Abstract

The phosphonium-decorated phenanthro-imidazolyl pyridine ligand, LP+Br, readily reacts with zinc(II) and cadmium(II) bromides to give inorganic–organic zero-dimensional compounds [LP+ZnBr2]2[ZnBr4] (1) and [(LP+)2Cd2Br4][CdBr4] (2), respectively, upon crystallization. These salts are moderately fluorescent in the solid state under ambient conditions (λem = 458 nm, Φem = 0.11 for 1; λem = 460 nm, Φem = 0.13 for 2). Their emission results from spin-allowed electronic transitions localized on the organic component with the negligible effect of [MBr4]2– and MBr2 units. Contrary to ionic species 1 and 2, lead(II) bromide affords a neutral and water-stable complex [(LP+)2Pb3Br8] (3), showing weak room-temperature phosphorescence arising from spin–orbit coupling due to the heavy atom effect. The emission, which is substantially enhanced for the amorphous sample of 3 (λem = 575 nm, Φem = 0.06), is assigned to the intraligand triplet excited state, which is a rare phenomenon among Pb(II) molecular materials.
Short abstract
A phosphonium-decorated diimine chromophore forms zero-dimensional hybrid complexes with zinc(II) and cadmium(II) bromides, which demonstrate intraligand fluorescence in the solid state. On the contrary, the use of lead(II) bromide leads to a zwitterionic compound with the [Pb3Br8]2− cluster core, showing room-temperature intraligand phosphorescence.
Introduction
Hybrid materials composed of ionic organic and metal halide components, where metal is a post-transition element from groups 12–15, demonstrate remarkably rich photophysical behavior and have been intensely investigated due to their potential optoelectronic applications, covering electroluminescent devices, solar cells, detectors, and sensors.1−5 Optical properties of such materials are generally defined by the architecture and composition of halometalate moieties along with characteristics of the lattice, which opens wide possibilities for tuning physical performance once structure–property relationships are understood. In particular, the large variability of organic ionic blocks affects connectivity within an anionic metal halide framework, which can adopt from zero- (0D) to three-dimensional (3D) topology in the crystalline state.6,7 Utilization of sterically demanding cations typically results in the formation of 0D hybrids, i.e., those comprising individual metal halide anions of mononuclear and cluster nature, spatially separated by organic “insulators”.8−11 Due to relatively nonrigid structural confinement imposed by the lattice, these compounds often exhibit broad photoemission with large Stokes shifts and can reach exceptionally high quantum yields.10,12−18
The majority of hybrid systems include nonconjugated organic cations, which play a negligible electronic role and influence photophysics by governing the crystal packing, the subtle structural features, and the local environment of the metal halide anionic components (Scheme 1A). Nevertheless, recently, there has been an increasingly growing number of reports on organic–inorganic compounds, which employ aromatic chromophores as cations having a noticeable or even dominating contribution to the luminescence of such ionic materials (Scheme 1B). Thus, the resonant energy transfer from one-dimensional (1D) lead chloride chains to organic molecules was proposed for broad-band luminescence demonstrated by [(aminoquinoline)H2][PbCl4] salt.19 The prevailing blue-to-white fluorescence originating from quaternized organic bases (various pyridines; pyrimidines; ethylene-, benzyl-, and stilbenyl amines; etc.) is observed for [MX3(dmso)] (M = Zn, X = Br; M = Cd, X = I),20 [ZnX4]2–,13,21,22 [InBr4]−,23 [PbCl5]3–,24 and [PbCl3]−25 derivatives, the quantum yields and photostability being substantially higher than those for parent organic halide salts. Alternatively, the related compounds [(aminopyridinium)][HgBr4],13 [diammoniumdiphenyl sulfone][SnCl6],26 and [bis(pyridinium)propane][Pb2X6]27 display broad-band luminescence due to the interplay of inorganic (self-trapped) and organic-localized excitons.
Scheme 1. Illustration of Inorganic–Organic Hybrids Based on Different Types of Organic Cations: (A) Aliphatic Nonemissive Cations; (B) Noncoordinating Cations with Chromophore Groups; and (C) Chromophore Cations with Coordinating Function.
Optical properties of the inorganic–organic ionic compounds are further broadened by populating the triplet excited states, leading to room-temperature phosphorescence.3,28 The lifetimes and quantum yields of organic ultralong photoemission can be readily modulated by means of the external heavy atom effect, i.e., varying the nature of the constituting halometalates, that was shown for complexes [benzoquinolinium][Pb2X5],29 [PPh4][ZnX4],30 and [PPh4][Cd2X6],31 the latter for X = Br reaches notably high Φem = 63% (τ = 37.85 ms).
Another approach to diversify the structural organization and photophysical characteristics of hybrid compounds potentially relies on the employment of cationic chromophores with coordinating function and yet remains virtually unexplored (Scheme 1C). The aromatic ligands, e.g., strongly binding N-donors (bi-/tripyridine, phenanthroline, and other heterocycles), decorated with remote positively charged groups (pyridinium/quinolinium, phosphonium, ammonium, etc.) are not exceptional and have been used in the design of a number of phosphorescent complexes of Ru, Re, Ir, and Pt for bioimaging applications,32 and some Cu(I) zwitterionic iodide clusters with delayed fluorescence.33,34 These sorts of constructing elements, however, have not penetrated the field of inorganic–organic systems built of post-transition metals, except sporadic examples like [(quinolinium terpyridine)ZnBr2][ZnBr4(dmso)],35 confirming the preparative feasibility of the given strategy.
From the photophysical viewpoint, it is well known that group 12 metal halides, which are used as starting reagents for the fabrication of halometalate-based hybrids, and the related salts MX2 (M = Zn, Cd) readily form fluorescent complexes with chelating N- and O-donors (e.g., N-heterocyclic and Schiff bases, salen-type ligands, etc.) adopting 4-, 5-, and 6-coordination numbers.36−45 Coordination of the metal centers may lead to the conformational modification46 and substantial perturbation of the electronic structure of organic molecules, resulting in, e.g., enhanced intramolecular charge transfer,39,40,47 appearance of stimuli-responsive and sensing ability,40,43−45,48,49 suppression of photoinduced electron transfer, and dramatic improvement of quantum efficiency typically accompanied by the bathochromic shift of the emission.36,39,46,47,50 On the other hand, the coordinated MX2 units tend to participate in intermolecular noncovalent interactions, which have a strong impact on the solid-state arrangement and the corresponding optical properties.41,42,51
In comparison to zinc and cadmium, the coordination chemistry of main group metals is less developed for the soft N-heterocyclic ligands and is illustrated by pseudo-octahedral complexes of Pb(II), forming 1D polymeric chains [Pb(diimine)(μ-X)2]n,52−56 the emission energy of which is regulated by aromatic system.56
To probe the viability of the concept illustrated in Scheme 1C and to utilize the cationic chromophore ligand in the synthesis of low-dimensional metal hybrids, as a case study, we have chosen easy-to-obtain and -modify phenanthrene fused with the chelating imidazolyl-pyridine motif, which readily coordinates to late transition metals as was earlier shown by us and other groups.42,57,58 The decoration of this fluorescent diimine ligand with the chemically robust tetraarylphosphonium group affords a bulky metal-binding cation, which was explored in the preparation of hybrid compounds derived from Zn, Cd, and Pb halide units.
Experimental Section
General Comments
1-(4-Bromophenyl)-2-(pyridin-2-yl)-1H-phenanthro[9,10-d]imidazole (LBr) was prepared according to the reported procedure.58 Other reagents were used as received. Toluene was distilled over Na-benzophenone ketyl under a nitrogen atmosphere prior to use. The solution 1H and 31P{1H} spectra were recorded on a JEOL 500 spectrometer. Microanalyses were carried out in the analytical laboratory of the University of Eastern Finland.
Triphenyl(4-(2-(pyridin-2-yl)-1H-phenanthro[9,10-d]imidazol-1-yl)phenyl) Phosphonium Bromide (LP+Br)
The synthesis was carried out in a pressure tube following the general protocol for the preparation of tetraarylphosphonium salts.59 The tube was charged with solid LBr (0.5 g, 1.11 mmol), triphenylphosphine (0.291 g, 1.11 mmol), and Pd2(dba)3 (6.4 mg, 0.007 mmol), which were degassed, and dry toluene (0.7 mL) was added under a nitrogen atmosphere. The reaction mixture was heated to 140 °C for 12 h, resulting in the gradual formation of a precipitate of the phosphonium salt. The suspension was cooled down to room temperature, the solvent was removed by filtration, and the crude product was washed with diethyl ether (2 mL × 5 mL). The solid was dried and further purified by column chromatography (silica gel 70–230 mesh, ⌀4 cm × 15 cm, eluent dichloromethane/methanol 40:1 → 10:1 v/v mixture) to afford a pale creamy amorphous material after evaporation of the volatiles (688 mg, 87%). 1H NMR (acetonitrile-d3, 298 K; δ): 8.86 (d, JHH 8.0 Hz, 1H), 8.79 (d, JHH 8.0 Hz, 1H), 8.75 (dd, JHH 7.8, 1.2 Hz, 1H), 8.39 (dt, JHH 7.9, 1.0 Hz, 1H), 8.24 (ddd, JHH 4.8, 1.8, 0.9 Hz, 1H), 7.97–7.92 (m, 5H), 7.92–7.86 (m, 3H), 7.82–7.74 (m, 13H), 7.70 (ddd, JHH 8.4, 7.1, 1.5 Hz, 1H), 7.60 (ddd, JHH 8.4, 7.0, 1.3 Hz, 1H), 7.39 (ddd, JHH 8.3, 7.0, 1.2 Hz, 1H), 7.31 (ddd, JHH 7.6, 4.8, 1.2 Hz, 1H), 7.22 (dd, JHH 8.4, 1.0 Hz, 1H). 31P{1H} NMR (acetonitrile-d3, 298 K; δ): 23.7 (s). 1H NMR (N,N-dimethylformamide-d7, 298 K; δ): 9.01 (d, JHH 8.2 Hz, 1H), 8.95 (d, JHH 8.4 Hz, 1H), 8.80 (dd, JHH 8.0, 1.3 Hz, 1H), 8.48 (dt, JHH 7.9, 1.0 Hz, 1H), 8.34 (ddd, JHH 4.8, 1.7, 0.9 Hz, 1H), 8.29–8.25 (m, 2H), 8.20–8.14 (m, 2H), 8.10–7.94 (m, 15H), 7.88–7.81 (m, 2H), 7.75 (ddd, JHH 8.4, 7.0, 1.5 Hz, 1H), 7.64 (ddd, JHH 8.3, 7.0, 1.2 Hz, 1H), 7.45–7.38 (m, 2H), 7.28 (dd, JHH 8.3, 1.0 Hz, 1H). ESI-MS (m/z): [M]+ 623.2267 (calcd 623.2256). Anal calcd for C44H31BrN3P: C, 74.16; H, 4.38; N, 5.90. Found: C, 73.84; H, 4.69; N, 5.82.
[LP+ZnBr2]2[ZnBr4] (1)
A solution of ZnBr2 (47 mg, 0.21 mmol) in ethyl acetate (5 mL) was added to a solution of LP+Br (100 mg, 0.14 mmol) in acetonitrile (20 mL). The reaction mixture was stirred for 30 min, and a nearly transparent solution was filtered and left at room temperature for slow evaporation to give a white crystalline material (124 mg, 84%). 1H NMR (acetonitrile-d3, 1.69 × 10–3 M, 298 K; δ): δ 8.93 (d br, JHH 7.8 Hz, 1H), 8.90 (d, JHH 8.3 Hz, 1H), 8.85 (d, JHH 8.3 Hz, 1H), 8.70 (s br, 1H), 8.19–8.05 (m, 4H), 8.00–7.93 (m, 4H), 7.88–7.77 (m, 15H), 7.70–7.64 (m, 2H), 7.43 (ddd, JHH 8.2, 7.1, 1.0 Hz, 1H), 7.05 (d, JHH 7.4 Hz, 1H). 31P{1H} NMR (acetonitrile-d, 298 K; δ): 23.8 (s). Anal calcd for C88H62Br8N6P2Zn3·CH3CN: C, 50.47; H, 3.06; N, 4.58. Found: C, 50.63; H, 3.11; N, 4.76.
[(LP+)2Cd2Br4][CdBr4] (2)
CdBr2 (42 mg, 0.15 mmol) and LP+Br (71 mg, 0.10 mmol) were suspended in a mixture of acetonitrile (12 mL) and methanol (6 mL). The suspension was heated at ca 60 °C under stirring until complete dissolution of the solids (ca 15 min). The resulting solution was filtered and concentrated by slow evaporation at room temperature to give a white crystalline material (2), which was washed with acetonitrile (5 mL) and diethyl ether (2 mL × 5 mL) and dried (94 mg, 85%). 1H NMR (acetonitrile-d3, 298 K; δ): 8.89 (d, JHH 6.8 Hz, 1H), 8.85 (d, JHH 8.2 Hz, 1H), 8.79 (dd, JHH 8.1, 1.3 Hz, 1H), 8.69 (s br, 1H), 7.99–7.92 (m, 7H), 7.89 (td, JHH 7.8, 1.7 Hz, 1H), 7.83–7.77 (m, 13H), 7.75–7.69 (m, 2H), 7.63 (ddd, JHH 8.3, 7.1, 1.2 Hz, 1H), 7.53 (dd, JHH 8.7, 3.6 Hz, 1H), 7.38 (ddd, JHH 8.2, 7.1, 1.1 Hz, 1H), 7.09 (d, JHH 8.4 Hz, 1H). 31P{1H} NMR (acetonitrile-d3, 298 K; δ): 23.8 (s). Anal calcd for C88H62Br8Cd3N6P2: C, 47.15; H, 2.79; N, 3.75. Found: C, 47.08; H, 2.85; N, 3.91.
[(LP+)2Pb3Br8] (3)
A solution of LP+Br (70 mg, 0.10 mmol) in acetonitrile (1.5 mL) was added to a solution of PbBr2 (51 mg, 0.14 mmol) in dimethylformamide (3 mL). The resulting mixture was gently refluxed for ca 10 min, giving a nearly transparent pale yellow solution, which was filtered. Gas-phase diffusion of water into this solution at room temperature for 1 week produced 3 as a pale yellow crystalline material, which was washed with acetonitrile (2 mL × 5 mL) and diethyl ether (2 mL × 5 mL) and dried (85 mg, 71%). Anal calcd for C88H62Br8N6P2Pb3: C, 41.84; H, 2.47; N, 3.33. Found: C, 42.04; H, 2.55; N, 3.75.
X-ray Structure Determinations and Powder Measurements
The crystals of 1–3 were immersed in cryo-oil, mounted in a Nylon loop, and measured at a temperature of 150 K. The X-ray diffraction data were collected with a Bruker Kappa Apex II diffractometer using Mo Kα (λ = 0.71073 Å) radiation. The APEX260 program package was used for cell refinements and data reductions. A numerical or semiempirical absorption correction (SADABS)61 was applied to all data. The structures were solved by direct methods using the SHELXS-201862 program with the WinGX63 graphical user interface. Structural refinements were carried out using SHELXL-2018.62
Some of the crystallization solvent molecules in 2 could not be resolved unambiguously. The acetonitrile solvent molecules were refined with an occupancy of 0.5. The contribution of the missing solvent to the calculated structure factors was taken into account by a SQUEEZE routine64 of PLATON.65 The missing solvent was not taken into account in the unit cell content. All non-H atoms were anisotropically refined, and all hydrogen atoms were positioned geometrically and constrained to ride on their respective parent atoms with C–H = 0.95–0.98 Å and Uiso = 1.2–1.5Uequiv (parent atom). The crystallographic details are summarized in Table S1 (Supporting Information).
Powder X-ray powder diffraction (XRD) patterns were recorded with a Bruker Advance D8 diffractometer using a Cu Kα (1.54184 Å) radiation source (Bruker: 40 kV/ 40 mA). Divergence and receiving slits of 1.0 mm were used together with a Ni filter in the measurements. Prior to measurement, a sample was placed on a Si single-crystal zero background sample holder. The diffraction patterns were scanned from 5 to 90° in a 2θ scale using the locked couple technique in Bragg–Brentano geometry. A collection time of 2 s was used together with a step size of 0.05° per step.
Photophysical Measurements
The excitation and emission spectra for the solid samples were recorded with an Edinburgh FLS980 spectrofluorometer. Photoluminescence quantum yields and the absorption spectra of solid materials were measured by an integrating sphere (F-M01, Edinburgh) set in the sample chamber of the spectrometer (Edinburgh FLS980). Emission lifetimes within 1 ns–20 μs were collected by an inbuilt, time-correlated single photon counting (TCSPC) system (Edinburgh FLS980) coupled with a synchronized diode laser (EPL-375) as the pumping source. For phosphorescence lifetimes longer than 50 μs, an intensified charge-coupled detector (PI-MAX ICCD) coupled with a Q-switch laser was used instead. Both timings of ICCD and tunable laser are triggered by a pulse generator (DG-535, Stanford Research System) with signal jittering less than 5 ns. The excitation pulse was generated by an LT-2134 (532 nm, SHG of Nd:Yag laser, LOTIS Tii) followed by an LT-2211 tunable laser (345–532 nm, LOTIS Tii), and 365 nm was chosen as the excitation source with an fwhm of ∼20 ns.
Computational Details
Quantum chemical studies on all complexes were carried out using density functional theory (DFT). The PBE0 hybrid density functional66,67 in combination with the def2-TZVP basis set68 and the corresponding scalar-relativistic effective core potential69 were applied for Cd and Pb atoms. The full model of Cd complex 2 was optimized using a QM/MM approach within the ONIOM framework,70 since attempts to locate a minimum geometry with the isolated single-molecule model were unsuccessful. In the QM/MM model, the crystal structure was expanded in three dimensions and the central molecule was assigned as the QM part, while the surroundings were kept frozen and treated with MM using the UFF force field.71 Electronic embedding with the QEq scheme72 was used to account for polarization of the QM region. The QM/MM optimization was performed using Gaussian 16.73 All other calculations were performed with Orca 5.0.374 using single-molecule models in the optimizations. The geometry of the first excited singlet state was optimized with TD-DFT, while the ground state and first excited triplet state were optimized with DFT. For compound 3, the implicit CPCM solvation model75 with toluene as the solvent was used because the geometry optimization did not converge in a vacuum environment. To confirm that the obtained geometries corresponded to a true local minimum on the potential energy surface, frequency calculations were performed for all optimized structures. Furthermore, to study the excitation and emission behavior, single-point TD-DFT calculations were performed for all optimized structures using the range-separated LRC-ωPBEh76 functional together with the def2-TZVP basis set. The resolution of identity approximation together with the chain of spheres for exchange approximation (RIJCOSX)77,78 was used in all calculations with Orca to speed up the calculations.
Results and Discussion
Synthesis and Characterization
The tetraarylphosphonium unit was chosen due to its stability, bulkiness, and accessible synthesis, the conditions of which tolerate the N-heterocycles. Functionalization of the phenanthro-imidazolyl pyridine was performed according to the general method by Marcoux (Scheme 2 and the Supporting Information).59 The palladium-catalyzed reaction of the bromoaryl precursor (LBr)58 with a stoichiometric amount of triphenylphosphine afforded the corresponding phosphonium bromide (LP+Br) in good yield. Hybrid complexes of zinc(II) and cadmium(II) were obtained by treating the ZnBr2 and CdBr2 salts with LP+Br in ethyl acetate/acetonitrile and methanol/acetonitrile mixtures, respectively. Slow evaporation of the resulting solutions at room temperature produced colorless crystals of ionic species [LP+ZnBr2]2[ZnBr4] (1) and [(LP+)2Cd2Br4][CdBr4] (2) suitable for X-ray diffraction analysis. Due to the lower solubility of PbBr2, it was reacted with LP+Br in a dimethylformamide/acetonitrile solvent system. Subsequent gas-phase diffusion of water into the reaction mixture gave a neutral complex [(LP+)2Pb3Br8] (3) as a pale yellow crystalline material (Scheme 2). The formation of 1 and 3 was found to be selective and rather independent of the ratio of starting reagents, while in the case of 2, proper stoichiometry (LP+Br/CdBr2 2:3) has to be obeyed to minimize crystallization of side products. A slight excess of LP+Br in the synthesis of 3 was used to avoid precipitation of PbBr2. The iodide analogues of 1 and 2 can also be obtained in a similar fashion, but the crystalline materials visibly degrade under exposure to light and are virtually nonluminescent under ambient conditions and thus were omitted from this work. In yet another approach, we were unable to prepare the lead iodide complex in a reproducible manner and in pure form.
Scheme 2. Synthesis of the Cationic Diimine Ligand and Its Zn, Cd, and Pb Complexes.
The structures of 1–3 determined by XRD analysis are depicted in Figures 1–3, and selected bond lengths and angles are listed in Table S2 (SI). The repeating motif of salt 1 is composed of individual inorganic dianions [ZnBr4]2– and two crystallographically nonequivalent metal–organic cations [LP+ZnBr2]+. In turn, the latter constituents comprise the phosphonium-decorated ligand, coordinated to the zinc dibromide unit via chelating imidazolyl-pyridine function. In this cationic complex, the metal center expectedly adopts pseudo-tetrahedral coordination geometry, analogously to the neutral predecessor compounds we described earlier42 and to other related [Zn(diimine)Hal2] species.37,41,46,79
Figure 1.

Molecular (top) and packing extended (middle and bottom) structures of salt [LP+ZnBr2]2[ZnBr4] (1); thermal ellipsoids are shown at 50% probability.
Figure 3.

Molecular (top) and packing extended (bottom) structures of complex [(LP+)2Pb3Br8] (3); thermal ellipsoids are shown at 50% probability.
Cadmium complex 2 reveals the same stoichiometry as that found in 1, [LP+MBr2]2[MBr4], and is topologically similar to the zinc congener (Figure 2). The slightly distorted [CdBr4]2– tetrahedron is placed in voids formed by tetraarylphosphonium groups, which is apparently driven by the electrostatic attraction. The [LP+CdBr2]+ cations dimerize via the nonsymmetrical bromide bridges (μ2-Br–Cd distances are 2.640(2), 2614(2), 2.861(2), and 2.802(2) Å; Table S2) that provides pentacoordinate environment of Cd(II) ions and evidently follows a preference of [Cd(diimine)Hal2] units to have a higher coordination number than that of Zn(II) analogues.80−82
Figure 2.

Molecular (top) and packing extended (bottom) structures of salt [(LP+)2Cd2Br4][CdBr4] (2); thermal ellipsoids are shown at 50% probability.
The [MBr4]2– anions in 1 and 2 form networks of hydrogen bonds with organic molecules and are well separated from each other (Figures 1 and 2); the distances between the bromine atoms of adjacent anions exceed 10 Å (1) and 6.6 Å (2), which are likely too long for appreciable electronic coupling between these inorganic components.
In contrast to ionic compounds 1 and 2, lead(II) bromide produced overall neutral molecular entity 3, which can be seen as a zwitterionic complex built of two cationic ligands coordinated to an anionic bromoplumbate cluster featuring a discrete [Pb3Br8]2– core (Figure 3); the complex conceptually resembles the [Cu4I6]2+ clusters stabilized by [RP(pyridine)3]+ ligands.33 The anions of the same composition but of somewhat different geometry were found as 1D polymers in salts [PR4]2[Pb3Br8].83,84 The trimetallic motif contains a central seesaw [PbBr4]2– fragment, which is linked to the lateral PbBr2 units via bromide bridges. The geometry of tetracoordinate [PbBr4]2– is known for ns2 metal halides2,9 and is derived from a trigonal bipyramid with a 6s2 electron pair of Pb(II) being in the equatorial plane.85,86 The bond distances from the bridging halides Br(3) and Br(4) to the pendant Pb(1) atom (Br(3/4)–Pb(1) = 3.0893(4) and 3.2068(8) Å) are visibly longer than those within the [PbBr4]2– block (Br(3/4)–Pb(2) = 2.9932(4) and 2.7250(5) Å). Such deviation of μ2-Br–Pb separations due to nonsymmetrical coordination has been described for other bromoplumbate clusters.87,88
The binding of the LP+ ligand to each of the PbBr2, together with μ2-bromides, completes the highly distorted octahedral environment of the Pb(1) center. This metal atom lies almost in the plane of the pyridine; the Pb(1)–Npy(1) bond length (2.642(3) Å) indicates relatively strong interaction and is consistent with reported data on the Pb(II) diimine halide complexes.52,54 Bonding of the Pb(1) atom to the diimine function is accompanied by a substantial twisting of the pyridine ring with respect to the imidazole; the corresponding dihedral angle between the planes of these heterocycles equals 33.8° (Figure 3). This might be caused, to a large extent, by intermolecular arrangement, which is probably defined by multiple Br···H contacts (2.73–3.10 Å) due to the Coulombic forces between the bromide ligands and the positively charged tetraarylphosphonium motifs. As a result, the Pb(1) atom is substantially displaced from the plane of imidazole, leading to a longer Pb(1)–Nimi(2) distance (2.759(3) Å). This effect can be attributed to the proximity of the lone pair of Pb(II), which causes the elongation and weakening of the Pb–Nimi bonding interaction.89
The 1H NMR spectra recorded for compounds 1 and 2 in acetonitrile-d3 at room temperature confirm the presence of complex cations [LP+MBr2]+ in solution (Figure S1), although broadening of the signals for salt 1 upon dilution suggests some dynamic process. This is likely caused by easy dissociation of MBr2 units in coordinating solvent and is supported by electrospray ionization (ESI)-mass spectrometry (MS) measurements, which show only the signal of the free ligand. Complex 3 is scarcely soluble in most organic solvents. Its proton spectrum in a dimethylformamide-d6 solution is identical to that of the free ligand, indicating complete disintegration of the aggregate. Nevertheless, the dissociation does not lead to degradation of the mixture and 3 is efficiently reassembled upon crystallization or slow precipitation with water.
Photophysical Studies and Theoretical Analysis
For the free ligand in a dichloromethane solution, the lowest energy absorption bands (330–360 nm, Figure S2) can be assigned to the predicted S0 → S1 transition (λ = 340 nm) from density functional theory (DFT), which mainly involves charge transfer (ILCT) from imidazole to phosphonium-bearing phenylene and the empty orbital of the P–C bond (Figure S3). The radiative relaxation to the ground state S1 → S0 (λ = 424 nm, Table S3) has virtually identical CT character. In accordance with theoretical results, the observed ILCT fluorescence of LP+Br (λem = 422 nm, Φem = 0.1, Figure S2) lacks vibronic progression and demonstrates substantial Stokes shift. This behavior markedly differs from that of the neutral analogue L (2-pyridyl-1H-phenanthro[9,10-d]imidazole) with largely phenanthrene-centered structured emission in the UV region (λem = 370, 388 nm, Φem = 0.37).58
Dried crystals of acetonitrile-solvated 1 and 2 show similar room-temperature luminescence (Figure 4 and Table 1). Both compounds are moderate blue emitters, which exhibit broad and structureless signals maximized at 458 (1) and 460 nm (2) with quantum yields of 0.11 and 0.13, respectively. The luminescence of the ligand salt LP+Br in the solid (λem = 456 nm, Φem = 0.04, Figure S2) is substantially weaker, although the peak wavelength is practically unaffected by the presence of the metal units. The average lifetimes (τav = 1.3 and 1.4 ns) and radiative rate constants (kr ≈ ca 108 s–1) point to the singlet origin in the excited state, i.e., fluorescence. A similar enhancement of organic luminescence in the hybrid compounds due to a more rigid environment has been reported for other congener materials.21,22,24
Figure 4.
Room-temperature absorption, excitation, and emission (λexc = 380 nm) spectra of 1 (A) and 2 (B) in the solid state.
Table 1. Photophysical Properties of Compounds 1–3 in the Solid State at 298 K.
| λfluo, nm | τfluo, ns | λphos, nm | τphos, μs | Φem | kr,b s–1 | knr,b s–1 | |
|---|---|---|---|---|---|---|---|
| LP+Br | 456 | <1 | 0.04 | >4 × 107c | >9 × 108c | ||
| 1 | 458 | 1.3a | 0.11 | 8.5 × 107 | 6.8 × 108 | ||
| 2 | 460 | 1.4a | 0.13 | 9.3 × 107 | 6.2 × 108 | ||
| 3cryst | 440 | 2.1 | 517, 550 | 11 | 0.009 | ∼8 × 102d | ∼9 × 104 kd |
| 3amorph | 460 | 4.0 | 575 | 120 | 0.06 | ∼5 × 102d | ∼8 × 103d |
Amplitude-weighted average emission lifetimes for the biexponential decays determined by the equation τav = ∑Aiτi, Ai = weight of the ith exponent.
kr and knr were estimated by Φem/τ and (1 – Φem)/τ, respectively.
Estimated approximately due to the inability to accurately determine the lifetime.
Calculated by neglecting the contribution of fluorescence to the value of Φem.
Albeit relatively short intermolecular Br···π contacts (ca 3.3–3.4 Å, Figure 1) seen in salt 1, the phosphorescence induced by the external heavy atom effect, which was previously detected for LZnI2,42 cannot be identified in the room-temperature spectra of 1. The spectroscopic profiles for 1 and 2 resemble those of the complexes LZnX2 (X = Cl, I, OAc).42 Thus, the excitation and emission bands of 1 and 2 are likely associated with the [LP+MBr2]+ cations and are dominated by the organic component, whereas the [MBr4]2– anions do not contribute to the S1 state. DFT studies of the [LP+MBr2]+ components (the half model was used for the Cd derivative, Table S3) confirm that the S0 ↔ S1 electronic transitions have the same nature as those for the free ligand (Figures 5 and S4) and illustrate the insignificant influence of coordinated MBr2 units on the optical behavior of 1 and 2. Compared to the neutral species LZnX2, for which the π(phenanthrene) → π*(pyridine) charge transfer was computationally predicted for the low-energy excitation, the appearance of the electron-deficient phosphonium motif decreases the contribution from phenanthrene and pyridine orbitals, substantially changing the character of excitation and emission to π(imidazole) ↔ π*(phenylene-P) transitions.
Figure 5.
Electron density difference plots for the lowest energy excitation S0 → S1 and emission S1 → S0 in the cation [LP+ZnBr2]+ of compound 1 (isovalue 0.002 a.u., DFT-LRC-ωPBEh method, optimized S0 and S1 geometries). During the electronic transition, the electron density increases in the red areas and decreases in the blue areas.
Single crystals of complex 3 are poorly luminescent at room temperature under UV excitation, showing the main low-energy band maximized at 550 nm and a very weak high-energy signal around 440 nm (Figure 6) with a total quantum yield of 9 × 10–3 (Table 1). The latter minor band has a lifetime of 2.1 ns and can be correlated with the intraligand fluorescence of 1 and 2. The microsecond lifetime of the prevailing band (τobs = 11 μs) and the corresponding rate constant (kr ≈ 8 × 102 s–1) indicate a spin-forbidden process, i.e., phosphorescence. The barely resolved but notwithstanding discernible vibrational structure of the main band points to the triplet state responsible for the emission is localized within the conjugated system of the diimine ligand. This tentative assignment is also implicitly supported by the similarity and broadness of excitation spectra for 1, 2, and 3; for self-trapped exciton emission of zero-dimensional halides, relatively narrow excitation spectra have been typically reported.13,14,85,88,90
Figure 6.
Room-temperature absorption, excitation, and emission (λexc = 400 nm) spectra of the single crystals (A) and amorphous sample (B) of complex 3.
Remarkably, thorough grinding of single crystals of 3 or vacuum evaporation of a dimethylformamide solution of 3 generates a material with much brighter luminescence. The predominantly amorphous nature of the solid is confirmed by powder XRD measurement (Figure S5). The crystal-to-amorphous phase transition and the decrease of the crystal size do not considerably affect the excitation spectra but red-shift the maxima of both emission bands for ca 20–25 nm, retaining their approximate ratio. This is accompanied by a nearly 7-fold increase in the total intensity (Φem = 0.06) and the disappearance of the fine structure of the dominating low-energy band. The latter indicates a more significant charge transfer character of the triplet state in the amorphous sample than that in bulk crystals. The gain in quantum yield arises from the suppression of nonradiative decay that is manifested by nearly an order of magnitude drop of the corresponding rate constant for amorphous 3 (Table 1). Despite the fact that the disruption of the lattice is expected to produce a less rigid local environment, the decrease in knr probably results from the disappearance of intermolecular interactions between organic chromophores (Figure 3), which could quench the luminescence in the neat crystal, making this effect opposite to the one termed “crystallization-induced emission”, e.g., in [PbX2(4,4′-bipyridine N-oxide)]n coordination polymers.91 The switch of weak fluorescence in crystals to stronger phosphorescence in the glass phase was observed for lead(II) alkanoates and explained by the shortening of intermolecular Pb–Pb distances.92 Importantly, the dominating phosphorescence of amorphous 3 indicates that it is an intrinsic molecular property but not an effect of crystal packing. Furthermore, it confirms an efficient assembly of this complex upon solvent removal, notwithstanding its complete dissociation in solution that opens a way for facile solution processing of these species. In terms of potential practical applications of this sort of compound, the improvement of the photophysical performance in an amorphous state could simplify the fabrication and handling of uniform films for optoelectronic devices.
The presence of a weak prompt fluorescence signal propounds that S1 → S0 radiative relaxation competes with relatively slow intersystem crossing, eventually induced by the heavy atom effect of the lead bromide fragment. The same phenomena were observed in the 77 K spectra for both amorphous 3 and crystalline 3 (Figure S6). Their emission intensities increase significantly by lowering the temperature, which suppresses the thermally induced nonradiative deactivations. For amorphous 3, its low-temperature fluorescence and phosphorescence bands are blue-shifted ca 20 nm compared to those at room temperature and indicate excited-state destabilization at low temperatures, plausibly due to hindered intermolecular relaxation in the amorphous phase under cryogenic conditions. Support for this viewpoint is given by crystalline 3, where the intermolecular relaxation is insignificant, and therefore, the fluorescence and phosphorescence peaks remain unchanged by varying the temperature.
Theoretical analysis of 3 was performed using the implicit CPCM solvation model75 in the DFT calculations because the geometry optimization did not converge in a vacuum environment. The inclusion of toluene as a solvent helped in reaching satisfactory optimization of the geometries of S0 and T1 states. As the use of solvent is not justified experimentally, because it is absent in the crystal structure of 3, the character of the predicted S0 → S1 excitation yet remains ambiguous. Given the similarity of the experimental excitation spectra for compounds 1–3, it is reasonable to assume photoinduced IL charge transfer transition (imidazole → phenylene) for 3 as well. The T1 state for 3 seems to be mostly insensitive to the choice of solvent, and the T1 → S0 emission is governed by the electronic transitions of one of the diimine ligands, which occur within the imidazole ring and the nearest bonds around it (Figure 7). The calculated wavelength for the phosphorescence of 3 (λcalc = 557 nm, Table S3) matches well the observed ones. Regarding the intraligand triplet emission of 3, it is worth mentioning that such behavior is rare among lead compounds. Long-lived organic phosphorescence, as a rule, with millisecond lifetimes, has been abundantly reported for some Pb(II) coordination polymers,91 constructed mainly from aromatic carboxylate blocks.92−96 Alternatively, efficient room-temperature phosphorescence of organic components has been realized in lead perovskites by introducing chromophore cations29,97,98 or sensitizing agents.99 However, on a molecular scale, complex 3 is one of the so far very few molecular lead compounds exhibiting phosphorescence at ambient temperature.100,101
Figure 7.

Electron density difference plot for the lowest energy triplet emission T1 → S0 in complex 3 (isovalue 0.002 a.u., DFT-LRC-ωPBEh method using the CPCM solvation model75 with toluene as a solvent, optimized T1 geometry). During the electronic transition, the electron density increases in the red areas and decreases in the blue areas.
Conclusions
In this work, we have demonstrated the strategy of using the cationic diimine ligand, phosphonium-functionalized phenanthro-imidazolyl pyridine LP+Br, for the straightforward preparation of inorganic–organic hybrid complexes of zinc (1), cadmium (2), and lead (3). Compounds 1 and 2 represent zero-dimensional salts containing ligand-derived complex cations [LP+MBr2]+ and bromometalate anions [MBr4]2–. The lead bromide resulted in the assembly of a novel type of zwitterionic complex, in which the anionic cluster [Pb3Br8]2– is stabilized by two coordinated motifs LP+. The analysis of the photophysical behavior of 1–3 in the solid state showed that the properties of 1 and 2 are governed by the intraligand charge transfer, resulting in moderate blue fluorescence, which is visibly enhanced versus the parent organic salt LP+Br. On the other hand, complex 3 demonstrates a weak fluorescence signal along with room-temperature phosphorescence, which is a very rare case for Pb(II) molecular compounds. The phosphorescence mainly originates from an intraligand transition, while the lead and bromine atoms act as external heavy elements to boost the spin–orbit coupling. Notably, weak luminescence of crystalline 3 is drastically enhanced in the amorphous state, retaining the dominating role of the triplet excited state. Essentially, it allows ascribing the observed photoemission of 3 to the molecular properties but not to the consequence of crystal packing. Thus, it potentially opens ways for new structural chemistry and for tuning the optical characteristics of related hybrid and zwitterionic species by employing positively charged organic ligands possessing different chromophore fragments and a variable number of cationic groups. Despite presenting a case study of one particular ligand, the described concept likely can be applied to a large selection of organic ligands that will allow generation of novel ionic crystalline materials, zwitterionic aggregates, and heterometallic systems with rich photophysical behavior. Although 3 undergoes complete dissociation in solution, it is efficiently assembled upon solvent removal, which is presumably facilitated by the electrostatic attraction of inorganic and organic blocks. No less important is the stability of 3 toward water that enables convenient fabrication, handling, and solution processing of the complex without special precautions.
Acknowledgments
Financial support from the Academy of Finland (decision 317903, I.O.K.; decision 340584, T.E. and A.J.K.; Flagship Programme, Photonics Research and Innovation PREIN, decision 320166) and computational resources from the Finnish IT Center for Science (CSC) are gratefully acknowledged.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.inorgchem.2c02867.
Crystal data and structure refinement for 1–3; selected structural parameters for 1–3; and additional spectroscopic and computational data and figures (PDF)
Author Contributions
The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript.
The authors declare no competing financial interest.
Supplementary Material
References
- Parola S.; Julián-López B.; Carlos L. D.; Sanchez C. Optical Properties of Hybrid Organic-Inorganic Materials and their Applications. Adv. Funct. Mater. 2016, 26, 6506–6544. 10.1002/adfm.201602730. [DOI] [Google Scholar]
- McCall K. M.; Morad V.; Benin B. M.; Kovalenko M. V. Efficient Lone-Pair-Driven Luminescence: Structure–Property Relationships in Emissive 5s2 Metal Halides. ACS Mater. Lett. 2020, 2, 1218–1232. 10.1021/acsmaterialslett.0c00211. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao R.; Kodaimati M. S.; Yan D. Recent advances in persistent luminescence based on molecular hybrid materials. Chem. Soc. Rev. 2021, 50, 5564–5589. 10.1039/D0CS01463J. [DOI] [PubMed] [Google Scholar]
- Han Y.; Yue S.; Cui B.-B. Low-Dimensional Metal Halide Perovskite Crystal Materials: Structure Strategies and Luminescence Applications. Adv. Sci. 2021, 8, 2004805 10.1002/advs.202004805. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ganose A. M.; Savory C. N.; Scanlon D. O. Beyond methylammonium lead iodide: prospects for the emergent field of ns2 containing solar absorbers. Chem. Commun. 2017, 53, 20–44. 10.1039/C6CC06475B. [DOI] [PubMed] [Google Scholar]
- Lin H.; Zhou C.; Tian Y.; Siegrist T.; Ma B. Low-Dimensional Organometal Halide Perovskites. ACS Energy Lett. 2018, 3, 54–62. 10.1021/acsenergylett.7b00926. [DOI] [Google Scholar]
- Kundu J.; Das D. K. Low Dimensional, Broadband, Luminescent Organic-Inorganic Hybrid Materials for Lighting Applications. Eur. J. Inorg. Chem. 2021, 2021, 4508–4520. 10.1002/ejic.202100685. [DOI] [Google Scholar]
- Li M.; Xia Z. Recent progress of zero-dimensional luminescent metal halides. Chem. Soc. Rev. 2021, 50, 2626–2662. 10.1039/D0CS00779J. [DOI] [PubMed] [Google Scholar]
- Zhou C.; Xu L.-J.; Lee S.; Lin H.; Ma B. Recent Advances in Luminescent Zero-Dimensional Organic Metal Halide Hybrids. Adv. Opt. Mater. 2021, 9, 2001766 10.1002/adom.202001766. [DOI] [Google Scholar]
- Zhou L.; Liao J.-F.; Kuang D.-B. An Overview for Zero-Dimensional Broadband Emissive Metal-Halide Single Crystals. Adv. Opt. Mater. 2021, 9, 2100544 10.1002/adom.202100544. [DOI] [Google Scholar]
- Sun S.; Lu M.; Gao X.; Shi Z.; Bai X.; Yu W. W.; Zhang Y. 0D Perovskites: Unique Properties, Synthesis, and Their Applications. Adv. Sci. 2021, 8, 2102689 10.1002/advs.202102689. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou C.; Worku M.; Neu J.; Lin H.; Tian Y.; Lee S.; Zhou Y.; Han D.; Chen S.; Hao A.; Djurovich P. I.; Siegrist T.; Du M.-H.; Ma B. Facile Preparation of Light Emitting Organic Metal Halide Crystals with Near-Unity Quantum Efficiency. Chem. Mater. 2018, 30, 2374–2378. 10.1021/acs.chemmater.8b00129. [DOI] [Google Scholar]
- Yangui A.; Roccanova R.; McWhorter T. M.; Wu Y.; Du M.-H.; Saparov B. Hybrid Organic–Inorganic Halides (C5H7N2)2MBr4 (M = Hg, Zn) with High Color Rendering Index and High-Efficiency White-Light Emission. Chem. Mater. 2019, 31, 2983–2991. 10.1021/acs.chemmater.9b00537. [DOI] [Google Scholar]
- Li Q.; Chen Z.; Li M.; Xu B.; Han J.; Luo Z.; Tan L.; Xia Z.; Quan Z. Pressure-Engineered Photoluminescence Tuning in Zero-Dimensional Lead Bromide Trimer Clusters. Angew. Chem., Int. Ed. 2021, 60, 2583–2587. 10.1002/anie.202009237. [DOI] [PubMed] [Google Scholar]
- Morad V.; Yakunin S.; Benin B. M.; Shynkarenko Y.; Grotevent M. J.; Shorubalko I.; Boehme S. C.; Kovalenko M. V. Hybrid 0D Antimony Halides as Air-Stable Luminophores for High-Spatial-Resolution Remote Thermography. Adv. Mater. 2021, 33, 2007355 10.1002/adma.202007355. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qi Z.; Chen Y.; Guo Y.; Yang X.; Zhang F.-Q.; Zhou G.; Zhang X.-M. Broadband white-light emission in a one-dimensional organic–inorganic hybrid cadmium chloride with face-sharing CdCl6 octahedral chains. J. Mater. Chem. C 2021, 9, 88–94. 10.1039/D0TC04731G. [DOI] [Google Scholar]
- Wei J.-H.; Liao J.-F.; Zhou L.; Luo J.-B.; Wang X.-D.; Kuang D.-B. Indium-antimony-halide single crystals for high-efficiency white-light emission and anti-counterfeiting. Sci. Adv. 2021, 7, eabg3989 10.1126/sciadv.abg3989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Feng S.; Ma Y.; Wang S.; Gao S.; Huang Q.; Zhen H.; Yan D.; Ling Q.; Lin Z. Light/Force-Sensitive 0D Lead-Free Perovskites: From Highly Efficient Blue Afterglow to White Phosphorescence with Near-Unity Quantum Efficiency. Angew. Chem., Int. Ed. 2022, 61, e202116511 10.1002/anie.202116511. [DOI] [PubMed] [Google Scholar]
- Barkaoui H.; Abid H.; Yangui A.; Triki S.; Boukheddaden K.; Abid Y. Yellowish White-Light Emission Involving Resonant Energy Transfer in a New One-Dimensional Hybrid Material: (C9H10N2)PbCl4. J. Phys. Chem. C 2018, 122, 24253–24261. 10.1021/acs.jpcc.8b06850. [DOI] [Google Scholar]
- Roccanova R.; Houck M.; Yangui A.; Han D.; Shi H.; Wu Y.; Glatzhofer D. T.; Powell D. R.; Chen S.; Fourati H.; Lusson A.; Boukheddaden K.; Du M.-H.; Saparov B. Broadband Emission in Hybrid Organic–Inorganic Halides of Group 12 Metals. ACS Omega 2018, 3, 18791–18802. 10.1021/acsomega.8b02883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li K.-J.; Zhao Y.-Y.; Sun M.-E.; Chen G.-S.; Zhang C.; Liu H.-L.; Li H.-Y.; Zang S.-Q.; Mak T. C. W. Zero-Dimensional Zinc Halide Organic Hybrids with Excellent Optical Waveguide Properties. Cryst. Growth Des. 2022, 22, 3295–3302. 10.1021/acs.cgd.2c00117. [DOI] [Google Scholar]
- McWhorter T. M.; Zhang Z.; Creason T. D.; Thomas L.; Du M.-H.; Saparov B. (C7H11N2)2MBr4 (M=Cu, Zn): X-Ray Sensitive 0D Hybrid Metal Halides with Tunable Broadband Emission. Eur. J. Inorg. Chem. 2022, 2022, e202100954 10.1002/ejic.202100954. [DOI] [Google Scholar]
- Fattal H.; Creason T. D.; Delzer C. J.; Yangui A.; Hayward J. P.; Ross B. J.; Du M.-H.; Glatzhofer D. T.; Saparov B. Zero-Dimensional Hybrid Organic–Inorganic Indium Bromide with Blue Emission. Inorg. Chem. 2021, 60, 1045–1054. 10.1021/acs.inorgchem.0c03164. [DOI] [PubMed] [Google Scholar]
- Yangui A.; Roccanova R.; Wu Y.; Du M.-H.; Saparov B. Highly Efficient Broad-Band Luminescence Involving Organic and Inorganic Molecules in a Zero-Dimensional Hybrid Lead Chloride. J. Phys. Chem. C 2019, 123, 22470–22477. 10.1021/acs.jpcc.9b05509. [DOI] [Google Scholar]
- Zhang W.-F.; Pan W.-J.; Xu T.; Song R.-Y.; Zhao Y.-Y.; Yue C.-Y.; Lei X.-W. One-Dimensional Face-Shared Perovskites with Broad-Band Bluish White-Light Emissions. Inorg. Chem. 2020, 59, 14085–14092. 10.1021/acs.inorgchem.0c01861. [DOI] [PubMed] [Google Scholar]
- Dammak T.; Abid Y. Quasi-white light emission involving Förster resonance energy transfer in a new organic inorganic tin chloride based material (AMPS)[SnCl6]H2O. Opt. Mater. 2017, 66, 302–307. 10.1016/j.optmat.2017.02.020. [DOI] [Google Scholar]
- Sun X.-Y.; Yue M.; Jiang Y.-X.; Zhao C.-H.; Liao Y.-Y.; Lei X.-W.; Yue C.-Y. Combining Dual-Light Emissions to Achieve Efficient Broadband Yellowish-Green Luminescence in One-Dimensional Hybrid Lead Halides. Inorg. Chem. 2021, 60, 1491–1498. 10.1021/acs.inorgchem.0c02785. [DOI] [PubMed] [Google Scholar]
- Huang Q.; Yang S.; Feng S.; Zhen H.; Lin Z.; Ling Q. Multicolor Output from 2D Hybrid Perovskites with Wide Band Gap: Highly Efficient White Emission, Dual-Color Afterglow, and Switch between Fluorescence and Phosphorescence. J. Phys. Chem. Lett. 2021, 12, 1040–1045. 10.1021/acs.jpclett.0c03538. [DOI] [PubMed] [Google Scholar]
- Chaaban M.; Ben-Akacha A.; Worku M.; Lee S.; Neu J.; Lin X.; Vellore Winfred J. S. R.; Delzer C. J.; Hayward J. P.; Du M.-H.; Siegrist T.; Ma B. Metal Halide Scaffolded Assemblies of Organic Molecules with Enhanced Emission and Room Temperature Phosphorescence. J. Phys. Chem. Lett. 2021, 12, 8229–8236. 10.1021/acs.jpclett.1c02354. [DOI] [PubMed] [Google Scholar]
- Xu L.-J.; Plaviak A.; Lin X.; Worku M.; He Q.; Chaaban M.; Kim B. J.; Ma B. Metal Halide Regulated Photophysical Tuning of Zero-Dimensional Organic Metal Halide Hybrids: From Efficient Phosphorescence to Ultralong Afterglow. Angew. Chem., Int. Ed. 2020, 59, 23067–23071. 10.1002/anie.202010555. [DOI] [PubMed] [Google Scholar]
- Liu S.; Fang X.; Lu B.; Yan D. Wide range zero-thermal-quenching ultralong phosphorescence from zero-dimensional metal halide hybrids. Nat. Commun. 2020, 11, 4649 10.1038/s41467-020-18482-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee L. C.-C.; Lo K. K.-W. Luminescent and Photofunctional Transition Metal Complexes: From Molecular Design to Diagnostic and Therapeutic Applications. J. Am. Chem. Soc. 2022, 144, 14420–14440. 10.1021/jacs.2c03437. [DOI] [PubMed] [Google Scholar]
- Artem’ev A. V.; Pritchina E. A.; Rakhmanova M. I.; Gritsan N. P.; Bagryanskaya I. Y.; Malysheva S. F.; Belogorlova N. A. Alkyl-dependent self-assembly of the first red-emitting zwitterionic {Cu4I6} clusters from [alkyl-P(2-Py)3]+ salts and CuI: when size matters. Dalton Trans. 2019, 48, 2328–2337. 10.1039/C8DT04328K. [DOI] [PubMed] [Google Scholar]
- Zhu K.; Cheng Z.; Rangan S.; Cotlet M.; Du J.; Kasaei L.; Teat S. J.; Liu W.; Chen Y.; Feldman L. C.; O’Carroll D. M.; Li J. A New Type of Hybrid Copper Iodide as Nontoxic and Ultrastable LED Emissive Layer Material. ACS Energy Lett. 2021, 6, 2565–2574. 10.1021/acsenergylett.1c00993. [DOI] [Google Scholar]
- Bazany-Rodríguez I. J.; Salomón-Flores M. K.; Viviano-Posadas A. O.; García-Eleno M. A.; Barroso-Flores J.; Martínez-Otero D.; Dorazco-González A. Chemosensing of neurotransmitters with selectivity and naked eye detection of l-DOPA based on fluorescent Zn(ii)-terpyridine bearing boronic acid complexes. Dalton Trans. 2021, 50, 4255–4269. 10.1039/D0DT04228E. [DOI] [PubMed] [Google Scholar]
- Kozhevnikov D. N.; Shabunina O. V.; Kopchuk D. S.; Slepukhin P. A.; Kozhevnikov V. N. 5-Aryl-2,2′-bipyridines as tunable fluorophores. Tetrahedron Lett. 2006, 47, 7025–7029. 10.1016/j.tetlet.2006.07.111. [DOI] [Google Scholar]
- Ardizzoia G. A.; Brenna S.; Durini S.; Therrien B.; Veronelli M. Synthesis, Structure, and Photophysical Properties of Blue-Emitting Zinc(II) Complexes with 3-Aryl-Substituted 1-Pyridylimidazo[1,5-a]pyridine Ligands. Eur. J. Inorg. Chem. 2014, 2014, 4310–4319. 10.1002/ejic.201402415. [DOI] [Google Scholar]
- Wang X.-M.; Chen S.; Fan R.-Q.; Zhang F.-Q.; Yang Y.-L. Facile luminescent tuning of ZnII/HgII complexes based on flexible, semi-rigid and rigid polydentate Schiff bases from blue to green to red: structural, photophysics, electrochemistry and theoretical calculations studies. Dalton Trans. 2015, 44, 8107–8125. 10.1039/C5DT00057B. [DOI] [PubMed] [Google Scholar]
- Dong Y.; Fan R.; Chen W.; Wang P.; Yang Y. A simple quinolone Schiff-base containing CHEF based fluorescence ‘turn-on’ chemosensor for distinguishing Zn2+ and Hg2+ with high sensitivity, selectivity and reversibility. Dalton Trans. 2017, 46, 6769–6775. 10.1039/C7DT00956A. [DOI] [PubMed] [Google Scholar]
- Tsukamoto T.; Aoki R.; Sakamoto R.; Toyoda R.; Shimada M.; Hattori Y.; Kitagawa Y.; Nishibori E.; Nakano M.; Nishihara H. Mechano-, thermo-, solvato-, and vapochromism in bis(acetato-κ1O)[4′-(4-(diphenylamino)phenyl)](2,2′:6′,2″-terpyridine-κ3N,N′,N″)zinc(ii) and its polymer. Chem. Commun. 2017, 53, 9805–9808. 10.1039/C7CC05022D. [DOI] [PubMed] [Google Scholar]
- Hiscock L. K.; Joekar D.; Balonova B.; Tomas Piqueras M.; Schroeder Z. W.; Jarvis V.; Maly K. E.; Blight B. A.; Dawe L. N. Structures, Phase Behavior, and Fluorescent Properties of 3-Phenyl-1-(pyridin-2-yl)-1H-pyrazol-5-amine and Its ZnCl2 Complex. Inorg. Chem. 2019, 58, 16317–16321. 10.1021/acs.inorgchem.9b02765. [DOI] [PubMed] [Google Scholar]
- Temerova D.; Kisel K. S.; Eskelinen T.; Melnikov A. S.; Kinnunen N.; Hirva P.; Shakirova J. R.; Tunik S. P.; Grachova E. V.; Koshevoy I. O. Diversifying the luminescence of phenanthro-diimine ligands in zinc complexes. Inorg. Chem. Front. 2021, 8, 2549–2560. 10.1039/D1QI00149C. [DOI] [Google Scholar]
- Nugent J. W.; Lee H.; Lee H.-S.; Reibenspies J. H.; Hancock R. D. Mechanism of chelation enhanced fluorescence in complexes of cadmium(ii), and a possible new type of anion sensor. Chem. Commun. 2013, 49, 9749–9751. 10.1039/c3cc45829f. [DOI] [PubMed] [Google Scholar]
- Wang Z.; Zhu C.-Y.; Yin S.-Y.; Wei Z.-W.; Zhang J.-H.; Fan Y.-N.; Jiang J.-J.; Pan M.; Su C.-Y. A Metal–Organic Supramolecular Box as a Universal Reservoir of UV, WL, and NIR Light for Long-Persistent Luminescence. Angew. Chem., Int. Ed. 2019, 58, 3481–3485. 10.1002/anie.201812708. [DOI] [PubMed] [Google Scholar]
- Liang Q.-F.; Zheng H.-W.; Yang D.-D.; Zheng X.-J. A triphenylamine derivative and its Cd(ii) complex with high-contrast mechanochromic luminescence and vapochromism. CrystEngComm 2022, 24, 543–551. 10.1039/D1CE01319J. [DOI] [Google Scholar]
- Volpi G.; Priola E.; Garino C.; Daolio A.; Rabezzana R.; Benzi P.; Giordana A.; Diana E.; Gobetto R. Blue fluorescent zinc(II) complexes based on tunable imidazo[1,5-a]pyridines. Inorg. Chim. Acta 2020, 509, 119662 10.1016/j.ica.2020.119662. [DOI] [Google Scholar]
- Younes A. H.; Zhang L.; Clark R. J.; Zhu L. Fluorescence of 5-Arylvinyl-5′-Methyl-2,2′-Bipyridyl Ligands and Their Zinc Complexes. J. Org. Chem. 2009, 74, 8761–8772. 10.1021/jo901889y. [DOI] [PubMed] [Google Scholar]
- Divya K. P.; Sreejith S.; Balakrishna B.; Jayamurthy P.; Anees P.; Ajayaghosh A. A Zn2+-specific fluorescent molecular probe for the selective detection of endogenous cyanide in biorelevant samples. Chem. Commun. 2010, 46, 6069–6071. 10.1039/c0cc01159b. [DOI] [PubMed] [Google Scholar]
- Wang Z.; Zhu C.-Y.; Mo J.-T.; Xu X.-Y.; Ruan J.; Pan M.; Su C.-Y. Multi-Mode Color-Tunable Long Persistent Luminescence in Single-Component Coordination Polymers. Angew. Chem., Int. Ed. 2021, 60, 2526–2533. 10.1002/anie.202012831. [DOI] [PubMed] [Google Scholar]
- Kozhevnikov V. N.; Shabunina O. V.; Kopchuk D. S.; Ustinova M. M.; König B.; Kozhevnikov D. N. Facile synthesis of 6-aryl-3-pyridyl-1,2,4-triazines as a key step toward highly fluorescent 5-substituted bipyridines and their Zn(II) and Ru(II) complexes. Tetrahedron 2008, 64, 8963–8973. 10.1016/j.tet.2008.06.040. [DOI] [Google Scholar]
- Hao P.; Liu X.; Guo C.; Zhao G.; Li G.; Shen J.; Fu Y. Lattice solvent controlled photochromism of tripyridyl-triazine-based zinc bromide complexes. Inorg. Chem. Front. 2022, 9, 879–888. 10.1039/D1QI01242H. [DOI] [Google Scholar]
- Zhu H.-G.; Xu Y.; Yu Z.; Wu Q.-J.; Fun H.-K.; You X.-Z. The synthesis, structure of 1-D lead halide adducts: PbI2(L) (L = 2,2′-bipyridine, 1,10-phenanthroline). Polyhedron 1999, 18, 3491–3495. 10.1016/S0277-5387(99)00280-6. [DOI] [Google Scholar]
- Marandi F.; Garousi E.; Fun H.-K. Supramolecular organization in lead(II) halides of 4,4′-dimethoxy-2,2′-bipyridine ligand. J. Mol. Struct. 2013, 1049, 205–211. 10.1016/j.molstruc.2013.06.019. [DOI] [Google Scholar]
- Travis W.; Knapp C. E.; Savory C. N.; Ganose A. M.; Kafourou P.; Song X.; Sharif Z.; Cockcroft J. K.; Scanlon D. O.; Bronstein H.; Palgrave R. G. Hybrid Organic–Inorganic Coordination Complexes as Tunable Optical Response Materials. Inorg. Chem. 2016, 55, 3393–3400. 10.1021/acs.inorgchem.5b02749. [DOI] [PubMed] [Google Scholar]
- Thompson J. R.; Snider D.; Wren J. E. C.; Kroeker S.; Williams V. E.; Leznoff D. B. Correlating Structural Features and 207Pb NMR Parameters with the Stereochemical Activity of PbII Lone Pairs in Birefringent Pb[2,6-bis(benzimidazol-2-yl)pyridine] Complexes. Eur. J. Inorg. Chem. 2017, 2017, 88–98. 10.1002/ejic.201601235. [DOI] [Google Scholar]
- Yang Y.; Wu Y.; Mei D.; Qu Y. Lead-based crystal luminophores: Tuning emissive crystals and application for recording high temperature history with ratiometric and colorimetric signals. Chem. Eng. J. 2019, 358, 606–613. 10.1016/j.cej.2018.10.028. [DOI] [Google Scholar]
- Qiu K.; Ouyang M.; Liu Y.; Huang H.; Liu C.; Chen Y.; Ji L.; Chao H. Two-photon photodynamic ablation of tumor cells by mitochondria-targeted iridium(III) complexes in aggregate states. J. Mater. Chem. B 2017, 5, 5488–5498. 10.1039/C7TB00731K. [DOI] [PubMed] [Google Scholar]
- Kisel K. S.; Eskelinen T.; Zafar W.; Solomatina A. I.; Hirva P.; Grachova E. V.; Tunik S. P.; Koshevoy I. O. Chromophore-functionalized phenanthro-diimine ligands and their Re(I) complexes. Inorg. Chem. 2018, 57, 6349–6361. 10.1021/acs.inorgchem.8b00422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marcoux D.; Charette A. B. Palladium-Catalyzed Synthesis of Functionalized Tetraarylphosphonium Salts. J. Org. Chem. 2008, 73, 590–593. 10.1021/jo702355c. [DOI] [PubMed] [Google Scholar]
- APEX2 - Software Suite for Crystallographic Programs; Bruker AXS, Inc: Madison, WI, USA, 2010. [Google Scholar]
- Sheldrick G. M.SADABS-2008/1 - Bruker AXS Area Detector Scaling and Absorption Correction; Bruker AXS: Madison, Wisconsin, USA, 2008. [Google Scholar]
- Sheldrick G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3–8. 10.1107/S2053229614024218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Farrugia L. J. WinGX and ORTEP for Windows: an Update. J. Appl. Crystallogr. 2012, 45, 849–854. 10.1107/S0021889812029111. [DOI] [Google Scholar]
- Spek A. L. PLATON SQUEEZE: a tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 9–18. 10.1107/S2053229614024929. [DOI] [PubMed] [Google Scholar]
- Spek A. L.PLATON, A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2001. [Google Scholar]
- Perdew J. P.; Burke K.; Ernzerhof M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
- Adamo C.; Barone V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. 10.1063/1.478522. [DOI] [Google Scholar]
- Weigend F.; Ahlrichs R. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. 10.1039/b508541a. [DOI] [PubMed] [Google Scholar]
- Andrae D.; Häußermann U.; Dolg M.; Stoll H.; Preuß H. Energy-Adjusted Ab Initio Pseudopotentials for the Second and Third Row Transition Elements. Theor. Chem. Acta. 1990, 77, 123–141. 10.1007/BF01114537. [DOI] [Google Scholar]
- Dapprich S.; Komáromi I.; Byun K. S.; Morokuma K.; Frisch M. J. A new ONIOM implementation in Gaussian98. Part I. The calculation of energies, gradients, vibrational frequencies and electric field derivatives. J. Mol. Struct.: THEOCHEM 1999, 461-462, 1–21. 10.1016/S0166-1280(98)00475-8. [DOI] [Google Scholar]
- Rappe A. K.; Casewit C. J.; Colwell K. S.; Goddard W. A.; Skiff W. M. UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024–10035. 10.1021/ja00051a040. [DOI] [Google Scholar]
- Rappe A. K.; Goddard W. A. Charge equilibration for molecular dynamics simulations. J. Phys. Chem. A 1991, 95, 3358–3363. 10.1021/j100161a070. [DOI] [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian 16, Rev. C.01, Wallingford, CT; 2016.
- Neese F. Software update: The ORCA program system-Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606 10.1002/wcms.1606. [DOI] [Google Scholar]
- Barone V.; Cossi M. Quantum Calculation of Molecular Energies and Energy Gradients in Solution by a Conductor Solvent Model. J. Phys. Chem. A 1998, 102, 1995–2001. 10.1021/jp9716997. [DOI] [Google Scholar]
- Rohrdanz M. A.; Martins K. M.; Herbert J. M. A long-range-corrected density functional that performs well for both ground-state properties and time-dependent density functional theory excitation energies, including charge-transfer excited states. J. Chem. Phys. 2009, 130, 054112 10.1063/1.3073302. [DOI] [PubMed] [Google Scholar]
- Neese F. An improvement of the resolution of the identity approximation for the formation of the Coulomb matrix. J. Comput. Chem. 2003, 24, 1740–1747. 10.1002/jcc.10318. [DOI] [PubMed] [Google Scholar]
- Neese F.; Wennmohs F.; Hansen A.; Becker U. Efficient, approximate and parallel Hartree–Fock and hybrid DFT calculations. A ‘chain-of-spheres’ algorithm for the Hartree–Fock exchange. Chem. Phys. 2009, 356, 98–109. 10.1016/j.chemphys.2008.10.036. [DOI] [Google Scholar]
- Chen S.; Fan R.-Q.; Wang X.-M.; Yang Y.-L. Novel bright blue emissions IIB group complexes constructed with various polyhedron-induced 2-[2′-(6-methoxy-pyridyl)]-benzimidazole derivatives. CrystEngComm 2014, 16, 6114–6125. 10.1039/c4ce00382a. [DOI] [Google Scholar]
- Song Y.; Kim D.; Lee H.-J.; Lee H. Cd(II) and Zn(II) Complexes Containing N,N’-Bidentate N-(Pyridin-2-ylmethylene)cyclopentanamine: Synthesis, Characterisation and Methyl Methacrylate Polymerisation. Bull. Korean Chem. Soc. 2014, 35, 2929–2934. 10.5012/bkcs.2014.35.10.2929. [DOI] [Google Scholar]
- Kokina T. E.; Glinskaya L. A.; Tkachev A. V.; Plyusnin V. F.; Tsoy Y. V.; Bagryanskaya I. Y.; Vasilyev E. S.; Piryazev D. A.; Sheludyakova L. A.; Larionov S. V. Chiral zinc(II) and cadmium(II) complexes with a dihydrophenanthroline ligand bearing (−)-α-pinene fragments: Synthesis, crystal structures and photophysical properties. Polyhedron 2016, 117, 437–444. 10.1016/j.poly.2016.06.018. [DOI] [Google Scholar]
- Munshi S. J.; Saini J. K.; Ingle S.; Kumar S. B. Metal(II) chloride complexes containing a tridentate N-donor Schiff base ligand: syntheses, structures and antimicrobial activity. J. Coord. Chem. 2021, 74, 2004–2016. 10.1080/00958972.2021.1940157. [DOI] [Google Scholar]
- Klapötke T. M.; Krumm B.; Polbom K.; Rienäcker C. M. Synthesis and Characterization of Bromo- and Bromochloroplumbates(II). Crystal Structures of [Ph4E]2[Pb3Br8] (E = P, As) and [Ph4P][PbBrCl2] • CH3CN. Z. Naturforsch. B 2000, 55, 377–382. 10.1515/znb-2000-0506. [DOI] [Google Scholar]
- Gröger H.; Lode C.; Vollmer H.; Krautscheid H.; Lebedkin S. Bromoplumbate mit kettenförmigen und isolierten Anionen: (Bzl4P)2[Pb3Br8], (Bzl4P)2[Pb3Br8(dmf)2], (Bzl4P)[PbBr3], (Bzl4P)2[PbBr4] und (Bzl4P)4[Pb2Br6][PbBr4]. Z. Anorg. Allg. Chem. 2002, 628, 57–62. . [DOI] [Google Scholar]
- Morad V.; Shynkarenko Y.; Yakunin S.; Brumberg A.; Schaller R. D.; Kovalenko M. V. Disphenoidal Zero-Dimensional Lead, Tin, and Germanium Halides: Highly Emissive Singlet and Triplet Self-Trapped Excitons and X-ray Scintillation. J. Am. Chem. Soc. 2019, 141, 9764–9768. 10.1021/jacs.9b02365. [DOI] [PubMed] [Google Scholar]
- Lin H.; Zhou C.; Chaaban M.; Xu L.-J.; Zhou Y.; Neu J.; Worku M.; Berkwits E.; He Q.; Lee S.; Lin X.; Siegrist T.; Du M.-H.; Ma B. Bulk Assembly of Zero-Dimensional Organic Lead Bromide Hybrid with Efficient Blue Emission. ACS Mater. Lett. 2019, 1, 594–598. 10.1021/acsmaterialslett.9b00333. [DOI] [Google Scholar]
- Zhou J.; Li M.; Ning L.; Zhang R.; Molokeev M. S.; Zhao J.; Yang S.; Han K.; Xia Z. Broad-Band Emission in a Zero-Dimensional Hybrid Organic [PbBr6] Trimer with Intrinsic Vacancies. J. Phys. Chem. Lett. 2019, 10, 1337–1341. 10.1021/acs.jpclett.9b00238. [DOI] [PubMed] [Google Scholar]
- Febriansyah B.; Neo C. S. D.; Giovanni D.; Srivastava S.; Lekina Y.; Koh T. M.; Li Y.; Shen Z. X.; Asta M.; Sum T. C.; Mathews N.; England J. Targeted Synthesis of Trimeric Organic–Bromoplumbate Hybrids That Display Intrinsic, Highly Stokes-Shifted, Broadband Emission. Chem. Mater. 2020, 32, 4431–4441. 10.1021/acs.chemmater.9b03925. [DOI] [Google Scholar]
- Reger D. L.; Huff M. F.; Rheingold A. L.; Haggerty B. S. Control of structure in lead(II) complexes using poly(pyrazolyl)borate ligands. Stereochemically inactive lone pair in octahedral [HB(3,5-Me2pz)3]2Pb (pz = pyrazolyl ring). J. Am. Chem. Soc. 1992, 114, 579–584. 10.1021/ja00028a026. [DOI] [Google Scholar]
- Zhou C.; Lin H.; Neu J.; Zhou Y.; Chaaban M.; Lee S.; Worku M.; Chen B.; Clark R.; Cheng W.; Guan J.; Djurovich P.; Zhang D.; Lü X.; Bullock J.; Pak C.; Shatruk M.; Du M.-H.; Siegrist T.; Ma B. Green Emitting Single-Crystalline Bulk Assembly of Metal Halide Clusters with Near-Unity Photoluminescence Quantum Efficiency. ACS Energy Lett. 2019, 4, 1579–1583. 10.1021/acsenergylett.9b00991. [DOI] [Google Scholar]
- Toma O.; Mercier N.; Allain M.; Meinardi F.; Botta C. Lead(II) 4,4′-Bipyridine N-Oxide Coordination Polymers – Highly Phosphorescent Materials with Mechanochromic Luminescence Properties. Eur. J. Inorg. Chem. 2017, 2017, 844–850. 10.1002/ejic.201601230. [DOI] [Google Scholar]
- Martínez-Casado F. J.; Ramos-Riesco M.; Rodríguez-Cheda J. A.; Cucinotta F.; Fernández-Martínez A.; Garrido L.; Matesanz E.; Marchese L. Short lead(ii) soaps: from weakly fluorescent crystals to strongly phosphorescent and structurally varied vitreous phases. A thermal, structural and spectroscopic study. J. Mater. Chem. C 2014, 2, 9489–9496. 10.1039/C4TC01645A. [DOI] [Google Scholar]
- Liu X.; Zhai L.; Zhang W.-W.; Zuo J.-L.; Yang Z.-X.; Ren X.-M. Intense greenish phosphorescence emission under ambient conditions in a two-dimensional lead(ii) coordination polymer with a 1,1′-ethynebenzene-3,3′,5,5′-tetracarboxylate ligand. Dalton Trans. 2017, 46, 7953–7959. 10.1039/C7DT01626C. [DOI] [PubMed] [Google Scholar]
- Wu X.-S.; Wang Y.-X.; Li S.-Q.; Qian Y.; Zhai L.; Wang X.-Z.; Ren X.-M. Metal ion coordination enhancing quantum efficiency of ligand phosphorescence in a double-stranded helical chain coordination polymer of Pb2+ with nicotinic acid. Dalton Trans. 2018, 47, 14636–14643. 10.1039/C8DT03589J. [DOI] [PubMed] [Google Scholar]
- Gao X.-S.; Dai H.-J.; Ding M.-J.; Pei W.-B.; Ren X.-M. Stereochemically Active and Inactive Lone Pairs in Two Room-Temperature Phosphorescence Coordination Polymers of Pb2+ with Different Tricarboxylic Acids. Inorg. Chem. 2019, 58, 6772–6780. 10.1021/acs.inorgchem.9b00215. [DOI] [PubMed] [Google Scholar]
- Zhu L.-L.; Huang Y.-E.; Gong L.-K.; Huang X.-Y.; Qi X.-H.; Wu X.-H.; Du K.-Z. Ligand Control of Room-Temperature Phosphorescence Violating Kasha’s Rule in Hybrid Organic–Inorganic Metal Halides. Chem. Mater. 2020, 32, 1454–1460. 10.1021/acs.chemmater.9b04268. [DOI] [Google Scholar]
- Hu H.; Meier F.; Zhao D.; Abe Y.; Gao Y.; Chen B.; Salim T.; Chia E. E. M.; Qiao X.; Deibel C.; Lam Y. M. Efficient Room-Temperature Phosphorescence from Organic–Inorganic Hybrid Perovskites by Molecular Engineering. Adv. Mater. 2018, 30, 1707621 10.1002/adma.201707621. [DOI] [PubMed] [Google Scholar]
- Yang S.; Wu D.; Gong W.; Huang Q.; Zhen H.; Ling Q.; Lin Z. Highly efficient room-temperature phosphorescence and afterglow luminescence from common organic fluorophores in 2D hybrid perovskites. Chem. Sci. 2018, 9, 8975–8981. 10.1039/C8SC03563F. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gong L.-K.; Li J.-R.; Wu Z.-F.; Hu B.; Wang Z.-P.; Shen N.-N.; Hu Q.-Q.; Deng Z.-H.; Zhang Z.-Z.; Fu J.-J.; Du K.-Z.; Huang X.-Y. Enhancing the phosphorescence of hybrid metal halides through molecular sensitization. J. Mater. Chem. C 2019, 7, 9803–9807. 10.1039/C9TC03067K. [DOI] [Google Scholar]
- Strasser A.; Vogler A. Intraligand phosphorescence of lead(II) β-diketonates under ambient conditions. J. Photochem. Photobiol. A: Chem. 2004, 165, 115–118. 10.1016/j.jphotochem.2004.03.007. [DOI] [Google Scholar]
- Strasser A.; Vogler A. Optical properties of thallium(I), lead(II) and bismuth(III) hexafluoroacetylacetonates. Intraligand phosphorescence under ambient conditions. Inorg. Chem. Commun. 2004, 7, 528–530. 10.1016/j.inoche.2003.12.039. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.





