Next Article in Journal
A Single Dose of AC102 Reverts Tinnitus by Restoring Ribbon Synapses in Noise-Exposed Mongolian Gerbils
Next Article in Special Issue
Microheterogeneity in Liquid Water Associated with Hydrogen-Bond Cooperativity-IR Spectroscopic and MD Simulation Study of Temperature Effect
Previous Article in Journal
Rootstock–Scion Exchanging mRNAs Participate in Watermelon Fruit Quality Improvement
Previous Article in Special Issue
Unveiling the Nature and Strength of Selenium-Centered Chalcogen Bonds in Binary Complexes of SeO2 with Oxygen-/Sulfur-Containing Lewis Bases: Insights from Theoretical Calculations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structure–Property Relationships in Zwitterionic Pyridinium–Triazole Ligands: Insights from Crystal Engineering and Hirshfeld Surface Analysis

1
Departamento de Química, Facultad de Ciencias, Universidad de los Andes, Mérida 5101, Venezuela
2
Departamento de Química, Facultad de Ciencias, Universidad de Católica del Norte, Sede Casa Central, Av. Angamos 0610, Antofagasta 1270709, Chile
3
Departamento de Química, Universidad de Antofagasta, Av. Universidad de Antofagasta 02800, Campus Coloso, Antofagasta 1240000, Chile
4
Instituto Universitario de Ciencia de Materiales “Nicolás Cabrera” (INC), Universidad Autónoma de Madrid, Campus de Cantoblanco, 28049 Madrid, Spain
5
Departamento de Física Aplicada, Universidad Autónoma de Madrid, 28049 Madrid, Spain
6
Departamento de Ingeniería Química y Procesos de Minerales, Universidad de Antofagasta, Avenida Angamos 601, Antofagasta 1270300, Chile
7
Departamento de Química Inorgánica, Universidad Autónoma de Madrid, 28049 Madrid, Spain
8
Institute for Advanced Research Chemistry (IAdChem), Universidad Autónoma de Madrid, 28049 Madrid, Spain
9
Condensed Matter Physics Center (IFIMAC), Universidad Autónoma de Madrid, 28049 Madrid, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(11), 5123; https://doi.org/10.3390/ijms26115123
Submission received: 9 April 2025 / Revised: 18 May 2025 / Accepted: 22 May 2025 / Published: 27 May 2025

Abstract

This article discloses the synthesis of four new positional isomeric zwitterionic ligands exhibiting semi-flexible and flexible characteristics—n-pyridinium-1,2,3-triazole-4-carboxy-5-Acetate (n-PTCA), and n-methylpyridinium-1,2,3-triazole-4-carboxy-5-Acetate (n-MPTCA; where n = 3, 4)—which were derived from an aqueous solution of the corresponding sodium salts in an acidic medium (HCl). These compounds are successfully synthesized and characterized with FT-IR and multinuclear NMR spectroscopy; likewise, proper single crystals are obtained for each compound. All compounds adopt zwitterionic forms in the solid state, which are stabilized via intermolecular proton transfer processes involving HCl and solvent molecules. A single-crystal X-ray analysis revealed how positional isomerism and molecular flexibility influence the supramolecular topology. Specifically, 3-PTCA and 4-PTCA exhibit isomorphic hydrogen bond networks, while 3-MPTCA and 4-MPTCA display distinct packing motifs, attributed to the presence of a methylene spacer between the pyridinium and triazole rings. The Hirshfeld surface analysis quantitatively confirmed the dominance of O···H/H···O and N···H/H···N interactions in the solid-state architecture. These strong hydrogen-bonding networks are indicative of the potential proton-conductive behavior in the crystalline state, positioning these compounds as promising candidates for applications in proton-conducting materials. The structural insights gained underscore the pivotal role of molecular topology in tailoring crystal packing, with implications for the rational design of zwitterionic ligands in functional materials, including MOFs and coordination polymers. The calculated HOMO-LUMO energy gaps reveal a significant electronic variability among the ligands, influenced primarily by the positional isomerism and structural flexibility introduced by the methylene spacer.

1. Introduction

N-heterocycles play a key role in biological and synthetic chemistry, forming the backbone of nucleic acid structures and many pharmacologically active molecules [1,2]. Among them, azoles—particularly 1,2,3-triazoles—have garnered significant attention due to their chemical stability, electronic properties, and synthetic versatility [3]. The Cu(I)-catalyzed azide–alkyne cycloaddition (CuAAC), a prototypical “click” reaction, provides efficient and regioselective access to 1,4-disubstituted triazoles, making them valuable scaffolds in drug discovery, catalysis, and materials science [4,5].
Beyond their molecular reactivity, 1,2,3-triazole-based ligands have demonstrated significant potential in the design of multitopic coordination polymers (CPs) and metal–organic frameworks (MOFs), owing to their multidentate coordination modes and tendency to form directional hydrogen bonds [6,7,8]. Recent studies have explored their incorporation into functional materials such as corrosion inhibitors, luminescent complexes, and photovoltaic devices [9,10,11]. Their nitrogen-rich frameworks and geometrical rigidity render them excellent candidates for constructing ordered supramolecular architectures.
Concurrently, zwitterionic ligands—bearing spatially separated positive and negative charges—have emerged as promising building blocks in crystal engineering [12,13,14]. These molecules can form robust hydrogen bond networks and electrostatic interactions, favoring the self-assembly of extended crystal structures [15,16,17]. Despite their potential, the deliberate use of zwitterions in engineering supramolecular networks and coordination materials remains underexplored, particularly regarding the role of positional isomerism and molecular flexibility.
From a crystal engineering perspective, molecular design strategies that modulate parameters such as the spacer length, substitution pattern, and charge distribution allow precise control over intermolecular interactions [18,19,20]. These, in turn, influence crystal packing, hydrogen-bonding topology, and emergent functional properties—including proton conduction [21,22]. The synergy between positional isomerism and supramolecular organization has proven crucial for tailoring material behavior, particularly in hydrogen-bond-mediated proton-conducting solids [23,24,25].
Zwitterionic materials have recently attracted growing interest for their proton-conducting capabilities, making them promising candidates for applications such as proton-exchange membrane fuel cells (PEMFCs) and solid-state electrolytes. For instance, zwitterion-functionalized nanofiber composites have demonstrated an enhanced proton mobility due to continuous hydrogen bond pathways [26,27,28], while zwitterionic polymers have shown competitive proton conductivities and an improved stability under humid conditions [29,30,31]. A recent study reported a novel zwitterion-based membrane applied successfully in PEM fuel cells, highlighting their potential as next-generation functional materials [32,33,34].
Herein, this work reports the synthesis, crystal structures, and supramolecular properties of four novel zwitterionic ligands based on n-pyridinium-1,2,3-triazole-4-carboxy-5-Acetate (n-PTCA) and n-methylpyridinium-1,2,3-triazole-4-carboxy-5-Acetate (n-MPTCA; where n = 3,4). These compounds, obtained under acidic aqueous conditions from their corresponding sodium salts, exhibit either semi-flexible or flexible molecular backbones. The presence of a methylene spacer in selected isomers introduces conformational adaptability, which directly impacts the crystal packing and hydrogen bond arrangements.
Through a single-crystal X-ray diffraction and Hirshfeld surface analysis, we examine how isomeric variation and ligand flexibility influence non-covalent interaction patterns and the supramolecular architecture. Special attention is given to the role of hydrogen bond networks as potential proton conduction pathways, emphasizing the promise of these zwitterionic systems for crystal engineering and functional materials applications.

2. Results and Discussion

2.1. Syntheses

The zwitterionic ligands 3-, 4-PTCA, and 3-, 4-MPTCA were synthesized via a stepwise sequence starting from their corresponding ester precursors, which were obtained through a [3 + 2] dipolar cycloaddition between n-pyridyl azides (3-PTCA and 4-PTCA) or n-(azidomethyl)pyridine (3-MPTCA and 4-MPTCA) and diethyl 1,3-acetonedicarboxylate, following previously reported procedures [35,36,37,38]. The resulting esters were saponified with aqueous NaOH to generate corresponding sodium salts in quantitative yields. The subsequent acidification with dilute HCl (0.1 N) led to the formation of the zwitterionic ligands described above as hygroscopic solids, which were stored under dry or inert conditions to avoid degradation.
The final compounds are highly soluble in water and moderately soluble in protic solvents, such as ethanol and hot methanol. Their general synthetic pathway is depicted in Scheme 1, which highlights the three-step protocol involving the cycloaddition, base hydrolysis, and acid-mediated zwitterion formation.

2.2. Crystal Structure

The crystal data, data collection, and refinement parameters for all compounds are shown in Table 1.
The single-crystal X-ray diffraction confirmed that all four compounds crystallize in the zwitterionic form, with the positively charged pyridinium nitrogen and a deprotonated carboxylate group. The formation of these internal ion pairs likely results from the intermolecular proton transfer from both the carboxylic acid moiety and hydrochloric acid to the solvent and nitrogen sites. 3- and 4-PTCA crystallize in the orthorhombic space group Pna21, while 3-MPTCA adopts the orthorhombic P212121 and 4-MPTCA the monoclinic P21/c space group, all with Z = 4. These space groups reflect the influence of both positional isomerism and structural flexibility on the packing arrangement and intermolecular interactions. Bond distances and angles across all structures fall within typical values [39] and are consistent with CSD reference standards [40]. Proton positions were refined using a riding model, and both N–H and O–H groups participate in an extensive hydrogen-bonding network, where O- and N- bound H atoms were initially located in a difference Fourier map and were then added in idealized positions and further refined according to the riding model with O−H = 0.82 Å and N−H =0.86 Å and with Uiso(H) = 1.5Ueq(O) or 1.2Ueq(N).
Figure 1 illustrates the asymmetric units of all compounds, emphasizing the conformational variability introduced by the methylene spacer in 3- and 4-MPTCA. Dihedral angles between the triazole, pyridinium, and carboxylate groups (Table 2) reveal notable differences in the molecular geometry among isomers. These torsional differences serve as a foundation for understanding the observed diversity in the supramolecular assembly.
A CSD search revealed no previously reported zwitterionic structures for this specific triazole–pyridinium–carboxylate framework, highlighting the novelty of the present ligands.
Solid-state studies show the cooperative actions of non-covalent interactions. The molecular structure and crystal packing of all compounds are stabilized by intermolecular O−H···O, N−H···O, and O−H···O/N−H···O (O-atom carboxylate group) hydrogen bond interactions (see Table 3). In all compounds, the supramolecular structure is established as infinite one-dimensional chains and rings exhibiting a complex binary graph-set with varying degrees of the pattern extending along different directions. These supramolecular structures are held together by hydrogen bond interactions (C−H···N and C−H···O), giving rise to three-dimensional networks in all structures.
The compounds 3- and 4-PTCA have identical supramolecular structures, known as isographic hydrogen bond patterns [41]. The molecules in these compounds are connected by chains with the graph-set motifs of C(8); C(9) ;   C 2 1 (11); C 2 2 (17); and C 4 3 (28) and a non-centrosymmetric ring with a graph-set motifs of R 6 5 (44). Figure 2 shows a chain C(9) and a non-centrosymmetric ring R 6 5 (44) as a representative example of the supramolecular structure for these two compounds. The atomic, molecular, supramolecular, and spectroscopic properties of compounds 3- and 4-PTCA are very similar, as well as other physical properties such as solubility and melting points; however, the conformational properties and the fingerprint plots are different, as demonstrated by the dihedral angles indicated in Table 2 and Figures S1–S4, respectively. For further clarity, the isostructurality was quantitatively deciphered from the commonly used method, which uses the unit cell parameters of two crystal structures to calculate the unit cell similarity ( Π ) [42], when Π = 0 .02004, as being quite isostructural between them.
For compounds 3- and 4-MPTCA, the supramolecular structures do not exhibit isographic hydrogen bond patterns like those in compounds 3- and 4-PTCA. In the case of 3-MPTCA, the molecules are interconnected by chains and a ring with a graph-set motifs of C(8); C(10); C 2 1 (14); C 2 2 (18); R 6 6 (48); and R 6 6 (52), respectively. Figure 3 depicts C(8) chains and R 6 6 (52) rings, as representative supramolecular structure examples.
In the case of 4-MPTCA, the molecules are interconnected through a complex network of hydrogen-bonded chains and rings with the following graph-set motifs: C(8); C(11); R 2 2 (22); R 4 4 (30); C 4 4 (34); R 4 4 (38); R 6 6 (50); R 6 6 (56); C 2 1 (13); C 2 2 (19); C 4 3 (32); R 6 4 (42); R 6 5 (48); R 6 6 (54); C 1 2 (8); C 2 2 (22); C 3 4 (30); R 4 6 (38); R 5 6 (52); and R 6 6 (66). Figure 4 shows C(8) chains and R 2 2 (22) rings, as a representative supramolecular structure of 4-MPTCA. This compound shows a greater number of supramolecular interactions than its positional isomer, 3-MPTCA. The position of the nitrogen atom on the pyridinium ring in this compound allows for greater supramolecular diversity.
The geometrical parameters corresponding to these non-covalent interactions are summarized in Table 3.
The conformational properties demonstrated by the dihedral angles indicated in Table 2 and the fingerprint plots generated from the crystal structures (Figures S1–S4) account for their non-isostructural character.

2.3. Hirshfeld Surface and Topological Studies

Intermolecular interactions in all compounds were verified using the CrystalExplorer software (version 17.5; University of Western Australia: Perth, Australia) [43] through a Hirshfeld surface analysis [44] and two-dimensional fingerprint plots [45]. The normalized contact distance (dnorm) surfaces were generated by mapping the internal (di) and external (de) distances between surface points and the nearest atoms in the crystal lattice. On these surfaces, short contacts (stronger than the sum of van der Waals radii) are represented in red, contacts close to the sum appear in white, and longer contacts are shown in blue (Figure 5). These visualizations highlight key regions of non-covalent interactions.
The most significant interactions observed across all compounds are O–H···O, N–H···O, and C–H···O hydrogen bonds, which appear as intense red regions on the Hirshfeld surfaces. These contacts correspond to strong directional interactions that govern the supramolecular organization of the crystals and are consistent with the hydrogen bonds observed in the X-ray diffraction analysis.
To gain quantitative insight, fingerprint plots were constructed by correlating the di and de values in increments of 0.01 Å. These plots allow the decomposition of the surface into contributions from specific contact types (Figures S1–S4). In all structures, O···H/H···O interactions are the dominant contributors to the total surface area, accounting for 39.5–43.7%, followed by N···H/H···N, H···H and C···H/H···C interactions to a lesser extent.
For example, in compound 3-PTCA, the fingerprint plot shows the following contributions: O···H/H···O (42.5%), N···H/H···N (20.1%), H···H (14.1%), C···H/H···C (8.3%), and C···O/O···C (7.8%). Minor contributions include C···C (1.6%), C···N/N···C (2.6%), O···N (2.2%), N···N (0.7%), and O···O (0.1%), which have a minimal directional influence on the molecular packing (Tables S1–S4).
The O···H/H···O interactions are represented in the fingerprint plots as two sharp symmetrical spikes centered around de + di ≈ 1.5–1.7 Å, indicating strong hydrogen bonding. Similarly, N···H/H···N contacts manifest as symmetrical wings between 2.4 and 2.7 Å, while C···H/H···C contacts—contributing 8.3–11.1%—show broader wings at 3.0–3.3 Å, which is suggestive of possible C–H···π interactions within the crystal packing.
This combined structural and surface analysis confirms the central role of hydrogen bonding and positional isomerism in dictating the supramolecular organization of these zwitterionic ligands.
Finally, the energy framework [46] was analyzed to gain a better understanding of the crystal packing, topology, and supramolecular arrangement. This method allows for the calculation and comparison of different energy components, including repulsion (E_rep), electrostatic (E_ele), dispersion (E_dis), polarization (E_pol), and total energy (E_tot), based on the anisotropy of pairwise intermolecular interaction energies (see Figures S5–S8 and Tables S1–S4). The cylinder thickness represents the strength of the interactions, directly correlating with the energy magnitude and providing insights into the stabilization of the crystal packing [46].
The orientation of the tubes suggests that the framework formation is driven by translational or centrosymmetric elements. However, this arrangement also facilitates the emergence of additional weak interactions within the crystal structure.
The calculations revealed that dispersion interactions form a distorted zig-zag ladder-like topology in the compounds with a zst topology for compounds 3-PTCA and 4-PTCA, meanwhile compounds 3-MPTCA and 4-MPTCA have a dia and tcg-x topology, respectively (see Figure 6). A significant difference between E_ele and E_pol was observed, likely due to the presence of classical 1D hydrogen bond interactions.

2.4. Density Functional Theory (DFT) Calculations

The electronic structure and reactivity parameters of the synthesized ligands were investigated through DFT calculations. All calculations were performed using the hybrid B3LYP functional in combination with the def2-TZVP basis set (see Section 3.4). This level of theory was selected for its well-established accuracy in describing the electronic properties of organic molecules, particularly frontier molecular orbitals and global reactivity descriptors. The calculated electronic parameters, including frontier molecular orbitals (Highest Occupied Molecular Orbital—HOMO and Lowest Unoccupied Molecular Orbital—LUMO), ionization potentials, electron affinities, electronegativities, hardness, softness, and electrophilicity indices, are presented in Table 4. The calculated HOMO-LUMO energy gaps (Egap) for ligands were found to be 1.88 eV (3-PTCA), 2.02 eV (4-PTCA), 4.04 eV (3-MPTCA), and 3.10 eV (4-MPTCA). These values reveal significant electronic variability among the ligands, influenced primarily by the positional isomerism and structural flexibility introduced by the methylene spacer. Ligand 3-MPTCA exhibited the largest HOMO-LUMO gap, indicating a higher kinetic stability and lower chemical reactivity compared to ligands 3-PTCA, 4-PTCA, and 4-MPTCA. Conversely, ligand 3-PTCA displayed the narrowest HOMO-LUMO gap, suggesting a greater chemical reactivity and potential ease in forming coordination bonds. The global hardness (η), reflecting the resistance to electron transfer, varied accordingly, with 3-MPTCA having the highest hardness value (2.02 eV), which is consistent with its large HOMO-LUMO gap. On the other hand, ligand 3-PTCA exhibited the lowest hardness value (0.94 eV), corroborating its higher chemical reactivity. Additionally, ligand 3-MPTCA had the lowest electrophilicity index (4.53 eV), indicating the lowest propensity to accept electrons, while ligand 3-PTCA had the highest electrophilicity index (11.18 eV), marking it as the most electrophilic of the set.
Figure 7 displays the computed frontier molecular orbitals (HOMO and LUMO) obtained by DFT calculations for all ligands. In general, the HOMO is primarily localized on the triazole nitrogen atoms and the carboxylate oxygens, regions associated with electron-rich sites capable of coordination. Conversely, the LUMO predominantly encompasses the pyridinium and triazole rings, reflecting an electron-accepting character. The presence of the flexible methylene spacer in 3-MPTCA and 4-MPTCA slightly disrupts the extended conjugation, inducing subtle but meaningful differences in the orbital distribution compared to their more rigid analogs (3-PTCA and 4-PTCA). These differences suggest that both positional isomerism and ligand flexibility significantly modulate the electronic structure and coordination potential of these promising N/O mixed ligands.

3. Materials and Methods

3.1. Chemicals

Chemicals were A.R. grade and used without further purification (hydrochloric acid: 32%, chemically pure grade, Sigma-Aldrich, St. Louis, MO, USA). The syntheses of the ester precursor and the sodium salt of the carboxylate ligands were carried out according to the procedure previously described and reported [38].

3.2. Characterization

(3-PTCA) Colorless parallelepipeds, m.p. 241–242 °C, FT-IR (KBr pellet, cm−1): 3482(w) ν(N–H), 3401(w) (O–H), (3070(w), 3048(w) ν(Csp2–H) 2939(vw), 2925(vw) ν(Csp3–H),1467(m), ν(C=O and/or C=C), 1618(s), ν(C–N)1410(m), ν(N=N and/or -CH3), 1260(vw), ν(C–O), 1260(w), ν(Csp2–N), 551 (vw); 1H-NMR (300 MHz, DMSO-d6) δ 8.82 (dd, J = 4.8, 1.5 Hz, 1H, H2), 8.79 (d, J = 2.5 Hz, 1H, H4), 8.07 (ddd, J = 8.2, 2.6, 1.5 Hz, 1H, H5), 7.71 (dd, J = 8.2, 4.8 Hz, 1H, H6), 4.09 (s, 1H, -CH2-);13C-NMR (75 MHz, DMSO-d6) δ 169.31 (C11), 162.23(C9), 151.45(C4), 146.03(C2), 137.92 (C9-ipso), 137.29 (C1-ipso), 133.51(C6), 132.19(C8-ipso), 124.77(C5), and 29.69(-CH3 Acetic acid). (See Figures S9 and S10).
(4-PTCA) Colorless parallelepipeds, m.p. 244–245 °C, FT-IR (KBr pellet,cm−1): 3482(w) ν(N–H), 3195(vs) ν(O–H), 3195(w), 3057(m) ν(Csp2–H), 2990(w), 2927(m) ν(Csp3–H), 1614(vs) ν(C=O and/or C=C), 1597(s) ν(C–N), 1426(m) ν(N=N and/or -CH3), 1290(m) ν(C–O), 1248(w) ν(Csp2–N), 558 (vw); 1H-NMR (300 MHz, DMSO-d6) δ 2.51 (d, J = 6.1 Hz, 7H), 4.20 (s, 1H), 7.70 (d, J = 5.2 Hz, 1H), 8.88 (d, J = 5.0 Hz, 1H); 13C-NMR (75 MHz, DMSO-d6) δ 30.16 (C10), 119.60 (C14, C18), 137.15 (C1, C2), 142.68 (C6), 152.12 (C15, C17), 162.48 (C7), 169.58 (C11). (see Figures S11 and S12).
(3-MPTCA) Colorless flat needle, m.p. 234–235 °C, FT-IR (KBr pellet, cm−1): 3482(w) ν(N–H), 3459(vs) ν(O–H), 3112(w), 2993(w) ν(Csp2–H), 2930(m), 2993(w), ν(Csp3–H),1603(w),1637 (m) ν(C=O and/or C=C), 1637(m), ν(C–N), 1482(m) ν(N=N and/or -CH3), 1367(w) ν(C–O), 1221(s), ν(Csp2–N), 501 (m); 1H -NMR (300 MHz, DMSO) δ 4.21 (s, 2H, 16), 5.76 (s, 2H, 6), 7.44 (ddd, J = 7.9, 4.8, 0.8 Hz, 1H, 11), 7.73 (dt, J = 7.9, 2.0 Hz, 1H, 12), 8.54–8.65 (m, 2H, 13, 15); 13C-NMR (75 MHz, DMSO) δ 169.66 (C17), 162.66 (C7), 149.70 (C15), 149.54 (C13), 137.93 (C11), 136.86 (C1), 136.34 (C2), 131.45 (C10), 124.18 (C12), 48.90 (C6), and 29.55 (C16). (See Figures S13 and S14).
(4-MPTCA) Colorless flat needle, m.p. 239–240 °C, FT-IR (KBr pellet, cm−1): 3482(w) ν(N–H), 3461(vs) ν(O–H), 3193(w), 3019(w) ν(Csp2–H), 2936(w), 2990(m) ν(Csp3–H),1614(vs), 1597(s) ν(C=O and/or C=C), 1426(vs) ν(C–N), 1278(m) ν(N=N and/or -CH3), ν(C–O), ν(Csp2–N), 554(vw); 1H-NMR (300 MHz, DMSO-d6) δ 3.87 (s, 8H, 16), 4.15 (s, 0H), 5.80 (s, 0H, 6), 7.20–7.28 (m, 0H, 12, 14), 8.55–8.64 (m, OH, 11, 15); 13C -NMR (75 MHz, DMSO-d6) δ 29.55 (C16), 50.06 (C6), 122.79 (C11, C15), 137.18 (C2), 138.05 (C1), 144.63 (C10), 150.27 (C12, C14), 162.65 (C7), and 169.58 (C17). (See Figures S15 and S16).
FT-IR spectra in the range 400–4000 cm−1 were recorded on a Thermo Nicolet Avatar 300 spectrometer using KBr pellets (Thermo Scientific, Waltman, MA, USA). Colorless crystals of the title compounds suitable for X-ray diffraction analysis were selected and measured. Diffraction data were collected at 295(2) K on a Bruker D8 Venture diffractometer equipped with a Photon-III C14 detector, using graphite monochromated MoKα (λ = 0.71073 Å) and Cu-Kα (λ = 1.54178 Å) radiation (Bruker Co.; Billerica, MA, USA). The diffraction frames were integrated using the APEX4 package [47] and were corrected for absorptions with SADABS [48]. The crystal structures of the title compounds were solved by intrinsic phasing [49] using the OLEX2 software (version 15.0, O.dolomanov; Department of Chemistry, Durham University) [50] and refined with full-matrix least-squares methods based on F2 (SHELXL) [49]. All the non-hydrogen atoms were refined with anisotropic displacement parameters. The hydrogen atoms were included from calculated positions and refined riding on their respective carbon atoms with isotropic displacement parameters. The structures of 3-, 4-PTCA, and 3-MPTCA were solved as merohedral twin specimens with the twin law in the reciprocal space expressed by the matrix [−1 0 0; 0–1 0; 0 0 −1]. All geometrical calculations were performed using the program Platon [51].

3.3. Hirshfeld Surface and Topological Analysis

The Hirshfeld surface and corresponding 2D fingerprint plots [45] were computed using CrystalExplorer 17.3 [52], with the crystallographic information file (CIF) as input [44]. The normalized contact distance (dnorm) was determined based on de, di, and the van der Waals radii (VdW) of atoms, following Equation (1). Here, de and di represent the distances from the Hirshfeld isosurface to the nearest external and internal nuclei, respectively, while VdW values were obtained from the literature [53,54].
d n o r m = d i r i v d W r i v d W + d e r e v d W r e v d W
Additionally, energy framework calculations were performed at the B3LYP/6-31G(d,p) level of theory over a range of ±0.002 a.u. [55], using the TONTO computational package (D. Jayatilaka & D. J. Grimwood; Chemistry, School of Biomedical and Chemical Sciences, University of Western Australia, Crawley, Australia) [56], integrated into CrystalExplorer [55]. Hydrogen bond lengths were normalized to standard neutron diffraction values (C–H = 1.083 Å, N–H = 1.009 Å, and O–H = 0.983 Å). Intermolecular interaction energies for molecular pairs in the crystal packing were calculated at the B3LYP/6-31G(d,p) level within a 3.8 Å radius cluster around the molecule [57,58].
Finally, topological analysis was supported by the topcryst.com [59]. The RCSR three-letter codes [60] were used to designate the network topologies. Those nets, that are absent in the RCSR, are designated with the TOPOS NDn nomenclature [61], where N is a sequence of coordination numbers of all non-equivalent nodes of the net, D is periodicity of the net (D = M, C, L, and T for 0-,1-,2-,3-periodic nets), and n is the ordinal number of the net in the set of all non-isomorphic nets with the given ND sequence. To calculate the underlying nets, we used algorithms3, the application of which for specific structures is discussed in the article [62]. The TTD collection [62] was used to determine the topological type of the crystal structure.

3.4. Density Functional Theory Calculations

All DFT calculations were carried out using the ORCA software package (version 6.0.1), developed by Prof. Frank Neese and the Molecular Theory and Spectroscopy Group at the Max-Planck-Institut für Kohlenforschung, Mülheim an der Ruhr, Germany, and distributed by FACCTs GmbH (https://www.faccts.de/orca/, accessed on 21 May 2025) [63,64]. Geometry optimization and vibrational frequency analyses were performed employing the hybrid B3LYP functional with the def2-TZVP basis set. The optimized structures were confirmed as true minima on the potential energy surface by ensuring no imaginary frequencies were observed. Frontier molecular orbitals HOMO and LUMO were computed at the same theoretical level (B3LYP/def2-TZVP). These energies were utilized to calculate the global reactivity parameters (GRPs): ionization potential (IP), electron affinity (EA), global hardness (η), global electronegativity (X), chemical potential (μ), global electrophilicity (ω), and global softness (σ).

4. Conclusions

In this study, four novel zwitterionic ligands—3-PTCA, 4-PTCA, 3-MPTCA, and 4-MPTCA—were synthesized and structurally characterized through a single-crystal X-ray diffraction and Hirshfeld surface analysis. All compounds crystallize in the zwitterionic form in the solid state, which is stabilized by strong O–H···O and N–H···O hydrogen bonds that define robust supramolecular networks.
Despite their structural similarities, positional isomerism and the presence of the methylene spacer in the MPTCA ligands significantly influence the crystal packing. While 3-PTCA and 4-PTCA exhibit isostructural hydrogen bond networks, 3-MPTCA and 4-MPTCA display distinct supramolecular topologies, highlighting the impact of molecular flexibility on non-covalent interactions. The dihedral angle variations and graph-set analysis further support these structural differences.
Hirshfeld surface and fingerprint plot analyses revealed that O···H/H···O interactions are the dominant contributors to the crystal cohesion, followed by N···H/H···N and C···H/H···C contacts. These findings underscore the importance of hydrogen-bonding interactions in directing the self-assembly of these zwitterionic ligands.
From a crystal engineering perspective, these results demonstrate how fine-tuning molecular topology through an isomeric variation and spacer incorporation allows for the control over the supramolecular architecture. The presence of well-organized hydrogen bond networks also suggests the potential of these ligands for proton-conducting applications, particularly in the design of zwitterionic metal–organic frameworks (ZW-MOFs) or coordination polymers (CPs). Future work will focus on evaluating the physicochemical properties of these materials, including their proton conductivity under humid conditions.
The calculated HOMO-LUMO energy gaps (EGap) for ligands 3-PTCA, 3-PTCA, 3M-PTCA, and 4-MPTCA were found to be 1.88 eV (3-PTCA), 2.02 eV (3-PTCA), 4.04 eV (3M-PTCA), and 3.10 eV (4M-PTCA). These values reveal a significant electronic variability among the ligands, influenced primarily by the positional isomerism and structural flexibility introduced by the methylene spacer. Ligand 3-MPTCA exhibited the largest HOMO-LUMO gap, indicating a higher kinetic stability and lower chemical reactivity compared to ligands 3-PTCA, 4-PTCA, and 4M-PTCA. Conversely, ligand 3-PTCA displayed the narrowest HOMO-LUMO gap, suggesting a greater chemical reactivity and potential ease in forming coordination bonds.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms26115123/s1. CCDC 2317645-2317648 contains the supplementary crystallographic data of compounds 3-PTCA, 4-PTCA, 3-MPTCA, and 4-MPTCA, respectively. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_request/cif, accessed on 21 May 2025.

Author Contributions

Conceptualization, Methodology, Software, Validation: I.B., G.E.D., J.C., P.N., N.N., R.P., Y.L. and J.L.; Methodology, Validation, Writing—Reviewing and Editing: F.Z., P.A.-O. and I.B.; Methodology, Formal Analyses, Writing: I.B. and J.C.; Conceptualization, Writing—Reviewing and Editing, Software, Validation: G.E.D. and P.N.; Methodology, Writing—Reviewing and Editing: I.B. and G.E.D.; Draft preparation, Writing—Reviewing and Editing, Writing—Original Draft Preparation: N.N., R.P., I.B. and P.N.; Investigation, Methodological, Formal Analysis: G.E.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article and Supplementary Materials.

Acknowledgments

The authors thank ANID for the financial support from “Fondo Nacional de Desarrollo Científico y Tecnológico [FONDECYT (Chile), grant n°. 1210689 and 1170256 and FONDEQUIP EQM, grant n° 130021, 180024, and 210029] (I. Brito). This work was partially carried out by Iván Brito Bobadilla during a visit to the Universidad Autónoma de Madrid, supported by the MINEDUC-UA project, code: ANT 22991. This work has been supported by MCINN/AEI/10.13039/5011000011033 under the National Program of Sciences and Technological Materials [ Grant n° PID2022–138968NB-C21 and TED2021–131132B–C22] (P.Amo-Ocho; F.Zamora). The authors also acknowledge Vicerrectoría de Investigación, Innovación y Postgrado (VRIIP-UA) and Dirección de Gestión de la investigación (DGI-UA) of the Universidad de Antofagasta, Chile for the financial support [grant SEM-18]. P. Narea thanks ANID (Chile) for their support of a graduate fellowship (n°. 21190030). This work was partially related to G.E. Delgado’s visit at the Universidad de Antofagasta, supported by MINEDUC-UA project, code ANT 1856.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Kanupriya; Mittal, R.K.; Sharma, V.; Biswas, T.; Mishra, I. Recent Advances in Nitrogen-Containing Heterocyclic Scaffolds as Antiviral Agents. Med. Chem. 2024, 20, 487–502. [Google Scholar] [CrossRef]
  2. Egbujor, M.C.; Tucci, P.; Onyeije, U.C.; Emeruwa, C.N.; Saso, L. NRF2 Activation by Nitrogen Heterocycles: A Review. Molecules 2023, 28, 2751. [Google Scholar] [CrossRef]
  3. Juríček, M.; Kouwer, P.H.J.; Rowan, A.E. Triazole: A Unique Building Block for the Construction of Functional Materials. Chem. Commun. 2011, 47, 8740. [Google Scholar] [CrossRef]
  4. Fantoni, N.Z.; El-Sagheer, A.H.; Brown, T. A Hitchhiker’s Guide to Click-Chemistry with Nucleic Acids. Chem. Rev. 2021, 121, 7122–7154. [Google Scholar] [CrossRef]
  5. Jankovič, D.; Virant, M.; Gazvoda, M. Copper-Catalyzed Azide–Alkyne Cycloaddition of Hydrazoic Acid Formed In Situ from Sodium Azide Affords 4-Monosubstituted-1,2,3-Triazoles. J. Org. Chem. 2022, 87, 4018–4028. [Google Scholar] [CrossRef]
  6. Esposito, M.R.; Newar, R.; Kim, M.; Cohen, S.M. Incorporation of Low-Valent Metal Ions in Metal-Organic Frameworks and Coordination Polymers. Coord. Chem. Rev. 2025, 531, 216491. [Google Scholar] [CrossRef]
  7. Naik, A.D.; Dîrtu, M.M.; Railliet, A.P.; Marchand-Brynaert, J.; Garcia, Y. Coordination Polymers and Metal Organic Frameworks Derived from 1,2,4-Triazole Amino Acid Linkers. Polymers 2011, 3, 1750–1775. [Google Scholar] [CrossRef]
  8. Narea, P.; Cisterna, J.; Cárdenas, A.; Amo-Ochoa, P.; Zamora, F.; Climent, C.; Alemany, P.; Conejeros, S.; Llanos, J.; Brito, I. Crystallization Induced Enhanced Emission in Two New Zn(II) and Cd(II) Supramolecular Coordination Complexes with the 1-(3,4-Dimethylphenyl)-5-Methyl-1H-1,2,3-Triazole-4-Carboxylate Ligand. Polymers 2020, 12, 1756. [Google Scholar] [CrossRef]
  9. Bezerra Morais, P.A.; Javarini, C.L.; Valim, T.C.; Francisco, C.S.; de Freitas Ferreira, L.C.; Trancoso Bottocim, R.R.; Neto, Á.C.; Júnior, V.L. Triazole: A New Perspective in Medicinal Chemistry and Material Science. Curr. Org. Chem. 2022, 26, 1691–1702. [Google Scholar] [CrossRef]
  10. Kantheti, S.; Narayan, R.; Raju, K.V.S.N. The Impact of 1,2,3-Triazoles in the Design of Functional Coatings. RSC Adv. 2015, 5, 3687–3708. [Google Scholar] [CrossRef]
  11. Sehrawat, R.; Vashishth, P.; Bairagi, H.; Shukla, S.K.; Kumar, H.; Ji, G.; Mangla, B. Coordination Bonding and Corrosion Inhibition Characteristics of Chalcone Compounds for Metals: An Inclusive Review Based on Experimental as Well as Theoretical Perspectives. Coord. Chem. Rev. 2024, 514, 215820. [Google Scholar] [CrossRef]
  12. Sun, J.-K.; Zhang, J. Functional Metal–Bipyridinium Frameworks: Self-Assembly and Applications. Dalton Trans. 2015, 44, 19041–19055. [Google Scholar] [CrossRef]
  13. Huang, Y.; Cohen, T.A.; Sperry, B.M.; Larson, H.; Nguyen, H.A.; Homer, M.K.; Dou, F.Y.; Jacoby, L.M.; Cossairt, B.M.; Gamelin, D.R.; et al. Organic Building Blocks at Inorganic Nanomaterial Interfaces. Mater. Horiz. 2022, 9, 61–87. [Google Scholar] [CrossRef]
  14. Cameron, J.M.; Guillemot, G.; Galambos, T.; Amin, S.S.; Hampson, E.; Mall Haidaraly, K.; Newton, G.N.; Izzet, G. Supramolecular Assemblies of Organo-Functionalised Hybrid Polyoxometalates: From Functional Building Blocks to Hierarchical Nanomaterials. Chem. Soc. Rev. 2022, 51, 293–328. [Google Scholar] [CrossRef]
  15. Liu, Y.; Wang, L.; Zhao, L.; Zhang, Y.; Li, Z.-T.; Huang, F. Multiple Hydrogen Bonding Driven Supramolecular Architectures and Their Biomedical Applications. Chem. Soc. Rev. 2024, 53, 1592–1623. [Google Scholar] [CrossRef]
  16. Rehm, T.H.; Schmuck, C. Ion-Pair Induced Self-Assembly in Aqueous Solvents. Chem. Soc. Rev. 2010, 39, 3597. [Google Scholar] [CrossRef]
  17. Miao, Z.; Zhou, J. Multiscale Modeling and Simulation of Zwitterionic Anti-Fouling Materials. Langmuir 2025, 41, 7980–7995. [Google Scholar] [CrossRef]
  18. Anthony, A.; Desiraju, G.R.; Jetti, R.K.R.; Kuduva, S.S.; Madhavi, N.N.L.; Nangia, A.; Thaimattam, R.; Thalladi, V.R. Crystal Engineering: Some Further Strategies. Cryst. Eng. 1998, 1, 1–18. [Google Scholar] [CrossRef]
  19. Disa, A.S.; Nova, T.F.; Cavalleri, A. Engineering Crystal Structures with Light. Nat. Phys. 2021, 17, 1087–1092. [Google Scholar] [CrossRef]
  20. Zhang, C.; Yuan, L.; Liu, C.; Li, Z.; Zou, Y.; Zhang, X.; Zhang, Y.; Zhang, Z.; Wei, G.; Yu, C. Crystal Engineering Enables Cobalt-Based Metal–Organic Frameworks as High-Performance Electrocatalysts for H 2 O 2 Production. J. Am. Chem. Soc. 2023, 145, 7791–7799. [Google Scholar] [CrossRef]
  21. Yuan, Y.; Patel, R.K.; Banik, S.; Reta, T.B.; Bisht, R.S.; Fong, D.D.; Sankaranarayanan, S.K.R.S.; Ramanathan, S. Proton Conducting Neuromorphic Materials and Devices. Chem. Rev. 2024, 124, 9733–9784. [Google Scholar] [CrossRef]
  22. Mukherjee, D.; Saha, A.; Moni, S.; Volkmer, D.; Das, M.C. Anhydrous Solid-State Proton Conduction in Crystalline MOFs, COFs, HOFs, and POMs. J. Am. Chem. Soc. 2025, 147, 5515–5553. [Google Scholar] [CrossRef]
  23. Yang, Z.; Zhang, N.; Lei, L.; Yu, C.; Ding, J.; Li, P.; Chen, J.; Li, M.; Ling, S.; Zhuang, X.; et al. Supramolecular Proton Conductors Self-Assembled by Organic Cages. JACS Au 2022, 2, 819–826. [Google Scholar] [CrossRef]
  24. Shi, C.-Y.; Qin, W.-Y.; Qu, D.-H. Semi-Crystalline Polymers with Supramolecular Synergistic Interactions: From Mechanical Toughening to Dynamic Smart Materials. Chem. Sci. 2024, 15, 8295–8310. [Google Scholar] [CrossRef]
  25. Lugger, S.J.D.; Houben, S.J.A.; Foelen, Y.; Debije, M.G.; Schenning, A.P.H.J.; Mulder, D.J. Hydrogen-Bonded Supramolecular Liquid Crystal Polymers: Smart Materials with Stimuli-Responsive, Self-Healing, and Recyclable Properties. Chem. Rev. 2022, 122, 4946–4975. [Google Scholar] [CrossRef]
  26. Ying, Y.P.; Kamarudin, S.K.; Masdar, M.S. Silica-Related Membranes in Fuel Cell Applications: An Overview. Int. J. Hydrogen Energy 2018, 43, 16068–16084. [Google Scholar] [CrossRef]
  27. Chu, D.; Huang, C.-C.; Tsen, W.-C.; Gong, C. Biomass Composite Proton Exchange Membranes with High Mechanical Strength and Proton Conductivity Based on Chitosan and Zwitterion-Modified Palygorskite. Appl. Clay Sci. 2024, 253, 107367. [Google Scholar] [CrossRef]
  28. Chen, L.; Ren, Y.; Fan, F.; Wu, T.; Wang, Z.; Zhang, Y.; Zhao, J.; Cui, G. Artificial Frameworks towards Ion-Channel Construction in Proton Exchange Membranes. J. Power Sources 2023, 574, 233081. [Google Scholar] [CrossRef]
  29. Li, Q.; Wen, C.; Yang, J.; Zhou, X.; Zhu, Y.; Zheng, J.; Cheng, G.; Bai, J.; Xu, T.; Ji, J.; et al. Zwitterionic Biomaterials. Chem. Rev. 2022, 122, 17073–17154. [Google Scholar] [CrossRef]
  30. Qu, K.; Yuan, Z.; Wang, Y.; Song, Z.; Gong, X.; Zhao, Y.; Mu, Q.; Zhan, Q.; Xu, W.; Wang, L. Structures, Properties, and Applications of Zwitterionic Polymers. ChemPhysMater 2022, 1, 294–309. [Google Scholar] [CrossRef]
  31. He, G.; Li, Z.; Li, Y.; Li, Z.; Wu, H.; Yang, X.; Jiang, Z. Zwitterionic Microcapsules as Water Reservoirs and Proton Carriers within a Nafion Membrane To Confer High Proton Conductivity under Low Humidity. ACS Appl. Mater. Interfaces 2014, 6, 5362–5366. [Google Scholar] [CrossRef] [PubMed]
  32. Doobi, F.A.; Mir, F.Q. Exploring the Development of Natural Biopolymer (Chitosan)-Based Proton Exchange Membranes for Fuel Cells: A Review. Results Surf. Interfaces 2024, 15, 100218. [Google Scholar] [CrossRef]
  33. Jiao, K.; Xuan, J.; Du, Q.; Bao, Z.; Xie, B.; Wang, B.; Zhao, Y.; Fan, L.; Wang, H.; Hou, Z.; et al. Designing the next Generation of Proton-Exchange Membrane Fuel Cells. Nature 2021, 595, 361–369. [Google Scholar] [CrossRef] [PubMed]
  34. Nagao, Y. Proton-Conducting Polymers: Key to Next-Generation Fuel Cells, Electrolyzers, Batteries, Actuators, and Sensors. ChemElectroChem 2024, 11, e202300846. [Google Scholar] [CrossRef]
  35. Narea, P.; Hernández, B.; Cisterna, J.; Cárdenas, A.; Llanos, J.; Amo-Ochoa, P.; Zamora, F.; Priego, J.L.; Cortijo, M.; Delgado, G.E.; et al. Heterobimetallic Three-Dimensional 4d-4f Coordination Polymers Based on 5-Methyl-1-(Pyridyn-4-Ylmethyl)-1H-1,2,3-Triazole-3,4-Dicarboxylate. J. Solid. State Chem. 2022, 310, 123027. [Google Scholar] [CrossRef]
  36. Hernández, B.; Narea, P.; Espinoza, D.; Navarrete, A.; Aguirre, G.; Delgado, G.E.; Cárdenas, A.; Brito, I.; Cisterna, J. Novel Zn(II) and Cd(II) Coordination Polymers Derived from 1,2,3-Triazole-1,3-Diketone Ligand. Syntheses and Structural, Thermal, Computational, and Luminescent Studies. J. Solid. State Chem. 2022, 312, 123156. [Google Scholar] [CrossRef]
  37. Kamalraj, V.R.; Senthil, S.; Kannan, P. One-Pot Synthesis and the Fluorescent Behavior of 4-Acetyl-5-Methyl-1,2,3-Triazole Regioisomers. J. Mol. Struct. 2008, 892, 210–215. [Google Scholar] [CrossRef]
  38. Cisterna, J.; Araneda, C.; Narea, P.; Cárdenas, A.; Llanos, J.; Brito, I. The Positional Isomeric Effect on the Structural Diversity of Cd(II) Coordination Polymers, Using Flexible Positional Isomeric Ligands Containing Pyridyl, Triazole, and Carboxylate Fragments. Molecules 2018, 23, 2634. [Google Scholar] [CrossRef]
  39. Allen, F.H.; Kennard, O.; Watson, D.G.; Brammer, L.; Orpen, A.G.; Taylor, R. Tables of Bond Lengths Determined by X-Ray and Neutron-Diffraction. 1. Bond Lengths in Organic-Compounds. J. Chem. Soc. Perkin Trans. 1987, 2, S1–S19. [Google Scholar] [CrossRef]
  40. Groom, C.R.; Allen, F.H. The Cambridge Structural Database in Retrospect and Prospect. Angew. Chem. Int. Ed. 2014, 53, 662–671. [Google Scholar] [CrossRef]
  41. Etter, M.C.; MacDonald, J.C.; Bernstein, J. Graph-Set Analysis of Hydrogen-Bond Patterns in Organic Crystals. Acta Crystallogr. B 1990, 46, 256–262. [Google Scholar] [CrossRef] [PubMed]
  42. Fábián, L.; Kálmán, A. Volumetric Measure of Isostructurality. Acta Crystallogr. B 1999, 55, 1099–1108. [Google Scholar] [CrossRef]
  43. Wolff, S.K.; Grimwood, D.J.; McKinnon, J.J.; Turner, M.J.; Jayatilaka, D.; Spackman, M.A. Crystal Explorer, version 17.5; Crystal Explorer 17.5; The University of Western Australia: Perth, Australia, 2012. [Google Scholar]
  44. Spackman, M.A.; Jayatilaka, D. Hirshfeld Surface Analysis. CrystEngComm 2009, 11, 19–32. [Google Scholar] [CrossRef]
  45. Spackman, M.A.; McKinnon, J.J. Fingerprinting Intermolecular Interactions in Molecular Crystals. CrystEngComm 2002, 4, 378–392. [Google Scholar] [CrossRef]
  46. Mackenzie, C.F.; Spackman, P.R.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer Model Energies and Energy Frameworks: Extension to Metal Coordination Compounds, Organic Salts, Solvates and Open-Shell Systems. IUCrJ 2017, 4, 575–587. [Google Scholar] [CrossRef] [PubMed]
  47. Bruker AXS INC. APEX4 Package. APEX4, SAINT and SADABS; Bruker AXS INC: Madison, WI, USA, 2020. [Google Scholar]
  48. Sheldrick, G.M. SADABS, Software for Empirical Absorption Correction; University of Göttingen: Göttingen, Germany, 2000. [Google Scholar]
  49. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  50. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.K.; Puschmann, H. OLEX2: A Complete Structure Solution, Refinement and Analysis Program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  51. Spek, A.L. Single-Crystal Structure Validation with the Program. PLATON J. Appl. Crystallogr. 2003, 36, 7–13. [Google Scholar] [CrossRef]
  52. Spackman, P.R.; Turner, M.J.; McKinnon, J.J.; Wolff, S.K.; Grimwood, D.J.; Jayatilaka, D.; Spackman, M.A. CrystalExplorer: A Program for Hirshfeld Surface Analysis, Visualization and Quantitative Analysis of Molecular Crystals. J. Appl. Crystallogr. 2021, 54, 1006–1011. [Google Scholar] [CrossRef]
  53. Bondi, A. Van Der Waals Volumes and Radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  54. Batsanov, S.S. Van Der Waals Radii of Elements. Inorg. Mater. Transl. Neorg. Mater. Orig. Russ. Text. 2001, 37, 871–885. [Google Scholar] [CrossRef]
  55. Spackman, M.A.; McKinnon, J.J.; Jayatilaka, D. Electrostatic Potentials Mapped on Hirshfeld Surfaces Provide Direct Insight into Intermolecular Interactions in Crystals. CrystEngComm 2008, 10, 377–388. [Google Scholar] [CrossRef]
  56. Jayatilaka, D.; Grimwood, D.J. Tonto: A Fortran Based Object-Oriented System for Quantum Chemistry and Crystallography. In Proceedings of the Computational Science—ICCS 2003, Melbourne, Australia, 2–4 June 2003; pp. 142–151. [Google Scholar]
  57. Becke, A.D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  58. Tirado-Rives, J.; Jorgensen, W.L. Performance of B3LYP Density Functional Methods for a Large Set of Organic Molecules. J. Chem. Theory Comput. 2008, 4, 297–306. [Google Scholar] [CrossRef]
  59. Blatov, V.A.; Shevchenko, A.P.; Proserpio, D.M. Applied Topological Analysis of Crystal Structures with the Program Package ToposPro. Cryst. Growth Des. 2014, 14, 3576–3586. [Google Scholar] [CrossRef]
  60. O’Keeffe, M.; Peskov, M.A.; Ramsden, S.J.; Yaghi, O.M. The Reticular Chemistry Structure Resource (RCSR) Database of, and Symbols for, Crystal Nets. Acc. Chem. Res. 2008, 41, 1782–1789. [Google Scholar] [CrossRef]
  61. Alexandrov, E.V.; Blatov, V.A.; Kochetkov, A.V.; Proserpio, D.M. Underlying Nets in Three-Periodic Coordination Polymers: Topology, Taxonomy and Prediction from a Computer-Aided Analysis of the Cambridge Structural Database. CrystEngComm 2011, 13, 3947. [Google Scholar] [CrossRef]
  62. Shevchenko, A.P.; Blatov, V.A. Simplify to Understand: How to Elucidate Crystal Structures? Struct. Chem. 2021, 32, 507–519. [Google Scholar] [CrossRef]
  63. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA Quantum Chemistry Program Package. J. Chem. Phys. 2020, 152, 224108. [Google Scholar] [CrossRef]
  64. Neese, F. Software Update: The ORCA Program System—Version 5.0. WIREs Comput. Mol. Sci. 2022, 12, e1606. [Google Scholar] [CrossRef]
Scheme 1. General synthetical route of zwitterionic positional isomers.
Scheme 1. General synthetical route of zwitterionic positional isomers.
Ijms 26 05123 sch001
Figure 1. The asymmetric unit for all compounds with an anisotropic ellipsoid representation, together with an atom labeling scheme. The ellipsoids are drawn at a 30% probability level. Hydrogen atoms are depicted as spheres with arbitrary radii.
Figure 1. The asymmetric unit for all compounds with an anisotropic ellipsoid representation, together with an atom labeling scheme. The ellipsoids are drawn at a 30% probability level. Hydrogen atoms are depicted as spheres with arbitrary radii.
Ijms 26 05123 g001
Figure 2. The view shows the chain and ring with the graph-set motifs  C (9) (b) and R 6 5 (44) (a) as representative examples of the supramolecular structure of 3- and 4-PTCA.
Figure 2. The view shows the chain and ring with the graph-set motifs  C (9) (b) and R 6 5 (44) (a) as representative examples of the supramolecular structure of 3- and 4-PTCA.
Ijms 26 05123 g002
Figure 3. The view shows the chain and ring with the graph-set motifs C(8) (b) and R 6 6 (52) (a) as a representative example of the supramolecular structure of 3-MPTCA.
Figure 3. The view shows the chain and ring with the graph-set motifs C(8) (b) and R 6 6 (52) (a) as a representative example of the supramolecular structure of 3-MPTCA.
Ijms 26 05123 g003
Figure 4. The view shows chains and rings with the graph-set notation C(8) (a) and R 2 2 (22) (c) as a representative example of the supramolecular structure of 4-MPTCA.
Figure 4. The view shows chains and rings with the graph-set notation C(8) (a) and R 2 2 (22) (c) as a representative example of the supramolecular structure of 4-MPTCA.
Ijms 26 05123 g004
Figure 5. A view of the three-dimensional Hirshfeld surface of all compounds plotted over dnorm. Intermolecular contacts longer than the sum of the van der Waals radii are colored blue, and contacts shorter than the sum of the radii are colored red. Contacts equal to the sum of the radii are white.
Figure 5. A view of the three-dimensional Hirshfeld surface of all compounds plotted over dnorm. Intermolecular contacts longer than the sum of the van der Waals radii are colored blue, and contacts shorter than the sum of the radii are colored red. Contacts equal to the sum of the radii are white.
Ijms 26 05123 g005
Figure 6. The resulting topologies in all compounds.
Figure 6. The resulting topologies in all compounds.
Ijms 26 05123 g006
Figure 7. Visualizations of the HOMO and LUMO molecular orbitals for the synthesized ligands. In the isosurfaces, the red region indicates a high electron concentration, whereas the blue region represents a low electron concentration.
Figure 7. Visualizations of the HOMO and LUMO molecular orbitals for the synthesized ligands. In the isosurfaces, the red region indicates a high electron concentration, whereas the blue region represents a low electron concentration.
Ijms 26 05123 g007
Table 1. Crystal data for all compounds.
Table 1. Crystal data for all compounds.
3-PTCA4-PTCA
CCDC number23176472317648
Empirical formulaC10H8N4O4C10H8N4O4
Formula weight248.20248.20
Temperature/K297297
Crystal systemorthorhombicorthorhombic
Space groupPna21 (No 33)Pna21 (No 33)
a/Å12.0311(14)13.962(5)
b/Å17.673(2)17.017(6)
c/Å5.0309(4)4.4663(17)
Volume/Å31069.7(2)1061.2(7)
Z’11
Z44
ρcalcg/cm31.5411.554
μ/mm−11.0520.124
F (000)512512
Crystal size/mm30.11 × 0.21 × 0.300.15 × 0.20 × 0.38
RadiationCuKα (λ = 1.54178)MoKα (λ = 0.71073)
2Θ range for data collection/°4.4 to 59.22.8 to 26.5
Index ranges−10 ≤ h ≤ 13, −17 ≤ k ≤ 19, −5 ≤ l ≤ 5−17 ≤ h ≤ 17, −20 ≤ k ≤ 21, −5 ≤ l ≤ 5
Reflections collected11,7757489
Independent reflections1462 [Rint = 0.177]1916 [Rint = 0.047]
Data/restraints/parameters1462/0/1661916/0/165
Goodness-of-fit on F21.001.30
Flack parameter (x)0.6(1)0(3)
Final R indexes [all data]R1 = 0.0693, wR2 = 0.1904R1 = 0.0432, wR2 = 0.1719
Largest diff. peak/hole/e Å−30.25/-0.250.68/−0.65
3-MPTCA4-MPTCA
CCDC number23176452317646
Empirical formulaC11H10N4O4C11H10N4O4
Formula weight262.23262.23
Temperature/K297297
Crystal systemorthorhombicmonoclinic
Space groupP212121 (No 19)P21/c (No 14)
a/Å4.748(12)4.9510(7)
b/Å12.88(4)19.019(3)
c/Å18.02(5)11.9930(16)
β/°-93.526(7)
Volume/Å31102(5)1127.2(3)
Z’11
Z44
ρcalcg/cm31.5811.545
μ/mm−10.1241.030
F(000)544544
Crystal size/mm30.18 × 0.08 × 0.070.19 × 0.06 × 0.05
RadiationMoKα (λ = 0.71073)MoKα (λ = 0.71073)
2Θ range for data collection/°2.8 to 25.04.4 to 6.4
Index ranges−5 ≤ h ≤ 3, −15 ≤ k ≤ 15, −21 ≤ l ≤ 21−5 ≤ h ≤ 5, −22 ≤ k ≤ 19, −14 ≤ l ≤ 14
Reflections collected643913612
Independent reflections1902 [Rint = 0.211]1921 [Rint = 0.315]
Data/restraints/parameters1902/0/1741921/0/173
Goodness-of-fit on F20.971.07
Final R indexes [all data]R1 = 0.0712, wR2 = 0.1830R1 = 0.0554, wR2 = 0.1139
Largest diff. peak/hole/e Å−30.24/−0.290.33/−0.25
Table 2. Dihedral angles for all compounds (°).
Table 2. Dihedral angles for all compounds (°).
3-PTCA4-PTCA
C3-N1-C6-C10−44 (2)26.7 (8)
C3-C4-C5-O4−1.4 (1)26.7 (6)
C3-C2-C1-O1−7 (2)−2.0 (8)
3-MPTCA4-MPTCA
N1-C6-C7-C8 −66 (1)−39.8 (3)
C3-C2-C1-O1−4.0 (2)4.6 (3)
C3-C4-C5-O4−17 (1)−19.4 (3)
Table 3. Hydrogen bond interactions for all compounds.
Table 3. Hydrogen bond interactions for all compounds.
SampleD−H∙∙∙A *Symmetry Code D∙∙∙A (Å)D−H∙∙∙A (°)
3-PTCAO1−H1∙∙∙O3[1 − x, 1 − y, −1/2 + z]2.519 (12)173
N4−H4∙∙∙O3[−1/2 + x, 1/2 − y, 1 + z]2.565 (13)170
H1∙∙ O3-C5[1 − x,1 − y, 1/2 + z]2.519 (12)173
H4∙∙∙O3−C5[1/2 + x, 1/2 − y, −1 + z]2.565 (13)170
4-PTCAO1−H1∙∙∙O3[−x, −y, 1/2 + z]2.551 (5)162
N4−H4∙∙∙O3[1/2 + x, −1/2 − y,1 + z]2.640 (6)179
H1∙∙∙O3−C5[−x, −y, 1/2 + z]2.551 (5)163
H4∙∙∙O3−C5[−1/2 + x, −1/2 − y, −1 + z]2.640 (6)179
3-MPTCAO3−H3∙∙∙O1[−1/2 + x, 1/2 − y, 1 − z]2.434 (9)165
N4−H4∙∙∙O2[3/2 − x, 1 − y, −1/2 + z]2.628 (15)145
4-MPTCAO2−H2∙∙∙O3[1 + x, 1/2 − y, ½ − z]2.450 (3)133
N4−H4∙∙∙O1[1 − x, 1/2 + y, ½ − z]2.757 (3)142
(N)H∙∙∙O1−C1[1 − x, −1/2 + y,1/2 + z]2.757 (3)142
H2∙∙∙O3−C5[−1 + x, 1/2 − y, −1/2 + z]2.449 (3)133
* D = Donor and A = Acceptor.
Table 4. Visualizations of the HOMO and LUMO molecular orbitals for the synthesized ligands. All values are reported in eV.
Table 4. Visualizations of the HOMO and LUMO molecular orbitals for the synthesized ligands. All values are reported in eV.
Parameter3-PTCA4-PTCA3-MPTCA4-MPTCA
EHOMO−5.5285−5.6125−6.3038−5.7310
ELUMO−3.6468−3.5904−2.2618−2.6355
EGap 1.88172.02214.04203.0955
IP5.52855.61256.30385.7310
EA3.64683.59042.26182.6355
Hardness (η)0.94091.01112.02101.5477
Electronegativity (X)4.58774.60144.28284.1833
Chemical Potential (μ)−4.5877−4.6014−4.2828−4.1833
Electrophilicity (ω)11.184910.47104.53795.6532
Softness (σ)0.53140.49450.24740.3230
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

Delgado, G.E.; Cisterna, J.; Llanos, J.; Pulido, R.; Naveas, N.; Narea, P.; Amo-Ochoa, P.; Zamora, F.; León, Y.; Brito, I. Structure–Property Relationships in Zwitterionic Pyridinium–Triazole Ligands: Insights from Crystal Engineering and Hirshfeld Surface Analysis. Int. J. Mol. Sci. 2025, 26, 5123. https://doi.org/10.3390/ijms26115123

AMA Style

Delgado GE, Cisterna J, Llanos J, Pulido R, Naveas N, Narea P, Amo-Ochoa P, Zamora F, León Y, Brito I. Structure–Property Relationships in Zwitterionic Pyridinium–Triazole Ligands: Insights from Crystal Engineering and Hirshfeld Surface Analysis. International Journal of Molecular Sciences. 2025; 26(11):5123. https://doi.org/10.3390/ijms26115123

Chicago/Turabian Style

Delgado, Gerzon E., Jonathan Cisterna, Jaime Llanos, Ruth Pulido, Nelson Naveas, Pilar Narea, Pilar Amo-Ochoa, Félix Zamora, Yasna León, and Iván Brito. 2025. "Structure–Property Relationships in Zwitterionic Pyridinium–Triazole Ligands: Insights from Crystal Engineering and Hirshfeld Surface Analysis" International Journal of Molecular Sciences 26, no. 11: 5123. https://doi.org/10.3390/ijms26115123

APA Style

Delgado, G. E., Cisterna, J., Llanos, J., Pulido, R., Naveas, N., Narea, P., Amo-Ochoa, P., Zamora, F., León, Y., & Brito, I. (2025). Structure–Property Relationships in Zwitterionic Pyridinium–Triazole Ligands: Insights from Crystal Engineering and Hirshfeld Surface Analysis. International Journal of Molecular Sciences, 26(11), 5123. https://doi.org/10.3390/ijms26115123

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