Laura
Holzhauer
a,
Vitalii
Shekhovtsev
b,
Cecilia
Bruschi
c,
Mathis
Gunther
a,
Olaf
Fuhr
bd,
Patrick
Hodapp
e,
Claudia
Bizzarri
c,
Wolfgang
Wenzel
b,
Nicole
Jung
*ad and
Stefan
Bräse
*ac
aInstitute of Biological and Chemical Systems (IBCS-FMS), Karlsruhe Institute of Technology, Kaiserstraße 12, 76131 Karlsruhe, Germany. E-mail: nicole.jung@kit.edu; braese@kit.edu
bInstitute of Nanotechnology (INT), Karlsruhe Institute of Technology, Kaiserstraße 12, 76131 Karlsruhe, Germany
cInstitute of Organic Chemistry (IOC), Karlsruhe Institute of Technology, Kaiserstraße 12, 76131 Karlsruhe, Germany
dKarlsruhe Nano Micro Facility (KNMFi), Karlsruhe Institute of Technology, Kaiserstraße 12, 76131 Karlsruhe, Germany
eInstitute for Biological Interfaces 3 – Soft Matter Synthesis Laboratory (IBG3-SML), Karlsruhe Institute of Technology, Kaiserstraße 12, 76131 Karlsruhe, Germany
First published on 8th January 2025
The formation of novel complexes from so far non-investigated ligands and different metal centers is important for the development of new functional materials such as (photo)catalysts or biologically active compounds. Still, promising strategies to quickly and systematically investigate the complexation behavior of selected ligands are rare. We developed an NMR-based screening approach to monitor changes within reaction mixtures containing metals and ligands on a small scale via a simple but reliable protocol. Based on the obtained data, we could draw conclusions on the formation of 2-(1′,2′,3′-triazol-1′yl)quinoxaline-based transition metal complexes. The NMR screening results obtained were confirmed by the repetition of selected experiments on a large scale, and wherever possible, the type of complex obtained was evaluated by crystal structure elucidation. We could show the versatile complexation of the selected ligands with copper, gold, and silver in a bridging, chelating, or monodentate manner. The calculated theoretical properties supported the experimental results and enabled initial in silico predictions of similar compounds before conducting the complexation experiments and structural analyses.
Metal complexes incorporating N-heterocyclic compounds with 1,2,3 triazoles are known as catalysts,7,8 photosensitizers,9 and can be applied as live cell imaging probes10 and for tumor diagnosis11 amongst other applications.12,13 Thus, we aimed to gain further insights into the coordination chemistry of triazolylquinoxalines to facilitate the access to novel complexes and to possible new applications of so far unknown metal complexes. Two isomers of triazole-linked N-heterocycles can be distinguished; either the nitrogen or the carbon atoms of the triazole are bound to the quinoxaline. The less common nitrogen-bound form undergoes chelating complexation via the nitrogen in position 2 instead of position 1 or 3 as for the carbon-bound form. Therefore, metal complexes with these ligands show differences in their catalytical and photochemical activity, as well as their luminescence intensity, Stokes shift, and lifetime compared to their carbon-bound counterparts.7,14,15 Some of the rare examples of metal complexes with triazole-linked N-heterocycles bound via the nitrogen atom are triazolylquinoxaline-based rhenium complexes. These complexes have been previously described by Bertrand et al.15 and have further been explored by our group, employing pyrazine-substituted triazolylquinoxalines and triazolylimidazo-quinoxalines as ligands.16 The respective stable rhenium complexes are obtained by reaction of Re(CO)5Br with the desired ligand in toluene at 110 °C.
Encouraged by these results, we selected the triazolylquinoxaline-based ligands 5a–e to screen their complexation properties with several other metal centers. These compounds can be straightforwardly synthesized and possess a variation in either the substitution of the triazole or the pyrazine ring of the quinoxaline, allowing to investigate the influence of the molecular structure on the complexation properties of triazolylquinoxalines. The ligands were accessed via copper-catalyzed azide–alkyne cycloaddition (CuAAC) from the readily available tetrazolo[1,5-a]quinoxaline 2 as previously described in the literature (see Scheme 1).16,17
![]() | ||
Scheme 1 Synthesis of triazole-bearing ligands 5a–e for the complexation with different metal centers.15,16 The numbering of the triazole atom is indicated in blue. |
It was intended to investigate the complexation of the selected ligands with different metals by NMR spectroscopy, as this technique is commonly used and readily available in most chemical institutes. 1H NMR measurements present a rapid methodology to observe complexation due to the expected changes of the shifts assigned to the ligand when coordinating to a diamagnetic metal. For soluble complexes, down- or upfield shifting of selected signals – especially of hydrogen atoms close to the coordinating nitrogen center – is expected. In those cases where the obtained metal complex is less soluble, NMR measurements should reveal a strong depletion up to the complete disappearance of the signals of the ligands. Therefore, an NMR screening combines beneficial aspects for monitoring the complexation of insoluble and soluble complexes. The process allows the indirect detection of insoluble complexes and monitors the progress of the complexation where missing.
To exploit the potential of an NMR-based screening method for a systematic investigation of the coordination potential of different ligands to different metals, processes for homo- and heteroleptic complexes were defined. For a homoleptic screening with 12 different metal salts 6–17 (Table 1), stock solutions of the ligands 5a–c and an appropriate amount of metal salt were added to an NMR tube. Deuterated methanol was selected as a solvent due to its suitability for comparable experiments with benzotriazole as a ligand.18 The tubes were then shaken for 30 minutes to ensure thorough mixing of the metal precursor and the ligand for the reaction. Subsequently, NMR spectra of these samples were recorded. Successful complexation led to either a shift of the signal or a decrease of the ligand signals and their disappearance, indicating full conversion to one or more insoluble complexes. AgNO3, AgSO3CF3, AgCO2CF3, and RuCl3 promptly complexed with all three ligands, causing a shift for the aromatic signals and, most conspicuously, for the triazole hydrogen atom (8.7 ppm for ligand 5a, see Fig. 1). For AuBr3, the ligand signals disappeared almost completely due to the precipitation of the obtained complex. The NMR spectra of the successful experiments with ligand 5a are depicted in Fig. 2; the spectra for ligands 5b and 5c as well as for the remaining metals are given in the ESI (Fig. S2–S9†).
Metal ligand | AgNO3 (6) | Ag2CO3 (7) | AgSO3CF3 (8) | AgOCOCH3 (9) | AgOCOCF3 (10) | AuBr3 (11) | PdCl2 (12) | RuCl3 (13) | IrCl3 (14) | EuCl3 (15) | La(NO3)3 (16) | ZnCl2 (17) |
---|---|---|---|---|---|---|---|---|---|---|---|---|
5a | ✓ | x | ✓ | x | ✓ | ✓ | ✓* | ✓ | x | x | x | x |
5b | ✓ | x | ✓ | x* | ✓ | ✓ | ✓* | ✓ | x* | x | x | x |
5c | ✓ | x | ✓ | x | ✓ | ✓ | ✓* | ✓* | x | x | x | x |
When no complexation could be observed in the first measurement at ambient temperature, the samples were subjected to a follow-up experiment. The samples were transferred to a suitable reaction vial, heated to 65 °C for 4 h, and the 1H NMR measurements were repeated. After heating, a complexation with precipitation was observed for the samples with PdCl2 (see Fig. 2), but no change in 1H NMR was observed for the remaining metal centers. In the latter cases, the complexation was assumed to be unsuccessful. The shifts in the NMR spectra between different ligands and complexes with the same metal precursors are similar in all cases, leading to the assumption that the same coordination mode is present in all investigated complexes with the same metal precursor.
After the NMR-based screening experiments, the complexation with selected metals was repeated in batch synthesis to confirm the formation of the complexes and the expected type of coordination. Ligand 5a was employed for these experiments and was combined with all metal centers that showed successful complexation, according to Table 1. Using AgNO3 (6) as a precursor, a colorless solid precipitated and could be filtered from the reaction mixture to yield complex 18a in 72% yield. Similar to the NMR experiment, the bulk of the complex precipitated but was still soluble enough to be detected by 1HNMR. We were not able to crystallize the formed complex 18a from MeOH or MeOD. Suitable crystals were grown from a solution in DMSO-d6 allowing the elucidation of the product structure 18bvia single-crystal X-ray crystallography. The obtained metallacycle consists of two triazolylquinoxaline ligands, two silver atoms, and four nitrate ligands (see Scheme 2). The N3-atom of the triazole and the pyrazine nitrogen each coordinate to a silver atom, linking the dimer in an almost planar configuration. The coordination of the triazole through the N3 donor to the Ag(I) is consistent with related literature-known complexes.19 The complex crystallizes in the monoclinic space group P21/n with N–Ag bond lengths of 2.309 and 2.310 Å, only slightly longer than for reported dimeric benzotriazole analogues.18 A comparison of 18a and 18bvia powder X-ray diffractometry (PXRD) reveals that both complexes are monoclinic but cannot be considered as isostructural.
Using AgSO3CF3 (8) with the same ligand led to the formation of a coordination polymer (19) instead of the previously described dimer. A suitable single crystal was obtained from a diluted solution of the obtained precipitate in MeOH. The oxygen atoms of the nitrate counterion are replaced by further complexation to the next quinoxaline unit (see Scheme 3), creating a one-dimensional polymer where the silver atom is coordinated to four nitrogen atoms originating from the pyrazine and the N3-atom of the triazole. The triflate anion is incorporated as a counterion into the structure without direct bonding to the silver atom. The replacement of the counterion might be due to the more electron-deficient nature of the triflate oxygen compared to those in the nitrate counterion, causing the coordination of the quinoxaline to be favored above that of the oxygen. With N–Ag bond lengths of 2.2532 Å to 2.4168 Å, the four bonds are slightly distorted but still in the same range as the related dimeric complex. The chemical shifts of the complexation products of ligand 5a with AgNO3 and AgSO3CF3 differ slightly. Similar to the experiments of ligand 5a with AgNO3 and AgSO3CF3, the reaction of 5a with AgOCOCF3 gave a precipitate, indicating the formation of a complex. Nevertheless, no suitable single crystal could be obtained in the batch experiment. A similar coordination mode as for the silver complexes 18b and 19 is expected; the obtained mass data and the similar chemical shifts in the 1H NMR (see Fig. S10 and Table S1 in the ESI†) support this evaluation.
![]() | ||
Scheme 3 Synthesis of the polymeric complex 19 and ORTEP diagram of the complex with the thermal ellipsoids shown at 50% probability. |
Complexation attempts with AuBr3 (11) and ligand 5a were carried out in methanol at 25 °C. As already observed in the NMR screening, the complex 20 precipitated quickly as a red solid and could be obtained in 78% by filtration from the reaction mixture. A single crystal suitable for X-ray diffraction was obtained from the slow evaporation of a diluted methanol solution under ambient conditions. The complexation of the square planar Au(III) occurs via the N3-atom of the triazole without the involvement of the pyrazine nitrogen atoms, whereas all three bromine atoms remain bound to the gold (Scheme 4). The complex crystallizes in the monoclinic P21/c space group with an Au–N distance of 2.040 Å. The observed complexation mode and the distance between the nitrogen and metal center are analogous to the results obtained for a CH2-spaced complex reported in the literature.20 For PdCl2 and RuCl3, no suitable single crystals could be grown from the precipitated reaction mixture in the batch experiment. Due to the low solubility of the palladium complex, no conclusion regarding the complexation mode could be drawn from the NMR data. Nevertheless, the obtained mass data indicates the formation of a dimeric complex with two involved triazolylquinoxaline ligands similar to literature-reported benzotriazole palladium complexes18,21 and other triazole-containing complexes.22–24 For the ruthenium complex, a trimeric complex with three ligand molecules and one central ruthenium atom is assumed to be present in analogy to the literature.25,26 The assumed complexation through the triazole nitrogen is supported by the downfield shift of the proton linked to the triazole in the 1H NMR spectrum, similar to the other complexes (see also Fig. S11 and S16 in the ESI†).
The approach described for homoleptic complexes was expanded to heteroleptic complexes. For the investigation of triazolylquinoxaline-based copper–phosphine complexes, it was decided to additionally include ligands 5d–e with substituted aromatic residues on the triazole as we deemed those to be the most promising examples for a potential application in photocatalysis in analogy to previously published examples.9,27 Stock solutions of bis[(2-diphenylphosphino)phenyl] ether (DPEPhos, 23) or 4,5-bis(diphenylphosphino)-9,9-dimethylxanthene (XantPhos, 24) were mixed in an NMR tube with either [Ag(MeCN)4]BF4 or [Cu(MeCN)4]BF4 as precursors. Here, DCM-d2 was employed, following previous reports on heteroleptic Cu(I) and Ag(I) complexes.27,28 The solutions were shaken for 20–30 min, then a solution of the triazolylquinoxaline ligand was added, and the mixture was shaken for another 20–30 min before the 1H NMR measurement was conducted. A successful complexation of the triazolylquinoxaline ligand was observed for all the combinations (see Table 2 and Fig. 3), in accordance with literature describing similar complexes.29 Again indicated by the shift of the aromatic signals of the quinoxaline and, most remarkably, the triazole hydrogen atom. The spectra for experiments with ligands 5a and 5b–e are respectively given in Fig. 3 and in the ESI section (Fig. S12–S15†).
Ligand | [Ag(MeCN)4]BF4 (21) | [Cu(MeCN)4]BF4 (22) | ||
---|---|---|---|---|
DPEPhos (23) | XantPhos (24) | DPEPhos (23) | XantPhos (24) | |
5a | ✓ | ✓ | ✓ | ✓ |
5b | ✓ | ✓ | ✓ | ✓ |
5c | ✓ | ✓ | ✓ | ✓ |
5d | ✓ | ✓ | ✓ | ✓ |
5e | ✓ | ✓ | ✓ | ✓ |
The heteroleptic triazolylquinoxaline-phosphine silver complex 25 was synthesized in batch and fully analyzed, and the structure could be confirmed via single-crystal X-ray diffraction (Scheme 5). A suitable crystal was obtained through vapor diffusion with DCM and diethyl ether. The X-ray analysis highlighted that compound 5a acted as monodentate ligand, being coordinated only through the N5-nitrogen of the triazole moiety.
![]() | ||
Scheme 5 Synthesis of the heteroleptic phosphine silver complex 25 and ORTEP diagram of the complex with the thermal ellipsoids shown at 50% probability. |
Three heteroleptic copper complexes 26a, 26b, 26c were synthesized following a previously published procedure,30 adding the diimine ligand to a dried dichloromethane solution of the Cu(I) precursor [Cu(MeCN)4]BF4 and the chelating DPEPhos ligand, under an argon atmosphere. After 4 hours of reaction, the solvent was removed under vacuum, and the three copper complexes were obtained as red powders. Lastly, the three mononuclear complexes formed crystals by slow diffusion of a cyclohexane layer into a concentrated dichloromethane solution, achieving high purity with reaction yields of 77% for 26a, 84% for 26b and 76% for 26c (Scheme 6). The obtained crystals were suitable for single-crystal X-ray crystallography. These copper complexes crystallize in the triclinic system with a P space group. The three compounds show similar distances between the metal center and the central nitrogen atom of the triazole unit (2.063 Å for 26a, 2.060 Å for 26b, 2.067 Å for 26c), while the distance between the copper and the nitrogen atom of the quinoxaline ring is slightly smaller for 26a (2.061 Å) than the ones of 26b (2.089 Å) and 26c (2.084 Å), that are almost identical. The distances between the copper and the coordinative P atoms of the chelating diphosphine are quite similar for all three compounds (26a: 2.256 Å, 26b: 2.220 Å, 26c: 2.213 Å for Cu–P1 and 26a: 2.210 Å, 26b: 2.275 Å, 26c: 2.286 Å for Cu–P2). Lastly, the three complexes show similar bite angles N1–Cu–N4 (26a: 78.93°, 26b: 77.74°, 26c: 77.6°) and P1–Cu–P2 (26a: 115.67°, 26b: 113.15°, 26c: 117.44°). These values are typical of a pseudo-tetrahedral geometry and match those of previously published copper complexes with related structures.9,31 The present chelating coordination mode is in contrast to the monodentate complexation observed for the heteroleptic phosphine silver complexes (see above).
When comparing the 1H NMR spectra of all complexes elucidated above, a complexation-dependent shift of the isolated hydrogen atoms of the pyrazine and triazole is visible, showing the power of NMR screening techniques to distinguish between coordination modes (see Fig. S16, ESI†). Depending on the involvement of the quinoxaline unit in the complex, the pyrazine hydrogen atom is de-shielded or more shielded than the ligand.
In addition to the experimental screenings and batch syntheses, theoretical calculations with density functional theory (DFT) were conducted to analyze the screening results further. The computations were carried out with the hybrid range separated CAM-B3LYP functional32 combined with the def2-SVP basis set by Ahlrichs and coworkers including relativistic effective core potential for Au and Ag atoms.33 The energy of ligand 5a was calculated depending on the rotation around the connecting C–N bond, showing the difference in the stability of the two major conformations labeled as “bridging” and “chelating” (Fig. 4a). In the bridging conformation, the triazole nitrogen atoms face inwards and the rotational energy is minimal. This conformation is not only lower in energy according to the computational data, but also the preferred conformation observed experimentally in the crystal lattice (see Fig. S17, ESI†). The chelating conformation is presumably destabilized by the repulsion of the triazole and quinoxaline hydrogen atoms (see Fig. 4b). It can be assumed that the ligand will usually adopt the bridging conformation in the absence of a metal or in the presence of a mono-coordinated metal center, and will only remain in the less favorable chelating conformation if a stable chelate complex is formed with a suitable metal.
We intended to obtain theoretical information describing the stability and electronic distribution of complexes 18b–20, 25, and 26a compared to the properties of the ligand 5a alone. Fig. 5 gives an overview of the complex structures and the notation used for the following paragraphs.
![]() | ||
Fig. 5 (a) Moieties for the NICS values and notation for the NBO charges of ligand 5a. (b) Overview of experimentally confirmed complexation motifs. |
In coordination compounds, elucidating the bonding nature by modeling proves to be complicated. Our approach integrates (1) aromaticity of the ligand (NICS), (2) electrostatic (NBO), and (3) π-backdonation from the metal center to the click ligand (normal mode frequency) mechanisms to elucidate the diverse stabilization phenomena observed across various compounds. To evaluate the aromaticity based on the magnetic properties of the compound, the nucleus-independent chemical shifts (NICS)34 were computed. Herein, a more negative value is connected to higher aromaticity of the respective moiety and thus increased stabilization effects. The lowest overall shifts among the investigated complexes for the triazole, pyrazine, and phenyl moieties are observed for the gold complex 20 (Fig. 6). The same behavior can be noticed when calculating the out-of-plane NICSzz values, a more accurate variant for most applications.35 This indicates the highest stability for complex 20 among the investigated 1,2,3-triazolylquinoxaline complexes. The NICS values also help to explain why a monodentate complexation of silver tetrafluoroborate with the ligand in bridging conformation is observed for the heteroleptic silver complex 25. The NICSzz values for a hypothetical chelating complexation are higher than for the observed complexation in the bridging conformation, suggesting a lower stability for the chelating complexation (see Table S5, ESI†). The same trend is present when switching the ligand from 5a to 5c; despite the higher steric hindrance of the phenyl residue in the bridging conformation, the monodentate complexation is still assumed to occur in the silver complex.
Both 5a and 5c should be aromatic according to the Hückel rule (molecule 5c has 22 π electrons and 5a has 14 π electrons, both corresponding to 4n + 2 rule) and according to the NICS, but NICS alone is usually inadequate for assessing aromaticity due to its locality.36 Visualizing ring currents reveals the aromatic nature of molecules due to delocalized π electrons, aiding in the identification of aromatic systems. Utilizing the CTOCD (continuous transformation of origin of current density) method for gauge selection it is possible to visually inspect the non-local nature of the ring current as well as have structural information.37 The currents were broken down into currents originating from π manifold and currents from the remaining populated orbitals. When inspecting the molecular current maps (see Fig. S33 and S34, ESI†), it was found that the ring current is diatropic mostly from the π orbitals and consistency with NICS was verified.
The analysis of the natural bond orbitals (NBO) provides explanations for the distribution of atomic charges (see Table S4, ESI†). The N2 and N5 atoms are important for forming the chelate conformation, while the N3 and occasionally N4 atoms are important for the bridging mode. When the chelating conformation is stable, N2 and N5 atoms possess higher charges in this conformation than in the bridge conformation. When the bridging conformation is stable, the N3 and N4 atoms possess higher charges in the bridge conformation than in the chelate conformation. A higher charge on nitrogen atoms near the metal cation accounts for better charge neutralization in different conformations, building up their stabilities. To inspect π-backdonation to the ligand, we compared the frequencies of the normal mode possessing the same symmetry as the LUMO orbital of the ligand 5a (Fig. S20 in the ESI†) and found that the ligand is not a good π-acceptor as opposed to previous studies,38 which is reflected in a subtle change in the vibration frequency (Table S4, ESI†).
In addition, exemplary calculations on the complexation energies and Gibbs energies were conducted for complexes 18b–20, 25, and 26a, hypothetical complexes (entries 2, 6, and 7), and entries 9 and 10, without a crystal structure. The obtained energies should give information on the theoretical spontaneity of the respective complex formation step. For this purpose, the reaction of the ligand in the bridging conformation with the respective metal precursor was considered (see ESI for further information, pages 33–38†). The obtained energies support the experimental results; all formed complexes 18b–20, 25, and 26 have negative theoretical complexation and Gibbs energies. A complexation of 5a with silver acetate is not expected due to the positive free Gibbs energy of the reaction (Table 3, entry 2), whereas the exergonic reaction to the other silver complexes 18b, 19, and 25, summarized in Table 3, can occur. The strongly negative free energy of the silver nitrate complexation (18b, entry 1) might be related to simplifying the three-dimensional crystal lattice to a separated Ag+ and NO3− pair within the single-molecule modeling approach. A slight tendency toward the bridging conformation of the heteroleptic silver triazolylquinoxaline-phosphine complexation can be derived from entries 5 and 6 as well as 9 and 10. The observed complexation mode seems to be mostly favored due to the higher stability of the final complex (see section above), not only due to the energetically slightly favored formation. For the copper complex 26, the formation of the experimentally observed chelate complex is predicted to be slightly energetically favored compared to the bridging complexation (entries 7 and 8).
Entry | Ligand | Metal precursor | Complex | Variant | ΔE [kJ mol−1] | ΔG [kJ mol−1] |
---|---|---|---|---|---|---|
a Polymeric complex structure simplified for calculation as a tetramer with three silver atoms and four additional ammonia atoms to account for the next quinoxaline units. | ||||||
1 | 5a | AgNO3 | 18b | — | −1754.89 | −1548.31 |
2 | 5a | [AgCH3CO2]2 | — | — | −64.5 | 30.0 |
3 | 5a | AgSO3CF3 | 19 | — | −1288.74a | −921.14a |
4 | 5a | AuBr3 | 20 | — | −124.68 | −72.88 |
5 | 5a | [Ag(MeCN)4]BF4 | 25 | Bridge | −565.37 | −483.55 |
6 | 5a | [Ag(MeCN)4]BF4 | — | Chelate | −556.66 | −475.39 |
7 | 5c | [Cu(MeCN)4]BF4 | — | Bridge | −662.71 | −572.98 |
8 | 5c | [Cu(MeCN)4]BF4 | 26a | Chelate | −673.50 | −579.78 |
9 | 5c | [Ag(MeCN)4]BF4 | — | Bridge | −557.53 | −478.21 |
10 | 5c | [Ag(MeCN)4]BF4 | — | Chelate | −551.88 | −472.12 |
Footnote |
† Electronic supplementary information (ESI) available. CCDC 2336030–2336038. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d4dt02068e |
This journal is © The Royal Society of Chemistry 2025 |