Next Article in Journal
How Do the Surroundings of the C-NO2 Fragment Affect the Mechanical Sensitivity of Trinitroaromatic Molecules? Evidence from Crystal Structures and Ab Initio Calculations
Previous Article in Journal
Gone with the Wind—Adducts of Volatile Pyridine Derivatives and Copper(II) Acetylacetonate
Previous Article in Special Issue
Influence of Energetic Xe132 Ion Irradiation on Optical, Luminescent and Structural Properties of Ce-Doped Y3Al5O12 Single Crystals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Design and Synthesis of a New Photoluminescent 2D Coordination Polymer Employing a Ligand Derived from Quinoline and Pyridine

by
Andrzej Kochel
1,*,
Małgorzata Hołyńska
2,*,
Aneta Jezierska
1 and
Jarosław J. Panek
1
1
Faculty of Chemistry, University of Wrocław, ul. F. Joliot Curie 14, 50 383 Wrocław, Poland
2
ESA-ESTEC, Keplerlaan 1, 2200 AZ Noordwijk, The Netherlands
*
Authors to whom correspondence should be addressed.
Crystals 2025, 15(8), 691; https://doi.org/10.3390/cryst15080691
Submission received: 5 June 2025 / Revised: 23 July 2025 / Accepted: 24 July 2025 / Published: 30 July 2025
(This article belongs to the Special Issue Research Progress of Photoluminescent Materials)

Abstract

Application of organic ligand 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate with N/O donor atoms enabled solvothermal synthesis of a 2D Cu(II) coordination polymer, {Cu(L)BF4}n (L = deprotonated 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate). Both the ligand and its coordination polymer have been characterized. The condensed ring system of the applied ligand promotes the formation of coordination polymers rather than mononuclear species. The obtained 2D coordination polymer is photoluminescent with bathochromic/hypsochromic shifts in ligand absorption bands leading to a single absorption band at 465 nm. Density Functional Theory was employed to provide a theoretical description of the possible conformational changes within the ligand, with emphasis on the difference between the ligand conformation in its hydrochloride salt and in the polymer. Two models of polymer fragments were constructed to describe the electronic structure and non-covalent interactions. The Quantum Theory of Atoms in Molecules (QTAIM) was applied for this purpose. Using the obtained results, we were able to develop potential energy profiles for various conformations of the ligand. For the set of the studied systems, we detected non-covalent interactions, which are responsible for the spatial conformation. Concerning the models of polymers, electron spin density distribution has been visualized and discussed.

1. Introduction

Application of aminocarboxylate ligands with N/O-donor atoms and aromatic rings provides extensive possibilities for construction of 1-3D coordination polymers. The resulting products may display useful photochemical as well as magnetic properties. This is dependent on the used metal ions [1,2,3,4,5,6,7,8,9,10,11,12,13,14]. Examples of Cu(II) coordination polymers have been reported with applications for gas absorption or ion exchange, for instance (the 4,4′–bpy 4,4′-bipyridine ligand has been utilized in the synthesis of [Cu(SiF6)(4,4′–bpy)2]n [1], [{Cu(GeF6)(4,4′–bpy)2}·8H2O]n [5], [{Cu(BF4)2(4,4′–bpy)(H2O)2}·4,4′–bpy]n [6], (pzdc = pyrazine-2,3-dicarboxylate; dpyg = 1,2-Di(4-pyridyl)glycol) [Cu2(pzdc)2(dpyg)] [7], (tctfm-4,4′,4″,4‴tetracyanotetraphenylmethane) [{Cu(tctfm)}∙BF4∙xC6H5NO2] [8], {[Cu(4,4′-bpy)1.5]·NO3·1.5H2O}n [9], and (rot-N,N′-bis(3-pyridylmethyl)-1,5-diaminopentane dihydronitrate){[Cu(rot)(C2O4)0.5(H2O)23NO3·20H2O}n [10].
Coordination polymers are also applied in heterogeneous catalysis. Compounds (where used 4,4′-bpy 4,4′-bipyridine, carboxylato co-ligand (formato or acetato){[Cu3(4,4′–bpy)3(μ–OOCH)4(H2O)2](ClO4)2(H2O)6}n, {[Cu(4,4′–bpy)(μ–OOCH)(NO3)]}n, {[Cu2(4,4′–bpy)2(μ–OOCCH3)3](PF6)(H2O)}n, being 2D coordination polymers, are effective catalysts to cyanosililation of aldehydes, which is a convenient synthesis method for cyanohydrins [11]. On the other hand, a compound with a ligand (bznbz, 1,3–bis–(5,6–dimetylobenzimidazol–1–il–metyleno)benzene){[Cu(bzmbz)2]·(NO3)2}n catalyzes degradation of azo dyes in the presence of hydroxyl radicals OH·, obtained from H2O2 solution containing Fe2+ ions under acidic conditions [12].
Another important field is luminescence of coordination polymers. In comparison to organic compounds used, among others, in the production of OLEDs, inorganic coordination compounds often display higher thermal stability, which allows for a wider application range. Depending on the used central metal ion and, in particular, its valence electrons configuration, luminescence properties of coordination polymers may be governed by excited states of inter-ligand charge transfer (IL), metal–ligand charge transfer (MLCT), ligand–metal charge transfer (LMCT), and ligand–ligand charge transfer (LLCT) character. One of the examples uses pyram—2-pirydyno-4-pirydylometyloamine {[Cu(PPh3)(pyram)1.5]∙0.5CHCl3∙ClO4 compound in which, in comparison to the free ligand (λem = 460 nm), the absorption band is shifted toward lower energies with 10 times higher intensity [13].
There is also much research on the electrical properties of coordination polymers. The reference materials are conductive metals with conductivity of 104–105 S cm–1, normally increasing with a decrease in temperature. An important and interesting example is the Cu(I) coordination polymer with ligand dcnqi—diiminum N,N′–dicyano–1,4–benzoquinone [Cu(dcnqi)2]n with a structure comprising seven superposed diamond-type lattices [14]. At 295 K, monocrystals of this compound display electrical conductivity values from 800 to 1000 S cm–1.
Herein are new insights into the coordination chemistry of a new ligand based on quinoline, functionalized with carboxylate and pyridyl groups, employed as a precursor of a novel two-dimensional coordination polymer. Introduction of a substituted pyrazine in this ligand results in increased coordination capabilities. The resulting Cu(II) polymer displays photoluminescent and antiferromagnetic properties. The experimental findings are supported by quantum-chemical simulations with the aid of DFT (Density Functional Theory) [15,16]. Based on theoretical investigations, it was possible to analyze conformational changes of the ligand as well as electronic density distribution. The network of non-covalent interactions was covered by means of Quantum Theory of Atoms in Molecules (QTAIM) [17,18].

2. Materials and Methods

2.1. Synthesis of Crystalline Materials

2-(3-Ethyl-pyrazin-2-yl)quinoline-4-carboxylate acid [LH] 1

In this study, 35 ml of double-distilled water and 10 ml of ethanol were combined in a flask (50 ml) equipped with a reflux condenser. To this mixture, isatin (1.22 g, 8.30 mmol; Sigma-Aldrich, Poznan, Poland), 2-acetyl-3-ethyl-pyrazine (0.80 g, 5.32 mmol; Aldrich), and KOH (2.24 g, 40.00 mmol; POCH, Gliwice, Poland) were added (Scheme 1). The mixture was stirred magnetically for a couple of minutes. It was then heated at 100 °C for 11 h. After cooling, 2 M HCl was added to adjust the pH to 4–5. A dark-yellow precipitate formed and was rinsed with distilled water. This procedure was continued until the precipitate’s color changed to light yellow. A powder was obtained as a result of drying this precipitate in a desiccator. This powder could not be crystallized from water. It was possible to dissolve it in 2 M HCl. This action yielded a crystalline hydrochloride salt.
Yield (pure ligand): 70% (based on isatin).
Elemental analysis calculated (%) for C16 H13 N3 O2
C 69.35 H 4.88, N 14.62; found: C 69.10, H 4.20, N 14.25.
IR (Nujol, cm−1): 3401 s νsOH; 1621 s νsCOO; 1600 s ν(ring); 1584 s νasCOO; 1551 m, 1515 s δCNC; 1403 s νCC; 1344 m νsCOO; 1303 w, 1244 w νs COO/νCC; 1153 m νsCOO/δCOH; 1110 m νsOH; 1059 w νsCOO; 870 w νCN; 849 w νCN; 819 w νCN; 790 m δOCO; 765 m νCN/δOCO; 724 w δOCO; 680 w δ(ring); 663 m δCH; 642 m δCH; 591 m δCH; 520 w νCC; 492 w, 468 m δCNC; 434 w δCNC; 414 w δCNC; 278 w, 254 w, 247 w δCH; 227 w δC=C(ring); 203 w, 177 w, 150 w, 140 w, 133 w, 127 w, 121 w, 106 w, 100 w, 81 vw, 74 vw, 56 vw νasCOO.
1H NMR (500.13 MHz, methanol-d4, 300.0 K): δ = 1.43 (t, 3H, J = 7.5 Hz), 3.29–3.34 (m, 2H), 7.89–7.94 (m, 1H), 8.02–8.07 (m, 1H), 8.32 (d, 1H, J = 8.5 Hz), 8.73 (s, 1H), 8.80 (d, 1H, J = 1.7 Hz), 8.87 (s, 1H), 8.99 (d, 1H, J = 8.6 Hz) ppm.
13C NMR (125.77 MHz, methanol-d4, 300.0 K): δ = 13.6, 29.1, 124.2, 126.6, 127.1, 127.5, 128.0, 131.2, 133.5, 141.7, 143.9, 144.0, 146.6, 150.2, 158.5, 167.9 ppm.
Synthesis of {Cu(L)BF4}n (L = (L = deprotonated 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate) 2.
Double-distilled water was applied, and Cu(BF4)2⋅H2O and 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate chloride dihydrate (organic ligand) were used for the synthesis (Scheme 2).
Teflon-lined pressure reactor Berghof 100 was fed with 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate chloride dihydrate (0.350 g, 1.17 mmol), 10 cm3 double distilled water, and Cu(BF4)2 • xH2O (anhydrous basis - 0.20 g, 0.843 mmol anhydrous basis).
The reactor was closed and flushed with nitrogen gas for 20 minutes and subsequently heated. During the first day, heating was carried out at 100 °C. For the next 4 days the reactor temperature was at 145 °C. Then, the reactor was turned off and slowly cooled down at 5 °C per hour. The coordination polymer product formed green–blue crystals at 73% yield. When other Cu(II) sources were used, no crystalline products were obtained. The reagent quantities have been optimized for the highest yields.
IR [KBr, cm−1]: 3431 s ν(CH), 2262 m ν(CH), 2011 w ν(OH), 1618 m ν(C=O), 1559 m νas(COO), 1517 m δ(CNC), 1407 m ν(CC)/ν(CN), 1380 m νs(COO), 783 m ν(CC), 650 m δ(CH)/ν(COO), 630 m δ(CH), 533 w ν(CC), 481 m ν(CC), 470 s νs(CuN)/νs(CuO).
Elemental analysis calculated (%) for {Cu(L)BF4}n: C 44.83 H 3.02, N 9.79; found: C 44.20, H 2.98, N 9.10.

2.2. IR Spectroscopy

IR spectra were recorded with a Bruker VERTEX 70 FTIR spectrometer (Bruker, Billerica, MA, USA) for the starting organic ligand and its Cu(II) polymer.

2.3. X-Ray Diffraction Studies

2.3.1. X-Ray Diffraction Experiment

Diffraction data were collected on a Rigaku Oxford Diffraction XtaLAB Synergy-R DW diffractometer equipped with a HyPix ARC 150° Hybrid Photon Counting (HPC) detector using CuKα (λ = 1.5418 Å) at 100 K. Data collection, cell refinement, data reduction, and analysis were carried out with the CRYSALISPRO (Rigaku Oxford Diffraction Ltd., Abingdon, UK, 2020). An analytical absorption correction was applied with the use of CRYSALISPRO RED [19].
CCDC numbers of compounds 1 (2443543 ligand) and 2 (2443544 2D polymer) contain the supplementary crystallographic data for this paper. These data are provided free of charge by The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif.
Selected X-ray data are shown in Tables S1–S11.

2.3.2. Crystal Structure Refinement

Both crystal structures were solved with Intrinsic Phasing using the SHELXT program and refined using SHELXL (2015 release) [20,21] with anisotropic thermal parameters for non-H atoms. In the final refinement cycles, all H atoms were treated as riding atoms in geometrically optimized positions.

2.4. Photoluminescence Spectra

The photoluminescence spectra of the applied organic ligand and for the title polymer for solid crystalline samples anchored to quartz tubes were recorded with an FLSP920 Spectrometer from Edinburgh Instruments, equipped with a 450 W Xenon arc lamp as an excitation source.

2.5. Studies of the Magnetic Properties

Magnetic susceptibility data were recorded on a Quantum Design MPMS-XL5 SQUID magnetometer over the 300–1.8 K temperature range. Magnetic data were corrected for diamagnetic contributions, which were estimated from Pascal’s constants, and for temperature-independent paramagnetism, estimated at 60 × 10−6 emu/mol for the Cu2+ ion.

2.6. NMR

NMR spectra were recorded on a Bruker 500 MHz device. For results, see Figure S1a,b.

2.7. Computational Methodology

The starting geometry of the analyzed ligand was extracted from the crystal structure (see text above) and modified accordingly to prepare suitable structures for further modeling. The same procedure was applied to prepare small and larger models of the polymer. The quantum-chemical simulations were performed based on Density Functional Theory (DFT) [15,16]. The simulations were carried out using M06 functional [22] and the def2-TZVP basis set for the ligand and the small model of the polymer in order to optimize the geometry of the systems studied [23,24]. Concerning the larger model of the polymer, the same functional but SDD basis set was used [25,26,27]. Harmonic vibrational calculations were performed as well to confirm that the obtained structures correspond with the minima on the Potential Energy Surface (PES). Next, the conformational analysis was performed for the ligand. Three torsional angles were scanned, and potential energy profiles were obtained. The method called “scan with optimization” was applied for this purpose. The procedure is based on geometry optimization of the remaining parts of the molecules while the chosen dihedral angle is scanned (in our case, the increment was set to 10°). The wavefunctions for further electron density studies were generated on the basis of the above-mentioned levels of theory. For this part of the simulations, the Gaussian 16, Rev. C.01 suite of programs was used [28]. The electronic structure analysis was carried out on the basis of the Quantum Theory of Atoms in Molecules (QTAIM) [15,16]. The presence of bond and ring critical points (BCPs and RCPs) allowed the detection of non-covalent interactions responsible for the molecular spatial constitution. It provided a qualitative picture of the electron density partitioning. The AIMAll program was employed for the QTAIM analysis, topological analysis, and results presentation [29]. The figures showing the obtained data were prepared using the VMD 1.9.3 [30] and GaussView [31] programs.

3. Results and Discussion

3.1. Crystal Structure

In this study, 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate was used for the synthesis. This ligand was isolated as hydrated chloride (Scheme 3), able to form coordination polymers with its N/O donor atoms and additionally displaying photoluminescent properties. The ligand obtained in Pfitzinger synthesis is not crystalline; therefore, for its structure determination, it was crystallized from aqueous solution in the form of chloride salt.
The resulting crystal structure 1 comprises protonated ligand cations, chloride counterions, and water molecules (Figure 1, Supplementary Tables S1–S6). The ligand protonated atom is the pyridyl N atom, whereas the quinoline N atom remains not protonated. The carboxylic group is twisted by 38.93(3)° (O2 C14 C1 C2 torsion angle). The carboxylic H atom is a donor to a hydrogen bond to chloride ion O1–H1B···Cl1 [1 − x, 1 − y, 1 − z] (Table S10). On the other hand, for the 2-ethylpyrazine moiety bonded to the quinoline ring via the C10 atom, the N1-C8-C10-C13 and C9-C8-C10-N2 angle values are 16.14(2) and 13.13(3)°, respectively. The orientation of quinoline and pyrazine rings is affected by the pyrazin ethyl group, constituting some kind of steric hindrance. The angle between quinoline and pyrazine ring planes is 12.08(3)°. The presence of water molecules and chloride ions results in the formation of a 3D hydrogen bonding network.
Figure 2 displays graph set motifs that can be distinguished in the crystal structure 1.
The Cu(II) coordination polymer was obtained in solvothermal synthesis in a Berghoff reactor in accordance with the scheme below.
A 2D Cu(II) coordination polymer with a single Cu2+ linked with donor atoms N/O of aminocarboxylate ligand (Figure 3) was obtained via a solvothermal route. This compound is displayed in Scheme 4 with the formula {Cu(L) BF4}n.
Each Cu2+ ion is tetracoordinate with three N atoms and one carboxylate O atom in its coordination sphere (the relevant bond lengths are Cu(1)−N(3) 1.9226(19); Cu(1)−O(1)−x,−1/2+y,1/2−z 2.4470(19); Cu(1)−N(1)x,1/2−y,−1/2+z 2.0375(19); and Cu(1)−N(2)x,1/2−y,−1/2+z 2.0309(19) Å, Table S8). The Cu2+ lies in a special position on a symmetry center so that only one coordinated N and O atom are symmetrically independent. The coordination sphere is in the shape of a trigonal pyramid. Additionally, in the crystal structure, C–H···O and C–H···F hydrogen bond types can be found (Table S11).
The polymer forms a 2D network along the [100] and [010] directions (Figure 4).
Within this network, weak stacking interactions between aromatic rings are observed with a centroid-to-centroid distance of 3.349(3) Å, as illustrated in Figure 5.
Furthermore, characteristic tubular arrangements are formed in the crystal structure (Figure 6a,b).
Figure 7 further displays the formation of channels along [100], whereas Figure 8 visualizes a layered arrangement of molecules along [001].

3.2. Photoluminescence Properties

Luminescence properties of the ligand and Cu(II) CPs were tested in solid state at room temperature. In the ligand spectrum (Figure 9), emission bands observed at 425 nm can be interpreted in terms of π π* transitions. For the coordination polymer (Figure 10), the absorption band is shifted to 465 nm and could be interpreted as a ligand–metal charge transfer (LMCT) band with delocalization of electron density from the ligand via its N atoms to Cu(II) d orbitals. Similar effects have been observed for copper(II) triazolate coordination polymer with cyanide as a co-ligand, reported by Zhu et al. [32]. The photoluminescence spectrum of the Cu(II) Schiff base complexes was described by Rani et al., who also noted a shift in polymer spectrum toward higher wavelengths [33]. For binuclear copper complexes with carboxylate ligands, photoluminescence properties have also been reported. Carboxylate ligands are donors of electron pairs, which lead to the auxochromic effect, increasing the intensity of light absorption [34]. Zink et al. reported on dinuclear copper(I) halide complexes with P-N ligands and proposed to apply these compounds for singlet harvesting in OLEDs [35].

3.3. Magnetic Properties

Figure 11 shows selected Cu···Cu distances whereas Figure 12 denotes model of antiferromagnetic interactions transmitted via ligand.
Magnetic susceptibility values for the coordination polymer were determined with a SQUID magnetometer at a 1.8–300 K temperature range. Based on the χmT experimental curve, at 300 K the corresponding value is 0.425 cm3 K mol−1 (Figure 13), pointing out that there are weak antiferromagnetic interactions between Cu2+ ions, transmitted via bonds within the ligand. In order to quantify these interactions based on spin Hamiltonian H = –2JS1S2, the Bleaney–Bowers Equation (1) was used [36], taking into account mainly the nearest-neighbor interactions:
χ m = N β 2 g 2 3 k T 1 + 1 3 exp 2 J k T 1
Knowing the two-dimensional crystal structure of 1, a molecular field model correction (2) was introduced [37]:
χ m corr = χ m 1 2 z J χ m N β 2 g 2
In the above equation, χm is defined as molar susceptibility, zJ’ characterizes magnetic interactions between Cu(II) ions in the crystal lattice, N is the Avogadro number, g = the spectroscopic splitting factor, β = the Bohr magneton, and k = the Boltzmann constant. On the χmT curve, the corresponding values decrease under 50 K, which may be caused by additional weak antiferromagnetic interactions between Cu(II) ions, supported by Neel temperature at 2.5 K. The following values resulted from application of the Bleaney–Bowers model: 2J = −1.07 ± 0.02 cm–1, zJ′ = –0.2 ± 0.1cm–1, and R = 3.66 10−3 (R = ∑[(χmT)obsd − (χmT)cald]2 ∑(χmT)obsd]2). Figure 12 illustrates a model of antiferromagnetic interactions between Cu2+ ions.

3.4. Molecular Modeling

Density Functional Theory (DFT) was employed for the theoretical description of the discussed ligand and two polymer models. The first relevant question is the conformational preference of the ligand, since the crystal structures of the ligand (in the hydrochloride form) and the polymer exhibit rotation of the pyrazine ring with respect to the quinoline skeleton. Initial optimization of the two models yielded the structures shown in Figure 14a,b. The structure from the left panel, corresponding to the free ligand, is more stable than the structure from the right panel, corresponding to the ligand coordinating the copper cation in the polymer. The difference is, however, only 5.0 kcal·mol−1, indicating that neither strong intramolecular attractive forces within the ligand nor significant steric hindrances exist. The lack of strong intramolecular interactions will be shown in Figure 18, on the basis of the QTAIM analysis, where the ligand exhibits only weak C-H···O and C-H···N contacts. The rotation of the pyrazine ring seems not hindered, and the form allowing better coordination capabilities and overall crystal packing is preferred in the polymer.
In Figure 15, the conformational analysis results are presented. We have investigated three dihedral angle rotations—the dihedral angle responsible for the carboxylic group rotation and two dihedral angles associated with the pyrazine ring rotation (N1-C8-C10-C13 and C9-C8-C10-C13). As it is shown in Figure 15, the carboxylic group exhibits free rotations; therefore, the obtained potential energy profile is symmetric. The most favorable conformation is the starting one obtained from the geometry optimization (see Figure 15 a,b,c). The degree of interaction of the carboxylic π electrons with the delocalized aromatic system is indicated by the height of the rotation barrier, 3.5 kcal·mol−1, which is not large. Concerning the pyrazine ring rotation, there is a steric hindrance visible in both presented charts, leading to a 9.5 kcal·mol−1 barrier for rotation. These two charts are similar, belonging to the same rotation, but with different conditioning during the scan. However, the most stable conformation in both cases is the one obtained from the geometry optimization and crystal structure (where the carboxylic and ethyl groups are located on the opposite sides).
In the next step, the models of polymers were investigated. We have constructed two models with two and three ligands attached to the Cu ion (denoted as small and large models, respectively). The obtained structures after the quantum-chemical simulations at the DFT/M06/def2-TZVP and DFT/M06/SDD are presented in Figure 16a,b.
For the polymers, spin density distribution analysis was performed, and the obtained results are presented in Figure 17a–d.
The spin density is not located only on the Cu metal center. For the small model, the spin density is located only on one of the two ligands, specifically the moiety binding by two nitrogen atoms. In the case of the larger model, the spin density is distributed to all three ligands, but its larger values are limited to the immediate vicinity of the binding nitrogen atoms. The relatively small range of the spin density distribution for the large model corresponds well with the also rather small experimentally derived exchange interaction, 2J = −1.07 ± 0.02 cm–1.
In order to complete our theoretical study, the topology analysis based on electron density partitioning was carried out. The obtained molecular graphs for all studied systems are presented in Figure 18. An application of the QTAIM theory enabled the detection of Bond and Ring Critical Points (BCPs and RCPs) in the case of the studied ligand and the small model of the polymer. Concerning the large model of the polymer, cage points (CPs) were found as well. The dotted lines indicate the presence of non-covalent interactions present in the studied molecules. As it is shown, the ligand spatial conformation is stabilized by interaction between hydrogen atoms from the ethyl group and quinoline ring with oxygen atoms from the carboxylic group and the nitrogen atom from the quinoline ring. Concerning the small model of the polymer, six non-covalent interactions were found. It is worth underlining that a BCP was found between the Cu ion and one of the hydrogen atoms from the ethyl group. In addition, being somehow in the middle, another hydrogen atom from the ethyl group is interacting with a hydrogen atom from the quinoline moiety. A stabilizing interaction between the oxygen atom from the carboxylic group and one of the hydrogens from the quinoline ring was also detected. Concerning the large model of the polymer, the presence of the third ligand introduced additional non-covalent interactions. As it is shown in Figure 18, two ligands of the polymer are in parallel arrangement. They interact with the Cu ion via oxygen from the carboxylic group and two nitrogen atoms from the pyrazine and quinoline moieties of the second ligand. As it is presented in Figure 18a–f, there is a network of non-covalent interactions between the parallelly arranged ligands, in line with their π-electron-rich nature, supporting their stacked spatial conformation in the studied model.

4. Conclusions

A new copper(II) 2D coordination polymer was synthesized via the solvothermal reaction between the luminescent aminocarboxylate ligand and a copper(II) precursor. This study enriches our knowledge of structural and luminescent properties of coordination polymers [38,39,40,41] with ligands derived from quinolinecarboxylic acid, possibly inspiring future research in this field. A holistic approach has been taken from design of the synthesis, through characterization of the product properties till computational modelling.
The DFT modeling of the ligand has revealed its preferred conformation and explained the conformational flexibility of the pyrazine ring so that the ligand can easily adapt its structure to provide better metal binding in the polymer. The spin density distribution maps do not exhibit long-range interactions away from the metal center. Interaction analysis within the QTAIM methodology reveals key factors for stabilization of the arrangement of ligands in the model.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst15080691/s1, Figure S1: (a) 1H NMR spectra of the ligand 2-pyridin-4-yl-quinoline-4-carboxylic acid HL (500 MHz, methanol-d4, 300 K) and, (b) 13C NMR spectrum of compound for {Cu(L)BF4}n (L = deprotonated 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate) 2 (125.77 MHz, methanol-d4, 300.0 K); Figure S2: Molecular graphs obtained based on QTAIM theory for the studied ligand and polymers (small and large models); Table S1: Crystal data and structure refinement; Table S2: Atomic coordinates (×104) and equivalent isotropic displacement parameters (Å2 × 103) for HL(2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate dihydrated chloride); Table S3: Bond lengths [Å] and angles [°] for HL; Table S4: Anisotropic displacement parameters (Å2 × 103) for HL; Table S5: Hydrogen coordinates (×104) and isotropic displacement parameters (Å2 × 103); Table S6: Torsion angles [°] for HL; Table S7: Atomic coordinates (×104) and equivalent isotropic displacement parameters (Å2 × 103) for {Cu(L)BF4}n (L = (L = deprotonated 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate) 2; Table S8: Bond lengths [Å] and angles [°] for {Cu(L)BF4}n; Table S9: Hydrogen coordinates (×104) and isotropic displacement parameters (Å2 × 103) for {Cu(L)BF4}n; Table S10: Hydrogen-bond geometry (Å, °) for the hydrated ligand hydrochloride (II) [2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate chloride dihydrate; Table S11: Hydrogen bonding geometry (Å, °) for {Cu(L)BF4}n (L = (L = deprotonated 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate) 2. coordination polymer.

Author Contributions

Conceptualization A.K.; methodology, A.K., A.J., and J.J.P.; validation, M.H.; investigation, A.K., A.J., and J.J.P.; writing—original draft preparation, A.K., A.J., J.J.P., and M.H.; writing—review and editing, A.K., A.J., and J.J.P.; visualization, A.K., A.J., and J.J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by IDUB grant (no: BPIDUB.17.2025; Wrocław University).

Data Availability Statement

The original contributions presented in this study are included in the article/supplementary material. Further inquiries can be directed to the corresponding authors.

Acknowledgments

A.J. and J.J.P. thank the Wrocław Center for Networking and Supercomputing (WCSS) in Wrocław for their generous CPU time resources and for allowing them to use the file-storage facilities.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Noro, S.; Kitagawa, S.; Kondo, M.; Seki, K. A New Methane Adsorbent, Porous Coordination Polymer [{CuSiF6(4,4′-bipyridine)2}n]. Angew. Chem. Int. Ed. 2000, 39, 2082–2084. [Google Scholar] [CrossRef]
  2. Peppas, A.; Sokalis, D.; Perganti, D.; Schnakenburg, G.; Falaras, P.; Philippopoulos, A.I. Sterically demanding pyridine-quinoline anchoring ligands as building blocks for copper(I)-based dye-sensitized solar cell (DSSC) complexes. Dalton Trans. 2022, 51, 15049–15066. [Google Scholar] [CrossRef]
  3. Appleby, M.V.; Cowin, R.A.; Ivalo, I.I.; Peralta-Arriaga, S.L.; Robertson, C.C.; Bartlett, S.; Fitzpatrick, A.; Dent, A.; Karras, G.; Diaz-Moreno, S.; et al. Ultrafast electronic, infrared, and X-ray absorption spectroscopy study of Cu(I) phosphine diimine complexes. Faraday Discuss. 2023, 244, 391–410. [Google Scholar] [CrossRef]
  4. Zhao, X.; Su, B.; Zhao, Q.; Lv, W.; Chen, L.; Huang, L.; Li, X.; Liao, S. Construction of pyridine-functionalized metal–organic frameworks for the detection of flazasulfuron. Acta Cryst. 2024, C80, 806–814. [Google Scholar] [CrossRef] [PubMed]
  5. Noro, S.; Kitaura, R.; Kondo, M.; Kitagawa, S.; Ishii, T.; Matsuzaka, H.; Yamashita, M. Framework Engineering by Anions and Porous Functionalities of Cu(II)/4,4‘-bpy Coordination Polymers. J. Am. Chem. Soc. 2002, 124, 2568–2583. [Google Scholar] [CrossRef]
  6. Blake, A.J.; Hill, S.J.; Hubberstey, P.; Li, W.S. Rectangular grid two-dimensional sheets of copper(II) bridged by both co-ordinated and hydrogen bonded 4,4′-bipyridine (4,4′-bipy) in [Cu(µ-4,4′-bipy)(H2O)2(FBF3)2]·4,4′-bip. J. Chem. Soc. Dalton Trans. 1997, 6, 913–914. [Google Scholar] [CrossRef]
  7. Kitaura, R.; Fujimoto, K.; Noro, S.I.; Kondo, M.; Kitagawa, S. A pillared-layer coordination polymer network displaying hysteretic sorption: [Cu2(pzdc)2(dpyg)]n (pzdc = pyrazine-2,3-dicarboxylate; dpyg=1,2-Di(4-pyridyl)glycol). Angew. Chem. Ed. Engl. 2002, 41, 133–135. [Google Scholar] [CrossRef]
  8. Hoskins, B.F.; Robson, R. Design and construction of a new class of scaffolding-like materials comprising infinite polymeric frameworks of 3D-linked molecular rods. A reappraisal of the zinc cyanide and cadmium cyanide structures and the synthesis and structure of the diamond-related frameworks [N(CH3)4][CuIZnII(CN)4] and CuI[4,4′,4″,4‴tetracyanotetraphenylmethane]BF4.xC6H5NO2. J. Am. Chem. Soc. 1990, 112, 1546–1554. [Google Scholar] [CrossRef]
  9. Yaghi, O.M.; Li, H. Hydrothermal Synthesis of a Metal-Organic Framework Containing Large Rectangular Channels. J. Am. Chem. Soc. 1995, 117, 10401–10402. [Google Scholar] [CrossRef]
  10. Lee, E.; Kim, J.; Heo, J.; Whang, D.; Kim, K. A Two-Dimensional Polyrotaxane with Large Cavities and Channels: A Novel Approach to Metal–Organic Open-Frameworks by Using Supramolecular Building Blocks. Angew. Chem. 2001, 113, 413–416. [Google Scholar] [CrossRef]
  11. Phuengphai, P.; Youngme, S.; Mutikainen, I.; Gamez, P.; Reedijk, J. A series of related 2D coordination polymers based on [copper(II)-4,4′-bpy-carboxylato] building blocks. Polyhedron 2012, 42, 10–17. [Google Scholar] [CrossRef]
  12. Geng, J.; Qin, L.; He, C.; Cui, G. Synthesis, crystal structures and catalytic properties of copper(II) and cobalt(II) coordination polymers based on a flexible benzimidazole ligand. Transit. Met. Chem. 2012, 37, 579–585. [Google Scholar] [CrossRef]
  13. Formica, M.; Fusi, V.; Giorgi, L.; Micheloni, M. New fluorescent chemosensors for metal ions in solution. Coord. Chem. Rev. 2012, 256, 170–192. [Google Scholar] [CrossRef]
  14. Sinzger, K.; Hunig, S.; Jopp, M.; Bauer, D.; Beitsch, W.; Wolf, C.H.; Kremer, R.K.; Metzenthin, T.; Bau, R.; Khan, S.I.; et al. The organic metal (Me2-DCNQI)2Cu: Dramatic changes in solid-state properties and crystal structure due to secondary deuterium effects. J. Am. Chem. Soc. 1993, 115, 7696–7705. [Google Scholar] [CrossRef]
  15. Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871. [Google Scholar] [CrossRef]
  16. Kohn, W.; Sham, L.J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138. [Google Scholar] [CrossRef]
  17. Bader Richard, F.W. Atoms in Molecules: A Quantum Theory; Oxford University Press: Oxford, MS, USA, 1994; ISBN 978-0-19-855865-1. [Google Scholar]
  18. Bader, R.F.W. A quantum theory of molecular structure and its applications. Chem. Rev. 1991, 91, 893–928. [Google Scholar] [CrossRef]
  19. Computer Programs: CrysAlis PRO, Rigaku OD, 2015; Rigaku Oxford Diffraction Ltd.: Abingdon, UK, 2020.
  20. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef]
  21. Sheldrick, G.M. A Short History of SHELX. Acta Cryst. 2008, A64, 112–122. [Google Scholar] [CrossRef]
  22. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  23. 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. [Google Scholar] [CrossRef] [PubMed]
  24. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef]
  25. Dunning, T.H., Jr.; Hay, P.J. Modern Theoretical Chemistry; Schaefer, H.F., III, Ed.; Plenum: New York, NY, USA, 1977; Volume 3, pp. 1–28. [Google Scholar]
  26. Igel-Mann, G.; Stoll, H.; Preuss, H. Pseudopotentials for main group elements (IIIA through VIIA). Mol. Phys. 1988, 65, 1321–1328. [Google Scholar] [CrossRef]
  27. Andrae, D.; Haeussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-adjusted ab initio pseudopotentials for the 2nd and 3rd row transition-elements. Theor. Chem. Acc. 1990, 77, 123–141. [Google Scholar] [CrossRef]
  28. 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.; et al. Gaussian 16, Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  29. Keith, T.A. Gristmill T.K. AIMAll, Version 19.10.12; TK Gristmill Software: Overland Park, KS, USA, 2019. Available online: https://aim.tkgristmill.com (accessed on 4 June 2025).
  30. Humphrey, W.; Dalke, A.; Schulten, K. VMD-Visual Molecular Dynamics. J. Mol. Graphics. 1996, 14, 33–38. [Google Scholar] [CrossRef] [PubMed]
  31. Dennington, R.; Keith, T.A.; Millam, J.M. GaussView, Version 6.1.1; Semichem Inc.: Shawnee Mission, KS, USA, 2019.
  32. Zhu, A.X.; Xu, Q.Q.; Liu, F.Y.; Qi, X.L. Synthesis, Characterization, and Photoluminescence Properties of a Planar Copper Triazolate Coordination Polymerwith Cyanide as Co-Ligand, Zeitschrift Für Anorg. Und Allg. Chem. 2011, 637, 502–505. [Google Scholar] [CrossRef]
  33. Rani, V.S.V.; Dhanasekaran, T.; Jayathuna, M.; Narayanan, V.; Jesudurai, D. Synthesis, characterization of Cu (II) Schiff base complexes: Optical and magnetic studies. Mater. Today Proc. 2018, 5, 8784–8788. [Google Scholar] [CrossRef]
  34. Hietsoi, O.; Filatov, A.S.; Dubceac, C.; Petrukhina, M.A. Structural diversity and photoluminescence of copper(I) carboxylates: From discrete complexes to infinite metal-based wires and helices. Coord. Chem. Rev. 2015, 295, 125–138. [Google Scholar] [CrossRef]
  35. Zink, D.M.; Bächle, M.; Baumann, T.; Nieger, M.; Kühn, M.; Wang, C.; Klopper, W.; Monkowius, U.; Hofbeck, T.; Yersin, H.; et al. Synthesis, structure, and characterization of dinuclear copper(I) halide complexes with P^N ligands featuring exciting photoluminescence properties. Inorg. Chem. 2013, 52, 2292–2305. [Google Scholar] [CrossRef]
  36. Bleaney, B.; Bowers, K.; Proc, K.D. Proceedings of the Royal Society of London. Ser. A Math. Phys. Sci. 1952, 214, 451–465. [Google Scholar] [CrossRef]
  37. Kahn, O. Molecular Magnetism; Wiley–VCH: New York, NY, USA, 1993. [Google Scholar] [CrossRef]
  38. Kachi-Terajjma, C.; Hagiwara, S. Synthesis, crystal structure and photophysical properties of chlorido-[(E)-3-hy-droxy-2-methyl-6-(quinolin-8-yldiazen-yl)phenolato]copper(II) monohydrate. Acta Crystallogr. E Crystallogr. Commun. 2022, 78, 473–476. [Google Scholar] [CrossRef] [PubMed]
  39. Kochel, A.; Hołyńska, M.; Twaróg, K.; Jezierska, A.; Panek, J.J.; Wojaczyński, J. A new mixed-valence CuI/CuII three-dimnsional coordination polymer constructed with an N,O-donor ligand generated via solvothermal synthesis: Structural features and magnetic properties. Acta Cryst. 2022, C78, 405–413. [Google Scholar] [CrossRef]
  40. Gogia, A.; Novikov, E.M.; Guzei, I.A.; Fonari, M.S. Crystal structure and characterization of a new one-dimensional copper(II) coordination polymer containing a 4-amino¬benzoic acid ligand. Acta Cryst. E 2024, 80, 330–334. [Google Scholar] [CrossRef]
  41. Scatena, R.; Massignani, S.; Lanza, A.E.; Zorzi, F.; Monari, M.; Nestola, F.; Pattinari, C.; Pandolfo, L. Synthesis of Coordination Polymers and Discrete Complexes from the Reaction of Copper(II) Carboxylates with Pyrazole: Role of Carboxylates Basicity. Cryst. Growth Des. 2022, 22, 1032–1044. [Google Scholar] [CrossRef]
Scheme 1. Synthesis of the 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate acid.
Scheme 1. Synthesis of the 2-(3-ethyl-pyrazin-2-yl)quinoline-4-carboxylate acid.
Crystals 15 00691 sch001
Scheme 2. Scheme of the synthesis of {Cu(L)BF4}n.
Scheme 2. Scheme of the synthesis of {Cu(L)BF4}n.
Crystals 15 00691 sch002
Scheme 3. Scheme of ligand hydrated chloride (1).
Scheme 3. Scheme of ligand hydrated chloride (1).
Crystals 15 00691 sch003
Figure 1. Molecular structure of the precursor ligand hydrated chloride. (Thermal ellipsoids are plotted at a 50% probability level).
Figure 1. Molecular structure of the precursor ligand hydrated chloride. (Thermal ellipsoids are plotted at a 50% probability level).
Crystals 15 00691 g001
Figure 2. View of graph set motif R 2,4(8) in the crystal structure 1.
Figure 2. View of graph set motif R 2,4(8) in the crystal structure 1.
Crystals 15 00691 g002
Figure 3. Asymmetric unit (a) and ligand coordination scheme (b).
Figure 3. Asymmetric unit (a) and ligand coordination scheme (b).
Crystals 15 00691 g003
Scheme 4. The obtained 2D coordination polymer.
Scheme 4. The obtained 2D coordination polymer.
Crystals 15 00691 sch004
Figure 4. Two-dimensional polymer network along the [100] and [010] directions.
Figure 4. Two-dimensional polymer network along the [100] and [010] directions.
Crystals 15 00691 g004
Figure 5. Stacking interactions in 2D coordination polymer.
Figure 5. Stacking interactions in 2D coordination polymer.
Crystals 15 00691 g005
Figure 6. Tubular polymeric arrangements in the crystal structure 2, (a,b) two different views.
Figure 6. Tubular polymeric arrangements in the crystal structure 2, (a,b) two different views.
Crystals 15 00691 g006
Figure 7. Extended crystal structure packing for 2 along [100].
Figure 7. Extended crystal structure packing for 2 along [100].
Crystals 15 00691 g007
Figure 8. Extended crystal structure packing for 2 along [001].
Figure 8. Extended crystal structure packing for 2 along [001].
Crystals 15 00691 g008
Figure 9. Photoluminescence emission spectra for ligand 1 excitation (black) wavelength λexc = 260 nm; excitation spectrum of ligand 1, recorded by monitoring emission (green) at λem = 425 nm) in the solid state at 297 K.
Figure 9. Photoluminescence emission spectra for ligand 1 excitation (black) wavelength λexc = 260 nm; excitation spectrum of ligand 1, recorded by monitoring emission (green) at λem = 425 nm) in the solid state at 297 K.
Crystals 15 00691 g009
Figure 10. Photoluminescence emission spectra for polymer 2 excitation (black) wavelength λexc = 365 nm; excitation spectrum of polymer 2, recorded by monitoring emission (green) at λem = 465 nm) in the solid state at 297 K.
Figure 10. Photoluminescence emission spectra for polymer 2 excitation (black) wavelength λexc = 365 nm; excitation spectrum of polymer 2, recorded by monitoring emission (green) at λem = 465 nm) in the solid state at 297 K.
Crystals 15 00691 g010
Figure 11. The shortest Cu(1)···Cu(1)i [i = x, 0.5 − y, 0.5 + z] distance is 6.5504(14) Å, justifying the approximation assuming only nearest-neighbor interactions.
Figure 11. The shortest Cu(1)···Cu(1)i [i = x, 0.5 − y, 0.5 + z] distance is 6.5504(14) Å, justifying the approximation assuming only nearest-neighbor interactions.
Crystals 15 00691 g011
Figure 12. Model of antiferromagnetic interactions transmitted via the organic ligand. Black arrows denote spin orientation.
Figure 12. Model of antiferromagnetic interactions transmitted via the organic ligand. Black arrows denote spin orientation.
Crystals 15 00691 g012
Figure 13. Temperature dependence of χm (Δ) and χmT (○) of (1). The solid line represents the calculated curve, according to the Bleaney–Bowers equation with the molecular field corrections (parameters seen in the text).
Figure 13. Temperature dependence of χm (Δ) and χmT (○) of (1). The solid line represents the calculated curve, according to the Bleaney–Bowers equation with the molecular field corrections (parameters seen in the text).
Crystals 15 00691 g013
Figure 14. Models of the ligands used in the theoretical investigations at the the DFT/M06/def2-TZVP level of theory: (a) conformation of the free ligand; (b) the ligand in the conformation taken from the polymer. Color coding: carbon—gray; nitrogen—blue; oxygen—red; hydrogen—white.
Figure 14. Models of the ligands used in the theoretical investigations at the the DFT/M06/def2-TZVP level of theory: (a) conformation of the free ligand; (b) the ligand in the conformation taken from the polymer. Color coding: carbon—gray; nitrogen—blue; oxygen—red; hydrogen—white.
Crystals 15 00691 g014
Figure 15. (ac) Potential energy profiles of dihedral angle rotation of the studied ligand obtained at the DFT/M06/def2-TZVP level of theory.
Figure 15. (ac) Potential energy profiles of dihedral angle rotation of the studied ligand obtained at the DFT/M06/def2-TZVP level of theory.
Crystals 15 00691 g015aCrystals 15 00691 g015b
Figure 16. (a,b) Models of polymers used in the theoretical investigations: left—small model; right—larger model. Color coding: carbon—gray; nitrogen—blue; oxygen—red; hydrogen—white; copper—brown.
Figure 16. (a,b) Models of polymers used in the theoretical investigations: left—small model; right—larger model. Color coding: carbon—gray; nitrogen—blue; oxygen—red; hydrogen—white; copper—brown.
Crystals 15 00691 g016
Figure 17. (ad) Spin density distribution obtained for the studied polymers (isosurface = 0.0004 e/Å3). Small model—upper part; larger model—lower part. Two visualizations were prepared for clarity. Spinα denoted in blue color, while spinβ is given in green. Color coding: carbon—gray; nitrogen—blue; oxygen—red; hydrogen—white; copper—light brown.
Figure 17. (ad) Spin density distribution obtained for the studied polymers (isosurface = 0.0004 e/Å3). Small model—upper part; larger model—lower part. Two visualizations were prepared for clarity. Spinα denoted in blue color, while spinβ is given in green. Color coding: carbon—gray; nitrogen—blue; oxygen—red; hydrogen—white; copper—light brown.
Crystals 15 00691 g017aCrystals 15 00691 g017b
Figure 18. (af) Molecular graphs obtained based on QTAIM theory for the studied ligand and polymers (small and large models). All models are presented using two different positions to make all non-covalent interactions detected by the topological analysis more visible. The dotted lines indicate the presence of non-covalent interactions. Color coding: carbon—gray; nitrogen—blue; oxygen—red; hydrogen—white; copper—light brown. The critical points are not shown for clarity. Figure S2 in the Supplement contains the location of the bond, ring, and cage critical points.
Figure 18. (af) Molecular graphs obtained based on QTAIM theory for the studied ligand and polymers (small and large models). All models are presented using two different positions to make all non-covalent interactions detected by the topological analysis more visible. The dotted lines indicate the presence of non-covalent interactions. Color coding: carbon—gray; nitrogen—blue; oxygen—red; hydrogen—white; copper—light brown. The critical points are not shown for clarity. Figure S2 in the Supplement contains the location of the bond, ring, and cage critical points.
Crystals 15 00691 g018
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kochel, A.; Hołyńska, M.; Jezierska, A.; Panek, J.J. Design and Synthesis of a New Photoluminescent 2D Coordination Polymer Employing a Ligand Derived from Quinoline and Pyridine. Crystals 2025, 15, 691. https://doi.org/10.3390/cryst15080691

AMA Style

Kochel A, Hołyńska M, Jezierska A, Panek JJ. Design and Synthesis of a New Photoluminescent 2D Coordination Polymer Employing a Ligand Derived from Quinoline and Pyridine. Crystals. 2025; 15(8):691. https://doi.org/10.3390/cryst15080691

Chicago/Turabian Style

Kochel, Andrzej, Małgorzata Hołyńska, Aneta Jezierska, and Jarosław J. Panek. 2025. "Design and Synthesis of a New Photoluminescent 2D Coordination Polymer Employing a Ligand Derived from Quinoline and Pyridine" Crystals 15, no. 8: 691. https://doi.org/10.3390/cryst15080691

APA Style

Kochel, A., Hołyńska, M., Jezierska, A., & Panek, J. J. (2025). Design and Synthesis of a New Photoluminescent 2D Coordination Polymer Employing a Ligand Derived from Quinoline and Pyridine. Crystals, 15(8), 691. https://doi.org/10.3390/cryst15080691

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop