Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Systematic investigation of the influence of electronic substituents on dinuclear gold(I) amidinates: synthesis, characterisation and photoluminescence studies

Frederic Krätschmer a, Xin Gui b, Michael T. Gamer a, Wim Klopper b and Peter W. Roesky *a
aInstitute of Inorganic Chemistry, Karlsruhe Institute of Technology, Engesserstr. 15, 76131 Karlsruhe, Germany. E-mail: roesky@kit.edu
bInstitute of Physical Chemistry, Karlsruhe Institute of Technology, Fritz-Haber-Weg 2, 76131 Karlsruhe, Germany

Received 9th November 2021 , Accepted 24th February 2022

First published on 10th March 2022


Abstract

Dinuclear gold(I) compounds are of great interest due to their aurophilic interactions that influence their photophysical properties. Herein, we showcase that gold–gold interactions can be influenced by tuning the electronic properties of the ligands. Therefore, various para substituted (R) N,N′-bis(2,6-dimethylphenyl)formamidinate ligands (pRXylForm; Xyl = 2,6-dimethylphenyl and Form = formamidinate) were treated with Au(tht)Cl (tht = tetrahydrothiophene) to give via salt metathesis the corresponding gold(I) compounds [pRXylForm2Au2] (R = –OMe, –Me, –Ph, –H, –SMe, and –CO2Me). All complexes showed intense luminescence properties at low temperatures. Alignment with the Hammett parameter σp revealed the trends in the 1H and 13C NMR spectra. These results showed the influence of the donor–acceptor abilities of different substituents on the ligand system which were confirmed with calculated orbital energies. Photophysical investigations showed their lifetimes in the millisecond range indicating phosphorescence processes and revealed a redshift with the decreasing donor ability of the substituents in the solid state.


Introduction

Due to its applications in various fields such as catalysis, medicine and small molecule activation, increasing research on molecular gold(I) compounds has been observed during the last few decades.1 In particular, dinuclear gold(I) complexes have been investigated in terms of their photophysical properties, which are influenced in many cases by aurophilic interactions.2–7 The concept of “aurophilic interactions” or "aurophilicity" was established by H. Schmidbaur in 1989 and describes the attractive interactions of gold cations.8,9 These attractive forces between cations, which are at first glance counter-intuitive, have been the subject of intensive theoretical investigations. Pyykkö showed by theoretical methods that aurophilicity is mainly a consequence of dispersion forces and is enhanced by relativistic effects.10 Although discussions about the concept are still ongoing,11–15 this interpretation has been widely accepted.16–21 Nowadays, it is common knowledge that such attractive forces are not only limited to gold cations but also observed in compounds containing other heavy-metal closed-shell cations.22,23 Thus, aurophilicity is just a special case of so-called metallophilicity. In addition, argentophilic and mercurophilic interactions have been observed and described.24,25 In general, aurophilic interactions are likely to exist if the distance of two gold atoms is below 3.5 Å, although the respective van der Waals distance (3.32 Å) is shorter.5,20,26,27 The binding energy of these interactions may reach values of up to 50 kJ mol−1, which can be divided into three cases: fully-supported (I), semi-supported (II) and unsupported (III) (Fig. 1).28,29 In comparison with mononuclear gold(I) compounds, different photophysical properties were observed in the presence of Au⋯Au28 or Au(I)⋯M interactions (M = heavy atom).30,31 The excited state of complexes with aurophilic Au⋯Au contacts has been suggested to be due to ligand-to-metal–metal charge-transfer (LMMCT).32–34
image file: d1dt03795a-f1.tif
Fig. 1 Classification of closed-shell aurophilic interactions into fully-supported (I), semi-supported (II) and unsupported (III).

An established method for the synthesis of dinuclear gold(I) complexes featuring “fully supported” aurophilic interactions is the employment of amidinate ligands.35–37 Following the synthesis of the first gold(I) bis(amidinate) complex by Dehnicke and co-workers,38 this field was expanded to the synthesis of several other bis(amidinate) gold(I) compounds.39–42 Pioneering work on the reactivity of dinuclear gold(I) amidinate compounds towards small molecules was carried out by Fackler et al. They investigated the oxidative addition of small molecules (e.g., Cl2, Br2, and I2) to the dinuclear gold(I) complex [Au2(2,6-Me2-Form)2], which resulted in the formation of complexes with an Au(II)⋯Au(II) bond, with the most stable and most commonly found Au24+ core.6,43–45

Another possibility to enhance the properties of dinuclear gold(I) complexes is the functionalisation of the NCN backbone. Previously, our group reported the photophysical properties of acetylide functionalised gold(I) complexes and the investigation of multinuclear coinage metal compounds.46–50

In general, amidinate ligands are available with a large variety of substituents. By using para-phenyl substituents on the nitrogen atom, the electronic properties of the resulting ligands can be selectively tuned without changing the steric demand.51 In light of these properties, we faced the challenge to synthesise various dinuclear gold(I) compounds, which are ligated by amidinate ligands having different electronic properties (Scheme 1). As a scale for the intensity of the electronic influence, the Hammett parameter was applied.52,53 Similar studies have been reported by Tunik et al., who investigated the influence of the ligand substituents of gold–copper alkynyl complexes on their photoluminescence properties. Using the Hammett parameter as the reference, they were able to reveal a substantial redshift of the emission maxima with the increase in the electron-donating properties of the ligands.54


image file: d1dt03795a-s1.tif
Scheme 1 Target complexes in which the amidinate ligands have different electronic properties but a similar steric demand upon coordination.

Herein, we report the synthesis of six different dinuclear gold complexes exhibiting fully supported aurophilic interactions by using a variety of formamidinate ligands with different substituents in para position on the 2,6-dimethylphenyl side groups. All complexes were fully characterised by SCXRD, multinuclear NMR spectroscopy, IR spectroscopy, ESI mass spectrometry, elemental analysis and photoluminescence experiments. In addition, theoretical studies regarding the influence of the respective substituents on the supported aurophilic interactions and luminescence properties were performed at the GW approximation and Bethe–Salpeter equation (GW/BSE) level.

Results and discussion

Synthesis and characterisation

The setup of the ligand framework is shown in Scheme 1. As a substituent of the nitrogen atoms, we used 2,6-xylidine (2,6-dimethylphenyl) as the basic scaffold and added different substituents (Scheme 2) at the para-position. By using this setup, we generated ligands with essentially the same steric demand upon coordination. The differently substituted amidine ligands L1–L6 were synthesised by an ethanol condensation reaction from the corresponding amines and triethyl orthoformate (Scheme 2).55–64
image file: d1dt03795a-s2.tif
Scheme 2 Synthesis of amidine ligands L1–L6via the ethanol condensation reaction.

While most of the reagents for the ligand synthesis are commercially available, the amine precursors for L1, L3 and L5 were synthesised according to literature-known procedures.59,62,63 In the case of L3, 4-phenyl-xylidine was obtained by the literature-known Suzuki coupling reaction of 4-bromo-xylidine with phenylboronic acid, followed by amidine condensation.64 We also established a second route for obtaining this ligand by synthesising N,N′-bis(4-bromo-2,6-dimethylphenyl)formamidine first,57 followed by cross coupling under similar conditions (Scheme 3). The advantage is the significantly better separation through column chromatography of the cross-coupling product. In the end, we were able to synthesise and fully characterise the new amidine ligands L3, L5 and L6.


image file: d1dt03795a-s3.tif
Scheme 3 Synthesis routes of L3 with Suzuki coupling before (left) and after (right) amidine formation.

Depending on the solvent used, the obtained 1H NMR spectra of L1–L6 show complex resonances due to E/Z isomerisation and tautomerisation.51 Ligands L1, L4 and L6 exist as two isomers in a 2[thin space (1/6-em)]:[thin space (1/6-em)]3 ratio when measured in C6D6, showing a pattern of three singlets in a 1[thin space (1/6-em)]:[thin space (1/6-em)]3[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio for the 2,6-methyl protons, while L2 and L3 exist in a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 ratio with a 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]1 pattern. The two small singlets belong to the 2,6-methyl protons of the E-isomer and the large singlet to the Z-isomer. These patterns were also found for the phenyl protons of L1, L2, L4 and L6 in the 1H NMR and 13C{1H} NMR spectra (see the ESI Fig. S1–S6, S11 and S12). In addition, two signals were observed for the NH proton, one at around 5 ppm and the other downfield shifted at around 7 ppm. In the 13C{1H} NMR spectra, the NCN carbon resonances appear between 140 and 150 ppm for all ligands. The N–H stretching mode occurs in the typical range at around 3200 cm−1 in the IR spectra (ESI Fig. S35–S40).51

Subsequently, the bis(amidinate) gold(I) complexes 1–6 were synthesised via salt elimination from the in situ generated potassium salt of ligands L1–L6 and Au(tht)Cl in THF (Scheme 4).


image file: d1dt03795a-s4.tif
Scheme 4 Synthesis of gold complexes 1–6.

Compounds [pRXylForm2Au2] R = −OMe (1), −Me (2), −Ph (3), −SMe (5) and −CO2Me (6) were crystallised by slow evaporation of THF (1–3 and 5) or toluene (6), while single crystals of [XylForm2Au2] (4) were obtained by a literature known procedure (Fig. 2).44 The complexes were obtained in good yields between 40% and 68%. Complete deprotonation of the ligands was confirmed by the disappearance of the N–H band in the IR spectra (ESI Fig. S41–S46) and the corresponding N–H resonance in the 1H NMR spectra (ESI Fig. S13–S26). The complexes are stable in the solid state under an inert atmosphere and exclusion of light for several months, but in solution, decomposition occurs within a few days. Compounds 1, 4 and 6 crystallise in the triclinic space group P[1 with combining macron], 2 in the monoclinic space group C2/c, 3 in P21/c and 5 in the trigonal space group R[3 with combining macron]c. The Au⋯Au distances range from 2.7079(7) Å in 1 to 2.7366(2) Å in 3 (Fig. 2). These compounds show similar Au⋯Au distances as reported in the literature for other bimetallic gold amidinate complexes, for example [(Me3SiC[triple bond, length as m-dash]CC(NDipp)2)2Au2] (2.7009(11) Å).46 The smallest amidinate bite angle (126.7(3)°) is seen for compound 1 and the largest for 4 (128.5(9)°). These values match well with the literature.44,46 The Au–N bond lengths for each compound (approximately 2.03 Å) are in good agreement with those of other literature known gold amidinates. Moreover, the N–C1 bond lengths (ca. 1.30 Å), which are approximately the same in all complexes, fit well with those of some other related gold(I) amidinate complexes.41,42,46 The twist angle (C1–Nx–Ph; x = 1 and 2) of the phenyl rings of the amidinates differs for each compound. The phenyl rings of 2 and 4–6 are twisted conrotatory, with compounds 4 and 6 having the largest difference between their angles on each side (30° and 1°), while compounds 2 and 5 have almost an identical twist angle of around 10°. In contrast, the phenyl rings of compounds 1 and 3 are turning disrotatory, having one smaller and one larger angle; for example, 1 (4° and 11°) (Fig. 3).


image file: d1dt03795a-f2.tif
Fig. 2 Molecular structure of compounds 1–3, 5 and 6 in the solid state. Hydrogen atoms and non-coordinating solvent molecules are omitted for clarity. Structural parameters are given in the ESI (Fig. S65–S69).

image file: d1dt03795a-f3.tif
Fig. 3 Rotation of the phenyl rings to each other.

Since the Au⋯Au distances are well below 3.5 Å, aurophilic interactions appear to exist in the sense of distance criterion, but one should be aware of the fact that we are dealing with fully supported aurophilic interactions. The short distance may very well result simply from the structural properties of the ligand. To investigate this in detail, we have performed Hartree–Fock calculations, which do not account for dispersion forces at all and post-Hartree–Fock calculations, which do account for dispersion (e.g., Møller–Plesset perturbation theory to the 2nd order, MP2, and spin-component-scaled MP2, SCS-MP2). Calculations were performed only for compound 4. However, its silver(I) counterpart (4-Ag), was also studied, because the argentophilic interactions are usually much smaller than the aurophilic ones. Results for the relevant structural parameters are shown in Table 1.

Table 1 Selected distances and angles of compound 4 and its silver(I) counterpart 4-Ag, as obtained as various levels of computation
Distance Methoda 4 4-Ag
a The basis set used is def2-TZVP. b In the frozen-core approximation.
M1⋯M1′, M = Au, Ag (Å) Hartree–Fock 2.852 294.5
PBE0 2.759 2.770
PBE0-D3(BJ) 2.752 2.763
SCS-MP2b 2.738 2.777
MP2b 2.702 2.743
M1⋯N1, M = Au, Ag (Å) Hartree–Fock 2.102 2.208
PBE0 2.044 2.101
PBE0-D3(BJ) 2.039 2.096
SCS-MP2b 2.009 2.061
MP2b 1.988 2.030
∠M1–N1–C1, M = Au, Ag (°) Hartree–Fock 122.6 124.6
PBE0 121.4 122.3
PBE0-D3(BJ) 121.5 122.3
SCS-MP2b 121.5 122.8
MP2b 121.3 122.5


It is clear from the computational results shown in Table 1 that in fully supported situations, the Au⋯Au distance criterion (<3.5 Å) is less suited to identify aurophilic interactions. The short distance is mainly a consequence of the ligand structure, i.e., the Au1–N1–C1 (cf.Table 1) and the N1–C1–N2 angles (between 127.0° and 128.5°). Very similar distances are found for Ag, although the dispersion forces should be much smaller than those for Au. From the viewpoint of the Hartree–Fock theory, it became clear that the electron-correlation effects are important (cf. SCS-MP2 and MP2 results), but they just lead to an overall shrinkage of the molecular structure. Nevertheless, Au⋯Au electronic interactions do exist and influence the photophysical properties. In the fully supported case, however, these interactions are enforced by the supporting ligands and should perhaps not be called “aurophilic interactions”, since the latter is solely due to the London dispersion interactions between filled 5d10 shells.

Neither the electronic influence of the substituents seems to have a significant impact on Au⋯Au distances as shown in Table 2. The electronic properties of these substituents can be described by the Hammett parameter σp with –OMe (1) being the strongest electron-donating group and –CO2Me (6) being the strongest electron withdrawing group. The Au⋯Au distances in the solid state structures are most likely influenced by the packing effects due to the steric demands of the substituents of the phenyl ring and cannot be ascribed to the electronic influence.

Table 2 Hammett parameter σp for the substituents compared with the Au⋯Au distance of compounds 1–6
  1 2 3
R –OMe –Me –Ph
σ p −0.27 −0.17 −0.01
Au⋯Au 2.7079(7) 2.7277(4) 2.7366(2)

  4 5 6
R –H –SMe –CO2Me
σ p 0.00 0.00 0.45
Au⋯Au 2.711(3) 2.7307(12) 2.7194(2)


In contrast to the 1H NMR spectra of L1–L6, well-defined signals are observed for 1–6. Due to the deprotonation, all isomers are transformed into only one symmetric species, which coordinates to the gold atoms. It turned out that the N2C(H) proton, which was detected in the range of 7.42 to 7.66 ppm, is a suitable probe for studying the electronic influence of the different substituents of the phenyl rings in the para-position on the ligand. Using the Hammett parameter, the signals for the proton in the backbone of compounds 1–6 can be lined up (except 3) in a linear correlation according to the withdrawing and donating ability of their substituents, resulting in a shift of around 0.2 ppm (Fig. 4). This is related to the shielding of the proton by electron density, which is influenced by the substituents. The proton signal of compound 3 cannot be aligned with the others when measured in THF-d8.


image file: d1dt03795a-f4.tif
Fig. 4 Influence of the Hammett parameter σp of several substituents on the chemical shift of the N2C(H) proton of compounds 1–6 measured in THF-d8.

By changing the solvent to CDCl3, a significant upfield shift from 7.66 ppm (THF-d8) to 7.53 ppm (CDCl3) is observed in the 1H NMR spectrum of 3. This leads to a conclusion that compared to the other complexes, compound 3 shows a different behaviour in THF solution. We suggest that this difference arises from the improved π-stacking interactions of the phenyl substituents in polar solvents.65

The NCN carbon atom of the ligand backbone is a sensitive probe for 13C{1H} NMR. Upon deprotonation of the ligand and coordination towards the gold centers, a strong downfield shift starting from around 140 ppm (NCNH) to about 170 ppm (NCN) was detected in the 13C{1H} NMR spectra. Compound 6 shows a slightly less shifted signal at around 169.5 ppm, while 1 shows the most significant downfield shifted resonance at 171.2 ppm. An opposite trend to that mentioned above for the 1H shifts can be observed for the NCN carbon signals, where electron-donating groups lead to a downfield shift of around 1.7 ppm from 6 to 1 (see the ESI Fig. S27 (right)).

Photoluminescence properties

We examined the luminescence properties of dinuclear gold(I) complexes 1–6 in the solid state and in frozen solution (THF) at 77 K. All compounds showed bright blue-green luminescence at low temperatures. UV-vis spectra were recorded for compounds 1–6 showing similar absorption with a band for 1, 2 and 4 at 256 nm, which is redshifted for 3 (283 nm), 5 (272 nm) and 6 (260 nm). An additional band for compound 6 was detected at 288 nm (see the ESI Fig. S47–S52). Note that the samples were measured in THF, and thus the band at 112 nm was referred to as the solvent.

In the photoluminescence emission (PL) and excitation (PLE) spectra, compounds 1–6 show broad absorption at low temperatures from UV (280 nm) up to 370 nm in the solid state and broad emission bands in the range from around 400 nm to 600 nm. Similar spectra were obtained for 1–2 and 4–6 in the solid state (Fig. 5). Only compound 3, while strongly red-shifted, shows well-defined emission bands featuring the typical vibronic pattern of luminescent gold compounds.66,67


image file: d1dt03795a-f5.tif
Fig. 5 Excitation and emission spectra of compounds 1–6 in the solid state at 77 K. The PLE/PL spectra were excited/recorded at the values given on the right side of the graph. Photographs of the PL samples: UV lamp (λexc = 256 nm).

The spectra of compounds 1–6 (except 4) in solution resemble the spectra in the solid state. For 4, a decrease in absorption at 310 nm in solution is seen. Moreover, the emission maximum of 4 in the solid state is observed at 425 nm with a second smaller band at 523 nm, which is reversed in solution with the emission maximum at a longer wavelength (519 nm) (Fig. 6). A blueshift in the emission spectra of 1, 3 and 4 can be observed from the solid state to solution, while those of 2, 5 and 6 are red-shifted (Table 3).


image file: d1dt03795a-f6.tif
Fig. 6 Excitation and emission spectra of compounds 1–6 in frozen solution at 77 K. The PLE/PL spectra were excited/recorded at the values given on the right side of the graph. Photographs of the PL samples: UV lamp (λexc = 256 nm).
Table 3 Values of excitation and emission λmax [nm] in the solid state and in THF at 77 K of compounds 1–6
  1 2 3 4 5 6
λ max exc, solid 313 343 349 316 350 358
λ max em, solid 416 419 501 425 443 451
λ max exc, THF 298 300 282 284 334 353
λ max em, THF 437 429 492 519 449 446


Another major difference between the emission in the solid state and in solution is the appearance of weak bands at higher wavelengths for 1, 5 and 6. While 6 shows an additional weak band at 700 nm in the solid state, which is not visible in solution, 1 and 5 feature extra bands in solution (570 nm for 1 and 690 nm for 5).

Table 3 shows a weak trend for λmax in the emission spectra in the solid state. Upon increasing the electron withdrawing ability (increasing Hammett parameter σp), the maximum emission wavelength tends to increase due to the electronic influence of the substituents leading to a shift of 35 nm for 1 to 6. This trend, however, is only observed in the solid state. In contrast to this, solvent effects overlay this weak trend. The largest difference can be seen for compound 4, where the emission maximum is shifted by around 100 nm. Compound 3 cannot be aligned with the others. A strong redshift for the emission maximum was detected, presumably due to the abovementioned π-interaction also seen for the 1H NMR signal of N2C(H) of compound 3 (Fig. 4).

Earlier, Tunik et al. reported an opposite result of the influence of the electronic properties represented by the Hammett parameter. They observed for gold–copper alkynyl complexes a bathochromic shift with the increase in donor ability.54 However, their ligand system and metal atom scaffold differ significantly from our rather simple dinuclear gold(I) core and amidinate ligation. This leads to the conclusion that the impact of the electronic influence can be altered with the ligand system.

Compounds 1–6 show phosphorescence, measured at λmax, with decay times that are in the range of microseconds from about 0.02 ms to 1.36 ms (see the ESI Fig. S34 and Table S1). Most of the compounds feature complicated and unusual PL kinetics, meaning that their decay traces can only be well fitted by at least two exponential components. Compound 3 exhibits a particularly long phosphorescence lifetime of 1.36 ms in THF and 1.23 ms in the solid state.

Quantum chemical calculations

To further investigate the photophysical properties of the dinuclear gold(I) amidinate complexes with different modifications on the ligand, quantum chemical calculations were performed with the TURBOMOLE program package.68 The ground-state equilibrium geometries were optimised at the PBE0-D3(BJ)/def2-TZVP level of theory,69–72 showing good agreement with the experimental data for the bond lengths in the complex but a slight divergence for the angles of the substituted phenyls, and an overall elongation of the Au⋯Au interaction (Cartesian coordinates can be found in the ESI). Excited-state calculations utilising the GW approximation and Bethe–Salpeter equation (GW/BSE)73,74 were performed at the one-component (1c, scalar-relativistic) level for absorption spectra, and the quasirelativistic two-component (2c, including spin–orbit coupling) level for emission spectra. Computations were carried out for hypothetical complexes in addition to complexes 1–6 with dimethylamino (complex 7), bromo (complex 8) and nitro (complex 9) substituents at the para-positions of the phenyl rings (see the ESI Fig. S70).

The calculated frontier orbital energies of complexes 1–6 show a clear decreasing trend with the increasing electron-withdrawing ability of the substituents (Fig. 7). This trend still remains after taking the two extreme cases 7 and 9 into account (see the ESI Table S6). Due to the similar drop in the HOMO and LUMO energies when bearing different substituents, the trend along the HOMO–LUMO gap of the gold complexes is rather weak, leading to only small differences in photoluminescence.


image file: d1dt03795a-f7.tif
Fig. 7 Frontier orbital energies (eV) with respect to the Hammett parameter of the substituents.

All complexes display metal-to-ligand charge-transfer (MLCT) absorption bands at around 240–270 nm (ESI Fig. S71), which fit with the experimental UV-vis spectra (Fig. 8). The two alkyl-substituted complexes 2 and 4 exhibit very similar absorption profiles with the MLCT peak centred at 240 nm. The introduction of the electron-donating methoxy groups does not alter the absorption profile of complex 1, resulting only in a slightly higher intensity of the MLCT peak. However, when the methoxy groups are replaced with methylthio groups, a significant redshift and increased intensity are observed in the predicted absorption spectrum of complex 5. On the other hand, after introducing the electron withdrawing methoxycarbonyl groups, the predicted absorption spectrum of complex 6 features a significantly redshifted and broader MLCT absorption band. Interestingly, the phenyl substituted complex 3 shows the most intense absorption peak. According to the unrelaxed difference density plots, the electron-donating and electron-withdrawing properties of the substituents are reflected in the change of electron density. However, the charge transfer from the gold centre to the aromatic groups of the amidinate ligands does not obviously differ and is actually quite similar among all the complexes.


image file: d1dt03795a-f8.tif
Fig. 8 Comparison of the experimental (blue) and calculated (black) UV-vis spectra.

The calculated emission maxima are in good agreement with the experimental data. The typical MLCT peak in the visible range is split into two distinct (shoulder) peaks, and the high-energy peak consists of one component of the most intense triplet excitation in the energetic range between S1 and T1, which starts to gain intensity from strong mixing with the singlet states. Natural transition orbital (NTO) analysis further confirms the MLCT character of the triplet excited states, and the NTOs are quite similar among all the complexes. Besides, all complexes turn out to present similar lifetimes in the range of tens or a few hundreds of microseconds, which also agree well with the experiment.

As shown in Fig. 9 and in Table S6, with the increasing electron-withdrawing abilities of the substituents, the absorption spectra tend to redshift while the emission spectra tend to blueshift. Therefore, compared to the measured emission spectra in the solid state, the calculated gas-phase emission spectra exhibit an opposite trend with respect to the Hammett parameter. When the solvent effects are accounted for in the computations by means of the conductor-like screening model (COSMO, ε = ∞),75 the trend is still opposite. Thus, solvent effects are not the reason for the opposite dependence on the Hammett parameter seen in the computations. Rather, we note that the frontier orbitals respond to the electron-donating and electron-withdrawing properties in a very similar manner (Fig. 7), leading to large uncertainties when computing the differences (e.g., HOMO–LUMO gap).


image file: d1dt03795a-f9.tif
Fig. 9 Emission maxima (nm) with respect to the Hammett parameter of the substituents.

The different behaviour of compound 3 indicates the presence of some sort of π-interaction. Also, the shifted λmax value of compound 4 from the solid state to solution arises most likely from solvent interactions.

For a detailed description of the computational methods, the calculated absorption and emission spectra, and a detailed analysis of the relevant excitation (including natural transition orbitals and unrelaxed difference density plots), we refer to the ESI.

Summary

We presented the synthesis and characterisation of several dinuclear gold(I) complexes. These complexes were easily obtained by deprotonation of an amidine ligand in situ and addition to the gold precursor in THF. We further investigated the influence of the electronic groups attached at the para position of the side groups of the amidine ligands on the behaviour of the dinuclear gold(I) complexes. The influence of the different substituents resulted in changed NMR shifts and luminescence properties. With the increasing Hammett parameter, we observed a downfield shift of the signal of the proton in the formamidinate backbone (N2C(H) proton). Furthermore, a redshift of the emission bands was seen in the solid state. In both cases, only the phenyl-substituted compound does not fit in this trend, most probably due to the π-interactions.

Quantum chemical calculations showed a clear trend for the frontier orbital energies and a less obvious trend for the HOMO–LUMO gap energies. Further calculations showed an MLCT band fitting with the UV-vis spectra and good agreement of the emission maxima with the experimental data. An inverse trend was observed for the emission maxima in the luminescence spectra between the experimental solid state measurement and the calculated gas-phase spectra, which cannot be explained by the missing solvent and intermolecular interactions in the calculations. Geometry optimisation of all compounds showed that the overall bonds are in good agreement, but the phenyl angles and Au⋯Au distances differ slightly from the experimental data.

Author contributions

The manuscript was written through the contributions of all authors. F. K. synthesised and analysed all presented compounds under the supervision of P. W. R. Solid state and solution PL analyses were conducted by F. K. and the data were interpreted by F. K. Theoretical investigations were performed by X. G. under the supervision of W. K. XRD data were refined by M. T. G. The original idea was from P. W. R. who supervised the work and interpreted the data. All authors have given approval to the final version of the manuscript.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

We thank Christoph Schoo for his help with the crystal data. Financial support from the DFG funded Transregional Collaborative Research Centre SFB/TRR 88 “Cooperative Effects in Homo and Heterometallic Complexes (3MET)” is gratefully acknowledged (projects C1 and C3).

Notes and references

  1. R. S. Ramon, N. Marion and S. P. Nolan, Chem. – Eur. J., 2009, 15, 8695–8697 CrossRef CAS PubMed.
  2. S. J. Freakley, Q. He, C. J. Kiely and G. J. Hutchings, Catal. Lett., 2014, 145, 71–79 CrossRef.
  3. A. S. Hashmi and G. J. Hutchings, Angew. Chem., Int. Ed., 2006, 45, 7896–7936 CrossRef PubMed.
  4. A. S. Hashmi and M. Rudolph, Chem. Soc. Rev., 2008, 37, 1766–1775 RSC.
  5. H. Schmidbaur and A. Schier, Chem. Soc. Rev., 2008, 37, 1931–1951 RSC.
  6. H. E. Abdou, A. A. Mohamed and J. P. Fackler, Jr., Inorg. Chem., 2007, 46, 9692–9699 CrossRef CAS PubMed.
  7. N. L. Coker, J. A. Bauer and R. C. Elder, J. Am. Chem. Soc., 2004, 126, 12–13 CrossRef CAS PubMed.
  8. H. Schmidbaur, Gold Bull., 1990, 23, 11–21 CrossRef CAS.
  9. F. Scherbaum, A. Grohmann, G. Müller and H. Schmidbaur, Angew. Chem., Int. Ed. Engl., 1989, 28, 463–465 CrossRef.
  10. P. Pyykkö and Y. Zhao, Angew. Chem., Int. Ed. Engl., 1991, 30, 604–605 CrossRef.
  11. M. B. Brands, J. Nitsch and C. F. Guerra, Inorg. Chem., 2018, 57, 2603–2608 CrossRef CAS PubMed.
  12. E. Andris, P. C. Andrikopoulos, J. Schulz, J. Turek, A. Ruzicka, J. Roithova and L. Rulisek, J. Am. Chem. Soc., 2018, 140, 2316–2325 CrossRef CAS PubMed.
  13. A. Lobato, M. A. Salvadó and J. M. Recio, Chem. Sci., 2021, 12, 13588–13592 RSC.
  14. C. Yin, U. Wedig and M. Jansen, Chem. Sci., 2021, 12, 13593–13596 RSC.
  15. A. Kovalevskiy, C. Yin, J. Nuss, U. Wedig and M. Jansen, Chem. Sci., 2019, 11, 962–969 RSC.
  16. P. Pyykko, Angew. Chem., Int. Ed., 2004, 43, 4412–4456 CrossRef PubMed.
  17. P. Schwerdtfeger, A. E. Bruce and M. R. M. Bruce, J. Am. Chem. Soc., 1998, 120, 6587–6597 CrossRef CAS.
  18. P. Pyykko, Chem. Soc. Rev., 2008, 37, 1967–1997 RSC.
  19. P. Pyykkö, Inorg. Chim. Acta, 2005, 358, 4113–4130 CrossRef.
  20. G. J. Hutchings, M. Brust and H. Schmidbaur, Chem. Soc. Rev., 2008, 37, 1759–1765 RSC.
  21. H. Schmidbaur, Gold Bull., 2000, 33, 3–10 CrossRef CAS.
  22. M. Jansen, Angew. Chem., Int. Ed. Engl., 1987, 26, 1098–1110 CrossRef.
  23. M. Jansen, J. Less-Common Met., 1980, 76, 285–292 CrossRef CAS.
  24. H. Schmidbaur and A. Schier, Angew. Chem., Int. Ed., 2015, 54, 746–784 CrossRef CAS.
  25. H. Schmidbaur and A. Schier, Organometallics, 2015, 34, 2048–2066 CrossRef CAS.
  26. S. Alvarez, Dalton Trans., 2013, 42, 8617–8636 RSC.
  27. H. Schmidbaur and A. Schier, Chem. Soc. Rev., 2012, 41, 370–412 RSC.
  28. C.-M. Che and S.-W. Lai, Gold Chemistry, 2009, pp. 249–281 Search PubMed.
  29. M. Contel, Angew. Chem., Int. Ed., 2010, 49, 250–251 CrossRef CAS.
  30. M. A. Omary, A. A. Mohamed, M. A. Rawashdeh-Omary and J. P. Fackler, Coord. Chem. Rev., 2005, 249, 1372–1381 CrossRef CAS.
  31. E. J. Fernandez, A. Laguna and J. M. Lopez-de-Luzuriaga, Dalton Trans., 2007, 1969–1981 RSC.
  32. V. Wing-Wah Yam, C.-K. Li and C.-L. Chan, Angew. Chem., Int. Ed., 1998, 37, 2857–2859 CrossRef PubMed.
  33. A. L. Balch, Angew. Chem., Int. Ed., 2009, 48, 2641–2644 CrossRef CAS PubMed.
  34. C. K. Li, X. X. Lu, K. M. Wong, C. L. Chan, N. Zhu and V. W. Yam, Inorg. Chem., 2004, 43, 7421–7430 CrossRef CAS PubMed.
  35. E. C. Taylor and W. A. Ehrhart, J. Org. Chem., 1963, 28, 1108–1112 CrossRef CAS.
  36. K. Kaji, H. Matsubara, H. Nagashima, Y. Kikugawa and S. Yamada, Chem. Pharm. Bull., 1978, 26, 2246–2249 CrossRef CAS.
  37. E. C. Taylor and W. A. Ehrhart, J. Am. Chem. Soc., 1960, 82, 3138–3141 CrossRef CAS.
  38. D. Fenske, G. Baum, A. Zinn and K. Z. Dehnicke, Z. Naturforsch., B: J. Chem. Sci., 1990, 45, 1273–1278 CrossRef CAS.
  39. F. A. Cotton, X. J. Feng, M. Matusz and R. Poli, J. Am. Chem. Soc., 1988, 110, 7077–7083 CrossRef CAS.
  40. E. Hartmann and J. Strähle, Z. Naturforsch., B: J. Chem. Sci., 1989, 44, 1–4 CrossRef CAS.
  41. J. P. Coyle, P. G. Gordon, A. P. Wells, D. J. Mandia, E. R. Sirianni, G. P. A. Yap and S. T. Barry, Chem. Mater., 2013, 25, 4566–4573 CrossRef CAS.
  42. T. J. J. Whitehorne, J. P. Coyle, A. Mahmood, W. H. Monillas, G. P. A. Yap and S. T. Barry, Eur. J. Inorg. Chem., 2011, 3240–3247 CrossRef CAS.
  43. H. E. Abdou, A. A. Mohamed and J. P. Fackler, J. Cluster Sci., 2007, 18, 630–641 CrossRef CAS.
  44. H. E. Abdou, A. A. Mohamed and J. P. Fackler, Jr., Inorg. Chem., 2005, 44, 166–168 CrossRef CAS PubMed.
  45. D. Y. Melgarejo, G. M. Chiarella, J. P. Fackler, Jr., L. M. Perez, A. Rodrigue-Witchel and C. Reber, Inorg. Chem., 2011, 50, 4238–4240 CrossRef CAS PubMed.
  46. T. J. Feuerstein, M. Poss, T. P. Seifert, S. Bestgen, C. Feldmann and P. W. Roesky, Chem. Commun., 2017, 53, 9012–9015 RSC.
  47. T. J. Feuerstein, T. P. Seifert, A. P. Jung, R. Muller, S. Lebedkin, M. M. Kappes and P. W. Roesky, Chem. – Eur. J., 2020, 26, 16676–16682 CrossRef CAS PubMed.
  48. M. Dahlen, M. Kehry, S. Lebedkin, M. M. Kappes, W. Klopper and P. W. Roesky, Dalton Trans., 2021, 50, 13412–13420 RSC.
  49. M. Dahlen, E. H. Hollesen, M. Kehry, M. T. Gamer, S. Lebedkin, D. Schooss, M. M. Kappes, W. Klopper and P. W. Roesky, Angew. Chem., Int. Ed., 2021, 60, 23365–23372 CrossRef CAS PubMed.
  50. M. Dahlen, N. Reinfandt, C. Jin, M. T. Gamer, K. Fink and P. W. Roesky, Chem. – Eur. J., 2021, 27, 15128–15136 CrossRef PubMed.
  51. M. P. Coles, Dalton Trans., 2006, 985–1001 RSC.
  52. L. P. Hammett, J. Am. Chem. Soc., 1937, 59, 96–103 CrossRef CAS.
  53. L. P. Hammett, Trans. Faraday Soc., 1938, 34, 156–165 RSC.
  54. J. R. Shakirova, E. V. Grachova, V. V. Gurzhiy, I. O. Koshevoy, A. S. Melnikov, O. V. Sizova, S. P. Tunik and A. Laguna, Dalton Trans., 2012, 41, 2941–2949 RSC.
  55. E. M. McGarrigle, S. P. Fritz, L. Favereau, M. Yar and V. K. Aggarwal, Org. Lett., 2011, 13, 3060–3063 CrossRef CAS PubMed.
  56. K. E. Krahulic, G. D. Enright, M. Parvez and R. Roesler, J. Am. Chem. Soc., 2005, 127, 4142–4143 CrossRef CAS PubMed.
  57. A. V. Zhukhovitskiy, M. G. Mavros, T. Van Voorhis and J. A. Johnson, J. Am. Chem. Soc., 2013, 135, 7418–7421 CrossRef CAS PubMed.
  58. H. Kinuta, M. Tobisu and N. Chatani, J. Am. Chem. Soc., 2015, 137, 1593–1600 CrossRef CAS PubMed.
  59. M. Micksch, M. Tenne and T. Strassner, Eur. J. Org. Chem., 2013, 6137–6145 CrossRef CAS.
  60. M. Cigl, A. Bubnov, M. Kašpar, F. Hampl, V. Hamplová, O. Pacherová and J. Svoboda, J. Mater. Chem. C, 2016, 4, 5326–5333 RSC.
  61. H. Meyer, Monatsh. Chem., 1904, 25, 1201–1214 CrossRef CAS.
  62. H. Nishioka, X. Liang, T. Kato and H. Asanuma, Angew. Chem., Int. Ed., 2012, 51, 1165–1168 CrossRef CAS PubMed.
  63. P. F. Ranken and B. G. McKinnie, J. Org. Chem., 1989, 54, 2985–2988 CrossRef CAS.
  64. P. J. Rayner, P. Norcott, K. M. Appleby, W. Iali, R. O. John, S. J. Hart, A. C. Whitwood and S. B. Duckett, Nat. Commun., 2018, 9, 4251 CrossRef PubMed.
  65. C. R. Martinez and B. L. Iverson, Chem. Sci., 2012, 3, 2191–2201 RSC.
  66. C. Kaub, S. Lebedkin, S. Bestgen, R. Köppe, M. M. Kappes and P. W. Roesky, Chem. Commun., 2017, 53, 9578–9581 RSC.
  67. C. Kaub, S. Lebedkin, A. Li, S. V. Kruppa, P. H. Strebert, M. M. Kappes, C. Riehn and P. W. Roesky, Chem. – Eur. J., 2018, 24, 6094–6104 CrossRef CAS PubMed.
  68. TURBOMOLE V7.5 2020, a development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989–2007, TURBOMOLE GmbH, since 2007, available from https://www.turbomole.org.
  69. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  70. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1997, 78, 1396–1396 CrossRef CAS.
  71. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
  72. S. Grimme, S. Ehrlich and L. Goerigk, J. Comput. Chem., 2011, 32, 1456–1465 CrossRef CAS PubMed.
  73. X. Gui, C. Holzer and W. Klopper, J. Chem. Theory Comput., 2018, 14, 2127–2136 CrossRef CAS PubMed.
  74. K. Krause and W. Klopper, J. Comput. Chem., 2017, 38, 383–388 CrossRef CAS PubMed.
  75. A. Klamt and G. Schüürmann, J. Chem. Soc., Perkin Trans. 2, 1993, 799–805 RSC.

Footnote

Electronic supplementary information (ESI) available. CCDC 2120940–2120944. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/d1dt03795a

This journal is © The Royal Society of Chemistry 2022
Click here to see how this site uses Cookies. View our privacy policy here.