foundation of Guilin University of Technology (grant no. GLUTQD2011039). The financial support provided by the China Scholarship Council (CSC) during a visit of Dr. Yan Li to San Jośe State University is also greatly acknowledged. G.M. acknowledges the National Institute of Health, Minority Biomedical Research Support (grant 1 SC3 GM089589-08) and the Henry Dreyfus Teacher-Scholar Award.

Footnotes

Electronic Supplementary Information (ESI) available: refinement details and crystallographic data for the X-ray diffraction studies. CCDC 1813714 to 1813719. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/x0xx00000x Accepted January 00, 2020.

Declaration of competing interest

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.

Appendix A. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.jlumin.2020.117835.

References

  • [1].Bünzli J-CG, On the design of highly luminescent lanthanide complexes, Coord. Chem. Rev 293–294 (2015) 19–47. [Google Scholar]
  • [2].SeethaLekshmi S, Ramya AR, Reddy MLP, Varughese S, Lanthanide complex-derived white-light emitting solids: a survey on design strategies, J. Photoch. Photobio. C 33 (2017) 109–131. [Google Scholar]
  • [3].Bünzli J-CG, Lanthanide light for biology and medical diagnosis, J. Lumin 170 (2016) 866–878. [Google Scholar]
  • [4].Utochnikova VV, The use of luminescent spectroscopy to obtain information about the composition and the structure of lanthanide coordination compounds, Coord. Chem. Rev 398 (2019) 113006. [Google Scholar]
  • [5].Ning YY, Zhu ML, Zhang JL, Near-infrared (NIR) lanthanide molecular probes for bioimaging and biosensing, Coord. Chem. Rev 399 (2019) 19. [Google Scholar]
  • [6].Teo RD, Termini J, Gray HB, Lanthanides: applications in cancer diagnosis and therapy, J. Med. Chem 59 (2016) 6012–6024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Long J, Guari Y, Ferreira RAS, Carlos LD, Larionova J, Recent advances in luminescent lanthanide based Single-Molecule Magnets, Coord. Chem. Rev 363 (2018) 57–70. [Google Scholar]
  • [8].Armelao L, Quici S, Barigelletti F, Accorsi G, Bottaro G, Cavazzini M, Tondello E, Design of luminescent lanthanide complexes: from molecules to highly efficient photo-emitting materials, Coord. Chem. Rev 254 (2010) 487–505. [Google Scholar]
  • [9].Bünzli J-CG, Eliseeva SV, Lanthanide NIR luminescence for telecommunications, bioanalyses and solar energy conversion, J. Rare Earths 28 (2010) 824–842. [Google Scholar]
  • [10].Santos HP, Gomes ES, dos Santos MV, D’Oliveira KA, Cuin A, Martins JS, Quirino WG, Marques LF, Synthesis, structures and spectroscopy of three new lanthanide β-diketonate complexes with 4,4′-dimethyl-2,2′-bipyridine. Near-infrared electroluminescence of ytterbium(III) complex in OLED, Inorg. Chim. Acta 484 (2019) 60–68. [Google Scholar]
  • [11].Petrosyants SP, Ilyukhin AB, Gavrikov AV, Mikhlina YA, Puntus LN, Varaksina EA, Efimov NN, Novotortsev VM, Luminescent and magnetic properties of mononuclear lanthanide thiocyanates with terpyridine as auxiliary ligand, Inorg. Chim. Acta 486 (2019) 499–505. [Google Scholar]
  • [12].Bai C, Fu X-Y, Hu H-M, He S, Wang X, Xue G-L, Construction of visible luminescent lanthanide coordination compounds with different stacking modes based on a carboxylate substituted terpyridyl derivative ligand, Inorg. Chim. Acta 506 (2020) 119550. [Google Scholar]
  • [13].Batrice RJ, Ridenour JA, Ayscue III RL, Bertke JA, Knope KE, Synthesis, structure, and photoluminescent behaviour of molecular lanthanide-2-thiophenecarboxylate-2,2’:6’,2”-terpyridine materials, CrystEngComm 19 (2017) 5300–5312. [Google Scholar]
  • [14].Carter KP, Pope SJA, Cahill CL, A series of Ln-p-chlorobenzoic acid-terpyridine complexes: lanthanide contraction effects, supramolecular interactions and luminescent behavior, CrystEngComm 16 (2014) 1873–1884. [Google Scholar]
  • [15].Zheng K, Ding L-W, Zeng C-H, Highly luminescent lanthanide complexes constructed by Bis-tridentate ligand and as sensor for Et2O, Inorg. Chem. Commun 95 (2018) 95–99. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Zeng C-H, Wu H, Luo Z, Yao J, Weak interactions cause selective cocrystal formation of lanthanide nitrates and tetra-2-pyridinylpyrazine, CrystEngComm 20 (2018) 1123–1129. [Google Scholar]
  • [17].Fugisawa FP, Ramos AP, de Sousa PC, Serra OA, Zaniquelli MED, Formation of thin luminescent Eu3+-LB films by in situ coordination with 2,3,5,6-tetra(2’-pyridyl)pyrazine and 1-octadecanol in pure and mixed Langmuir monolayers, J. Lumin 132 (2012) 1116–1121. [Google Scholar]
  • [18].Padgett CW, Walsh RD, Drake GW, Hanks TW, Pennington WT, New conformations and binding modes in halogen-bonded and ionic complexes of 2,3,5,6-tetra(2’-pyridyl)yoyrazine, Cryst. Growth Des 5 (2005) 745–753. [Google Scholar]
  • [19].Padgett CW, Pennington WT, Hanks TW, Conformations and binding modes of 2,3,5,6-tetra(2’-pyridyl)pyrazine, Cryst. Growth Des 5 (2005) 737–744. [Google Scholar]
  • [20].Campos-Fernandez CS, Smucker BW, Clerac R, Dunbar KR, Reactivity studies of 2,3,5,6-tetra(2-pyridyl) pyrazine (tppz) with first-row transition metal ions, Isr. J. Chem 41 (2001) 207–218. [Google Scholar]
  • [21].Bock H, Jaculi D, Electronentransfer und ionenpaar-bildung, 20[1,2] zweistufen-reduktion von tetra(2’-pyridyl)pyrazin unter aprotischen bedingungen sowie in Loesungen mit Li+-und Na+-leitsalzkationen, Z. Naturforsch. B Chem. Sci 46 (1991) 1091–1097. [Google Scholar]
  • [22].Goodwin HA, Lions F, Tridentate chelate compounds. II, J. Am. Chem. Soc 81 (1959) 6415–6422. [Google Scholar]
  • [23].Behnamfar MT, Hadadzadeh H, Simpson J, Darabi F, Shahpiri A, Khayamian T, Ebrahimi M, Rudbari HA, Salimi M, Experimental and molecular modeling studies of the interaction of the polypyridyl Fe(II) and Fe(III) complexes with DNA and BSA, Spectrochim. Acta A 134 (2015) 502–516. [DOI] [PubMed] [Google Scholar]
  • [24].Arana CR, Abruna HD, Monomeric and oligomeric complexes of ruthenium and osmium with tetra-2-pyridyl-1,4-pyrazine (TPPZ), Inorg. Chem 32 (1993) 194–203. [Google Scholar]
  • [25].Vogler LM, Scott B, Brewer KJ, Investigation of the phptochemical electrochemical and spectroelectrochemical properties of an iridium(III)/Ruthenium(II) mixed-metal complex bridged by 2,3,5,6-Tetrakis(2-pyridy1) pyrazine, Inorg. Chem 32 (1993) 898–903. [Google Scholar]
  • [26].Graf M, Greaves B, Stoecklievans H, Crystal and molecular structures of Cu(II) and Zn(II) complexes of 2,3,5,6-tetra(2-pyridyl)pyrazine (TPPZ): an old ligand revisited, Inorg. Chim. Acta 204 (1993) 239–246. [Google Scholar]
  • [27].Wan J, Cai S-L, Zhang K, Li C-J, Feng Y, Fan J, Zheng S-R, Zhang W-G, Anion- and temperature-dependent assembly, crystal structures and luminescence properties of six new Cd(II) coordination polymers based on 2,3,5,6-tetrakis(2-pyridyl)pyrazine, CrystEngComm 18 (2016) 5164–5176. [Google Scholar]
  • [28].Wen H-M, Wang J-Y, Zhang L-Y, Shi L-X, Chen Z-N, Syntheses, characterization and electrochemical and spectroscopic properties of ruthenium–iron complexes of 2,3,5,6-tetrakis(2-pyridyl)pyrazine and ferroceneacetylide ligands, Dalton Trans 45 (2016) 10620–10629. [DOI] [PubMed] [Google Scholar]
  • [29].Palion-Gazda J, Switlicka-Olszewska A, Machura B, Grancha T, Pardo E, Lloret F, Julve M, High-temperature spin crossover in a mononuclear six-coordinate cobalt(II) complex, Inorg. Chem 53 (2014) 10009–10011. [DOI] [PubMed] [Google Scholar]
  • [30].Nagashima T, Nakabayashi T, Suzuki T, Kanaizuka K, Ozawa H, Zhong Y-W, Masaoka S, Sakai K, Haga M.-a., Tuning of metal–metal interactions in mixed-valence states of cyclometalated dinuclear ruthenium and osmium complexes bearing tetrapyridylpyrazine or -benzene, Organometallics 33 (2014) 4893–4904. [Google Scholar]
  • [31].Sadati SM, Alizadeh R, Amani V, New binuclear Hg(II)-2,3,5,6-tetrakis-(2-pyridyl)pyrazine complexes: synthesis, characterization, luminescence study and crystal structure determination, J. Mol. Struct 1202 (2020) 127324–127330. [Google Scholar]
  • [32].Kuehl CJ, Da Re RE, Scott BL, Morris DE, John KD, Toward new paradigms in mixed-valency: ytterbocene-terpyridine charge-transfer complexes, Chem. Commun (2003) 2336–2337. [DOI] [PubMed] [Google Scholar]
  • [33].Carlson CN, Kuehl CJ, Da Re RE, Veauthier JM, Schelter EJ, Milligan AE, Scott BL, Bauer ED, Thompson JD, Morris DE, John KD, Ytterbocene charge-transfer molecular wire complexes, J. Am. Chem. Soc 128 (2006) 7230–7241. [DOI] [PubMed] [Google Scholar]
  • [34].Demir S, Nippe M, Gonzalez MI, Long JR, Exchange coupling and magnetic blocking in dilanthanide complexes bridged by the multi-electron redox-active ligand 2,3,5,6-tetra(2-pyridyl)pyrazine, Chem. Sci 5 (2014) 4701–4711. [Google Scholar]
  • [35].Yuasa J, Fukuzumi S, OFF-OFF-ON switching of fluorescence and electron transfer depending on stepwise complex formation of a host ligand with guest metal ions, J. Am. Chem. Soc 130 (2008) 566–575. [DOI] [PubMed] [Google Scholar]
  • [36].Sheldrick GM, A short history of SHELX, Acta Crystallogr. A 64 (2008) 112–122. [DOI] [PubMed] [Google Scholar]
  • [37].Sheldrick GM, SHELXT–Integrated space-group and crystal-structure determination, Acta Crystallogr. A 71 (2015) 3–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [38].Turner MJ, McKinnon JJ, Wolff SK, Grimwood DJ, Spackman PR, Jayatilaka D, Spackman MA, CrystalExplorer17, University of Western Australia, 2017. CrystalExplorer 17 (2017. [Google Scholar]
  • [39].Spackman MA, Jayatilaka D, Hirshfeld surface analysis, CrystEngComm 11 (2009) 19–32. [Google Scholar]
  • [40].Stephens PJ, Devlin FJ, Chabalowski CF, Frisch MJ, Ab initio calculation of vibrational absorption and circular dichroism spectra using density functional force fields, J. Phys. Chem 98 (1994) 11623–11627. [Google Scholar]
  • [41].Andrae D, Häußermann U, Dolg M, Stoll H, Preuß H, Energy-adjusted ab initio pseudopotentials for the second and third row transition elements, Theor. Chim. Acta 77 (1990) 123–141. [Google Scholar]
  • [42].Dolg M, Stoll H, Preuss H, A combination of quasirelativistic pseudopotential and ligand field calculations for lanthanoid compounds, Theor. Chim. Acta 85 (1993) 441–450. [Google Scholar]
  • [43].Dolg M, Stoll H, Savin A, Preuss H, Energy-adjusted pseudopotentials for the rare earth elements, Theor. Chim. Acta 75 (1989) 173–194. [Google Scholar]
  • [44].Hariharan PC, Pople JA, The influence of polarization functions on molecular orbital hydrogenation energies, Theor. Chim. Acta 28 (1973) 213–222. [Google Scholar]
  • [45].Gans PJ, Application of the Fisher-sykes attrition relation to the diamond lattice, J. Chem. Phys 56 (1972) 2253–2257. [Google Scholar]
  • [46].Alemany P, Casanova D, Alvarez S, Dryzun C, Avnir D, Continuous symmetry measures: a new tool in quantum chemistry, in: Parrill AL, Lipkowitz KB (Eds.), Reviewsin Computational Chemistry, 2017, pp. 289–352. Wiley, New York. [Google Scholar]
  • [47].Alvarez S, Alemany P, Casanova D, Cirera J, Llunell M, Avnir D, Shape maps and polyhedral interconversion paths in transition metal chemistry, Coord. Chem. Rev 249 (2005) 1693–1708. [Google Scholar]
  • [48].Zabrodsky H, Peleg S, Avnir D, Contionuous symmetry measures J. Am. Chem. Soc 114 (1992) 7843–7851. [Google Scholar]
  • [49].Wang Y, Shen P-P, Ren N, Zhang J-J, Geng L-N, Wang S-P, Shi S-K, A series of lanthanide complexes with different N-donor ligands: synthesis, structures, thermal properties and luminescence behaviors, RSC Adv 6 (2016) 70770–70780. [Google Scholar]
  • [50].Gou L, Wu Q-R, Hu H-M, Qin T, Xue G-L, Yang M-L, Tang Z-X, A new family of lanthanide terpyridine nitrate complexes: solvothermal syntheses, crystal structures and luminescent properties of [Ln(pytpy)(NO3)2(μ-OCH3)]2, Inorg. Chim. Acta 361 (2008) 1922–1928. [Google Scholar]
  • [51].Cotton SA, Noy OE, Liesener F, Raithby PR, Unequivocal characterisation of a [Ln(terpy)(NO3)3·(H2O)] complex. The synthesis and structure of [M(terpy) (NO3)3·(H2O) (M = Eu, Tb); a comparison with the structure of [Eu(bipy)2(NO3)3 and with other europium nitrate complexes {terpy = 2,2’:6’,2”-terpyridyl; bipy = 2,2’-bipyridyl}, Inorg. Chim. Acta 344 (2003) 37–42. [Google Scholar]
  • [52].Cai LL, Hu YT, Li Y, Wang K, Zhang XQ, Muller G, Li XM, Wang GX, Solid-state luminescence properties, Hirshfeld surface analysis and DFT calculations of mononuclear lanthanide complexes (Ln = EuIII, GdIII, TbIII, DyIII) containing 4′-phenyl-2,2′:6′,2′′-terpyridine, Inorg. Chim. Acta 489 (2019) 85–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Cotton SA, Raithby PR, Systematics and surprises in lanthanide coordination chemistry, Coord. Chem. Rev 340 (2017) 220–231. [Google Scholar]
  • [54].Palion-Gazda J, Gryca I, Maroń A, Machura B, Kruszynski R, Cyanate cadmium (II) coordination compounds with 2,3,5,6-tetrakis(2-pyridyl)pyrazine-Synthesis, structure and luminescent properties, J. Lumin 192 (2017) 713–719. [Google Scholar]
  • [55].Palion-Gazda J, Gryca I, Maroń A, Machura B, Kruszynski R, Thiocyanate cadmium(II) coordination compounds with 2,3,5,6-tetrakis(2-pyridyl)pyrazine-Synthesis, structure and luminescent properties, Polyhedron 135 (2017) 109–120. [Google Scholar]
  • [56].Wu SH, Burkhardt SE, Zhong YW, Abruna HD, Cyclometalated ruthenium oligomers with 2,3-di(2-pyridyl)-5,6-diphenylpyrazine: a combined experimental, computational, and comparison study with noncyclometalated analogous, Inorg. Chem 51 (2012) 13312–13320. [DOI] [PubMed] [Google Scholar]
  • [57].Yuste C, Canadillas-Delgado L, Ruiz-Perez C, Lloret F, Julve M, Dinuclear, tetranuclear and one-dimensional pyrazine-based copper(II) complexes: preparation, X-ray structure and magnetic properties, Dalton Trans 39 (2010) 167–179. [DOI] [PubMed] [Google Scholar]
  • [58].Kundu T, Sarkar B, Mondal TK, Fiedler J, Mobin SM, Kaim W, Lahiri GK, Carboxylate tolerance of the redox-active platform Ru(mu-tppz)Ru (n), where tppz = 2,3,5,6-Tetrakis(2-pyridyl)pyrazine, in the electron-transfer series (L)CIRu(mutppz)RuCl(L) (n), n = 2+, +, 0, −, 2−, with 2-picolinato, quinaldato, and 8-quinolinecarboxylato ligands (L-), Inorg. Chem 49 (2010) 6565–6574. [DOI] [PubMed] [Google Scholar]
  • [59].Trivedi M, Pandey DS, Rath NP, Binuclear copper and zinc complexes based on polypyridyl ligand 2,3,5,6-tetra(2-pyridyl)pyrazine (tppz): synthesis, spectral and structural characterization, Inorg. Chim. Acta 362 (2009) 284–290. [Google Scholar]
  • [60].Li Y, Liang FP, Jiang CF, Li XL, Chen ZL, 2D network coordination polymers of lanthanide with N-(2-pyridylmethyl)iminodiacetic acid: hydrothermal syntheses, crystal structures and luminescent properties, Inorg. Chim. Acta 361 (2008) 219–225. [Google Scholar]
  • [61].Guo K-K, Su S-D, Shen Y-L, Wang K, Li Y, Zhang X-Q, Li B, Zou H-H, Liang F-P, A Cu-3 cluster-based chain featuring linkages of acylhydrazone N-N single bonds and Clions: synthesis, structure and magnetic properties, J. Cluster Sci 30 (2019) 219–224. [Google Scholar]
  • [62].McKinnon JJ, Spackman MA, Mitchell AS, Novel tools for visualizing and exploring intermolecular interactions in molecular crystals, Acta Crystallogr. B 60 (2004) 627–668. [DOI] [PubMed] [Google Scholar]
  • [63].McKinnon JJ, Mitchell AS, Spackman MA, Hirshfeld surfaces: a new tool for visualising and exploring molecular crystals, Chem. Eur J 4 (1998) 2136–2141. [Google Scholar]
  • [64].Bouchene R, Bouacida S, Syntheses, supramolecular networks and Hirshfeld surface and thermal analyses of two new cadmium chloride coordination polymers with an N,O-chelating ligand, Acta Acta Cryst. C C75 (2019) 1–8. [DOI] [PubMed] [Google Scholar]
  • [65].Kumar M, Kariem M, Sheikh HN, Frontera A, Seth SK, Jassal AK, A series of 3D lanthanide coordination polymers decorated with a rigid 3,5-pyridinedicarboxylic acid linker: syntheses, structural diversity, DFT study, Hirshfeld surface analysis, luminescence and magnetic properties, Dalton Trans. 47 (2018) 12318–12336. [DOI] [PubMed] [Google Scholar]
  • [66].Nawrot I, Machura B, Kruszynski R, Coordination assemblies of Cd-II with 2,2’: 6’,2”-terpyridine (terpy), 2,3,5,6-tetra-(2-pyridyl)pyrazine (tppz) and pseudohalide ions-structural diversification and luminescence properties, CrystEngComm 17 (2015) 830–845. [Google Scholar]
  • [67].Hua KT, Xu JD, Quiroz EE, Lopez S, Ingram AJ, Johnson VA, Tisch AR, de Bettencourt-Dias A, Straus DA, Muller G, Structural and photophysical properties of visible- and near-IR-emitting tris lanthanide(III) complexes formed with the enantiomers of N,N’-Bis(1-phenylethyl)-2,6-pyridinedicarboxamide, Inorg. Chem 51 (2012) 647–660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Hashami Z, Martins AF, Funk AM, Jordan VC, Petoud S, Eliseeva SV, Kovacs Z, Lanthanide DO3A-tropone complexes: efficient dual MR/NIR imaging probes in aqueous medium, Eur. J. Inorg. Chem (2017) 4965–4968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [69].Ahmed Z, Iftikhar K, Red, orange-red and near-infrared light emitting ternary lanthanide tris -diketonate complexes with distorted C-4v geometrical structures, Dalton Trans. 48 (2019) 4973–4986. [DOI] [PubMed] [Google Scholar]
  • [70].Taha ZA, Ajlouni AM, Hijazi AK, Al-Rawashdeh NA, Al-Hassan KA, Al-Haj YA, Ebqa’ai MA, Altalafha AY, Synthesis and luminescent spectroscopy of lanthanide complexes with dimethylpyridine-2,6-dicarboxylate (dmpc), J. Lumin 161 (2015) 229–238. [Google Scholar]
  • [71].Xiao YH, Deng ZP, Zhu ZB, Huo LH, Gao S, Rare earth metal-organic complexes constructed from hydroxyl and carboxyl modified arenesulfonate: syntheses, structure evolutions, and ultraviolet, visible and near-infrared luminescence, Dalton Trans. 46 (2017) 16493–16504. [DOI] [PubMed] [Google Scholar]
  • [72].Andres J, Chauvin A-S, Lanthanides: luminescence applications. Encyclopedia of Inorganic and Bioinorganic Chemistry, John Wiley & Sons, Ltd., 2012. [Google Scholar]
  • [73].Puntus LN, Lyssenko KA, Antipin MY, Bünzli J-CG, Role of inner- and outer-sphere bonding in the sensitization of EuIII-luminescence deciphered by combined analysis of experimental electron density distribution function and photophysical data, Inorg. Chem 47 (2008) 11095–11107. [DOI] [PubMed] [Google Scholar]
  • [74].Zhang Y, Ali B, Wu J, Guo M, Yu Y, Liu Z, Tang J, Construction of metallosupramolecular coordination complexes: from lanthanide helicates to octahedral cages showing single-molecule magnet behavior, Inorg. Chem 58 (2019) 3167–3174. [DOI] [PubMed] [Google Scholar]
  • [75].Dong H-M, Li H-Y, Zhang Y-Q, Yang E-C, Zhao X-J, Magnetic relaxation dynamics of a centrosymmetric Dy2 single-molecule magnet triggered by magnetic-site dilution and external magnetic field, Inorg. Chem 56 (2017) 5611–5622. [DOI] [PubMed] [Google Scholar]
  • [76].Wang K, Chen Z-L, Zou H-H, Zhang S-H, Li Y, Zhang X-Q, Sun W-Y, Liang F-P, Diacylhydrazone-assembled {Ln11} nanoclusters featuring a “double-boats conformation” topology: synthesis, structures and magnetism, Dalton Trans. 47 (2018) 2337–2343. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Table S1-S8-FS1-FS16-

RESOURCES