DOI:
10.1039/D3DT03498D
(Paper)
Dalton Trans., 2024,
53, 4705-4718
Neutral 2-phenylbenzimidazole-based iridium(III) complexes with picolinate ancillary ligand: tuning the emission properties by manipulating the substituent on the benzimidazole ring†
Received
20th October 2023
, Accepted 15th December 2023
First published on 16th February 2024
Abstract
We report the synthesis and characterization of ten neutral bisheteroleptic iridium(III) complexes with 2-phenylbenzimidazole cyclometallating ligand and picolinate as ancillary ligand. The 2-phenylbenzimidazole has been modified by selected substituents introduced on the cyclometallating ring and/or on the benzimidazole moiety. The integrity of the complexes has been assessed by NMR spectroscopy, by high-resolution mass spectrometry and by elemental analysis. The complexes are demonstrated to be highly phosphorescent at room temperature and a luminescence study with comprehensive ab initio calculations allow us to determine the lowest emitting excited state which depends on the substituent nature and its position on the cyclometallating ligand.
Introduction
Organic light-emitting diodes (OLEDs) and light-emitting electrochemical cells represent very interesting technologies for lighting displays as such devices are able to work at low voltage.1–3 In these technologies, the excitons generated by the recombination of injected holes and electrons are in both singlet and triplet excited states, with a ratio of 1
:
3, making the theoretical external quantum efficiency (EQE) of only 25% for pure organic devices that can emit solely from the singlet excited state. The seminal work of Thompson and Forrest has demonstrated that phosphorescent emitters are able to convert a singlet exciton to a triplet one, therefore offering the possibility to harvest 100% of the exciton and raise the theoretical EQE to unity.4–6 Thus, since the early 2000s the search for highly emissive and color-tunable transition metal-based emitters has shown an impressive boom. Among the transition metal complexes, two metals display excellent potentials in lighting displays with complexes displaying very high quantum yields, relatively short lifetimes and high emission energy tunability, these being octahedral Ir(III) and square planar Pt(II).7 Those emission properties have been brought to light thanks to cyclometallation. Indeed, the metal–carbon bond with the strong σ donor ability from C−, along with the π-acceptor ability of pyridine, gives a very strong ligand field to these metals, leading to the abovementioned tremendous photophysical properties of the lowest excited state. Consequently, cyclometallated Ir(III) and Pt(II) complexes are studied or used in numerous applications, spanning from triplet emitters in electroluminescent devices,3,8–13 sensors,14–18 theragnostic and/or therapeutic agents19–24 to photosensitizers and photocatalysts14,25,26 to name a few examples.
Neutral Ir(III) complexes have been particularly studied and can be divided into three main types: tris-homoleptic fac/mer-Ir(C^N)3, where C is a cyclometallated carbon and N is a heterocyclic nitrogen; bis-heteroleptic Ir(C^N)2(LX), where LX represents an anionic ancillary ligand; and tris-heteroleptic of the form Ir(N^C^N)(C^N)X, where X is an anionic ligand, typically a chloride.27–30 The emission properties of Ir(III) complexes are often an intriguing interplay of emissive excited states, taking as reference the well-known fac-Ir(ppy)3 (hereafter denoted simply Ir(ppy)3, where ppy = 2-phenylpyridine). The lowest-energy absorption is of 1MLCT (metal-to-ligand charge transfer) nature and likewise the emissive level is recognized to be of 3MLCT nature (λem = 508 nm, τ = 1.6 μs in MeTHF at r.t.).31,32 Higher-lying excited states of 3IL (intra-ligand or ligand-centred) nature are also present and in several cases the energy separation with the triplet MLCT excited state is rather narrow or even “inverted”, 3IL being the lowest excited state. This is the case for the heteroleptic complex (thpy)2Ir(acac) (thpy = thienylpyridine, acac = acetylacetonate) displaying an r.t. emission at λem = 562 nm with τ = 5.3 μs in MeTHF, which is recognized to be a genuine 3IL emitter.28 In addition, the lowest-lying excited state of bis-heteroleptic Ir(III) complexes can also be the ligand-to-ligand charge transfer (3LL′CT, L′ = ancillary ligands) excited state. The 3CT state radiative deactivation results in a broad emission profile and goes along with a rigidochromism effect at low temperature (hypsochromic shift), while the 3IL radiative deactivation results in a structured emission profile and no rigidochromism is observed and even a bathochromic shift can be observed.12,33–36 In addition, the nature of the emitting excited states will also affect the radiative constant (kr), which is typically of the order of 2 × 105 s−1 when the emission emanates from 3MLCT/3LL′CT excited state and lower in the case of 3IL phosphorescence.36,37 However, frequently, cyclometallated Ir(III) complexes demonstrate an emission being a mixture of the 3MLCT and 3LC excited states.
The majority of the reported Ir(III) complexes, as for the Pt(II) ones, are derived directly from the introduction of substituent(s) on the 2-phenylpyridine ligand, and their photophysical properties are well established. On the other hand, complexes based on 2-phenylbenzimidazole as cyclometallating ligand represent a smaller family, but are not devoid of interest. Numerous host materials for phosphors are based on benzimidazole heterocycles for OLEDs regarding their good electron mobility with excellent thermal stability.38–41 From a synthesis point of view, this ligand is an attractive scaffold for cyclometallating ligands, as it presents three divergence points which can be independently modified: the introduction of alkyl or aryl can be performed on the secondary amine, on the phenyl ring or on the benzimidazole ring, and the synthesis does not require the use of palladium-catalysed cross-coupling reactions.42,43 Fine tuning of the emission properties has been achieved by the introduction of electron withdrawing/donating groups on the cyclometallating arene,41,44–56 while the modification of the benzimidazole moiety has been performed by ring expansion.57 For example, the introduction of –OCH3 and CN groups on the cyclometallating phenyl ring allowed tuning of the luminescence from 496 nm to 605 nm with quantum yield from 0.05 to unity.47 Recently, we focused our effort toward the modification of this moiety by the introduction of chosen substituents leading to highly emissive cationic Ir(III) complexes and the luminescence and electrochemical properties have been successfully tuned.58 In addition, we demonstrated that two complexes had emitting excited state that was sensitive to the solvent polarity and it was possible to switch from 3M/LLCT* to 3LC*. Herein, we report a series of neutral Ir(III) complexes featuring 2-phenylbenzimidazole cyclometallating (N^C) ligand and picolinate as ancillary ligand. The N^C ligands (Scheme 1) are designed to study the influence of the substituents’ (Cl, CF3, and OCH3) electron withdrawing/donating ability by tailoring their localization on the ligand, on either the phenyl or the benzimidazole or both through the synthesis of position isomers. It must be emphasized that the HOMO is usually localized on the Ir-ph moiety and the LUMO on the benzimidazole moiety.44,56 The experimental data are successively confronted with state-of-the-art computational methods leading to unambiguous attribution of the emitting exciting state.
 |
| Scheme 1 Top: proligand structures. Bottom: synthesis of the complexes. (i) EtOEtOH/H2O reflux; (ii) 2-picolinic acid, Na2C2O3 100 °C. | |
Results and discussion
Synthesis and characterization
The cyclometallating ligands (HLn, Scheme 1) and μ-dichloridodimers were synthesized following our previous report.59 The ten new complexes IrLn2 were obtained by reacting an excess of picolinic acid with adequate μ-dichloridodimer in the presence of sodium bicarbonate in a mixture of 2-ethoxyethanol/water at 100 °C overnight. After precipitation by water addition and filtration, the solids were purified by flash column chromatography on silica gel using mixtures of dichloromethane/methanol/triethylamine as eluent. All the complexes were characterized by 1H, 13C and 19F (when applicable) NMR, by HRMS and elemental analysis.
Crystallographic quality single crystals of IrL62, IrL92 and IrL102 have been obtained by slow vapor diffusion of diethyl ether or pentane in a concentrated solution of each complex in dichloromethane. The cell parameters of each complex are summarized in Table S1† and selected bond lengths and angles of the three complexes are presented in Table 1, along with those of complex [Ir(ppy)2pic]60 for comparison purpose. The crystallization space groups and Bravais lattices are monoclinic P21/m for IrL62 and IrL92, and triclinic P
for IrL102. Each asymmetric unit displays a single complex: four complexes are present in the unit cell for IrL62 and IrL92 and two in the case of IrL102. As expected, we observed in the lattice the two Δ and Λ isomers which arise from the reaction of the picolinate with the μ-dichlorido-bridged Ir dimer having the D2 symmetric ΔΔ and ΛΛ racemic mixture, where the two C^N ligands have a cis-C,C and trans-N,N configuration around the metal center.61–63 The resulting configuration for the three complexes is the expected mer-N3 cis-C,C trans-N,N within the two Δ and Λ isomers (Fig. 1). The reason for this outcome is due to the Ir–Ir distances being rather small below 4 Å, which leads to important steric hindrance and to the so-called trans effect of the Ir–C bonds, which induces preferential labilization of the bonds located in trans configuration. It results in the stereochemical positioning of Ir–C and Ir–N bonds trans to one another.27,61 The Ir–C and Ir–NC^N bond lengths displayed by the three complexes are similar, ranging from 1.995 to 2.020 Å, and so are the bond angles around the Ir core. The trans effect emanating from the strong σ-donating ability of the cyclometallating carbon affects the Ir–Npic bond lengths.34 The latter are roughly 2.13 Å for the three structures, much longer than the other Ir–NC^N lengths which are of the order of 2.04 Å. The bite angles of both cyclometallating ligand and ancillary ligand are similar through the series and comparable with that of [Ir(ppy)2pic]; the bite angles of ppy ligand and 2-phenylbenzimidazole are around 80°.60 The C–Ir–C′ angles are about 90.5° for the three complexes, of the same order as the one encountered in [Ir(ppy)2pic] (88.7°). Brought together, the different parameters are coherent with the expected octahedral coordination geometry, with slight distortions of the ligands caused by the formation of the five-membered metallacycles.34,64–67 A particularity observed for the three complexes is the strong interaction between the hydrogen atoms from CH3CHCH3 and the ortho H from the benzimidazole ring: these two atoms display a distance very inferior to 2.29 Å corresponding to the sum of the van der Waals radii (vdW), ranging from 1.913 Å to 2.055 Å. These strong interactions have been observed before in a cationic series of iridium complexes,59 and they could be the origin of the strong distortions observed in the cyclometallating ligand (∼13°), larger than the ones observed in [Ir(ppy)2pic] (∼4°). It is worth noting that these interactions are also observed in solution at room temperature, notably on the 1H NMR spectrum, as the iso-propyl's methyl groups are not equivalent, even at higher temperature, and the central H is observed at lower field than the expected chemical shift (3–4.5 ppm) in all the complexes, with multiplets appearing between 5.8 ppm and 5.56 ppm.
 |
| Fig. 1 X-ray molecular structure of complexes IrL62, IrL92 and IrL102. Thermal ellipsoids are plotted at the 50% probability level. H atoms and solvent molecules are omitted for clarity. | |
Table 1 Some relevant bonding and angle parameters for complexes IrL62pic, IrL92pic and IrL102pic along with [Ir(ppy)2(pic)]60 for comparison purpose
Complex |
[Ir(ppy)2(pic)]60 |
IrL62
|
IrL92
|
IrL102
|
Distortion C^N is defined by the angle between the mean planes of the benzimidazole moiety and the phenyl and distortion N^O (picolinato ligand) by the angle of the mean planes of the COO function and the corresponding pyridine. |
Ir–C (Å) |
2.003(6), 2.012(5) |
1.995(5), 2.020(5) |
2.003(5), 2.006(5) |
1.995(2), 2.009(2) |
Ir–NC^N (Å) |
2.041(5), 2.052(5) |
2.043(3), 2.049(3) |
2.031(3), 2.037(3) |
2.028(2), 2.031(2) |
2.031(3), 2.037(3) |
Ir–NN^O (Å) |
2.141(5) |
2.130(3) |
2.121(3) |
2.134(2) |
Ir–ON^O (Å) |
2.156(4) |
2.147(3) |
2.151(3) |
2.157(2) |
C–Ir–C′ (°) |
88.7(2) |
90.5(2) |
91.7(2) |
89.37(8) |
NC^N–Ir–NC^N (°) |
175.7(2) |
172.6(1) |
172.5(1) |
173.56(8) |
CγC^N–Ir–ON^O (°) |
95.4(2) |
94.8(2) |
93.5(2) |
97.93(7) |
CδC^N–Ir–NN^O (°) |
100.1(2) |
97.8(2) |
97.8(2) |
96.01(7) |
N^O bite angle (°) |
77.1(2) |
76.9(1) |
77.1(1) |
76.81(6) |
C^N bite angle (°) |
80.1(2), 81.3(2) |
79.2(2), 79.6(2) |
79.6(2), 79.6(2) |
79.14(7), 79.62(7) |
Distortion C^N (°) |
2.60, 5.71 |
15.18, 17.30 |
12.34, 12.90 |
8.95, 10.45 |
Distortion N^O (°) |
4.54 |
6.74 |
3.60 |
4.40 |
H30–H19 H3–H14 (Å) |
— |
2.055 2.025 |
2.048, 1.957 |
1.943, 1.913 |
One can notice that the crystal packings of the three complexes display several hydrogen bonds between adjacent complex molecules, involving the oxygen atoms of the picolinate ligand and with distances ranging from 2.397 to 2.679 Å that are inferior to the vdW sum (∑vdW(HAr–O) = 2.61 Å and ∑vdW(HAl–O) = 2.72 Å). Other hydrogen bonds are present involving fluorine atoms from the CF3 group and H⋯π interactions are also present. These interactions are characterized considering the distances H–X that are inferior to the vdW sum (∑vdW(HAr–CAr) = 2.79 Å, ∑vdW(HAr–F) = 2.56 Å, ∑vdW(HAl–F) = 2.67 Å, and ∑vdW(HAl–CAr) = 2.90 Å). The supramolecular bonds seem to be mainly driven by electrostatic interaction, with the exception of the abovementioned interaction between H30 and H19 that is due to structural hindrance.68,69
From a computational point of view, the relaxed ground state structures are in very good agreement with respect to experiment. For instance, for complex IrL62, the averaged Ir–C and Ir–NC^N computed (experimental) values are 2.002 (2.019) and 2.051 (2.039) Å. Also, the Ir–NN^O and Ir–O–N^O are simulated at 2.150 and 2.171 Å matching well the observed ones using XRD (2.130 and 2.147 Å). These results gave us confidence for the rationalization of the ground and excited state optoelectronic properties.
The electrochemical properties of the complexes have been studied by cyclic voltammetry (CV) in a deaerated 10−2 M solution of n-NBu4PF6 in MeCN as supporting electrolyte, using vitreous carbon as working electrode (5 mm) and Ag/AgNO3 (10−2 M) as reference electrode at a scan rate of 100 mV s−1. The redox potentials are given versus the reference electrode. CV traces are shown in Fig. S1,† and values are gathered in Table 2. In agreement with previous work on similar complexes, the oxidation peaks in the range 0.55–0.94 V are ascribed to the IrIII/IrIV couple, whereas the reduction affects principally the cyclometallating ligand.47,70 Complex IrL12 displays Ered at −2.38 V and Eox at 0.61 V, whereas the parent complex [Ir(ppy)2pic] has a smaller ΔEredox with Ered at −2.27 V and a Eox at 0.66 V.71 The differences can be explained by the fact that 2-phenylbenzimidazole is more electron rich than 2-phenylpyridine, which leads IrL12 to be more easily oxidized (i.e. the metal center is easier to oxidize) and, consequently, more difficult to be reduced. As expected, the electrochemical properties of the complexes are sensitive to the nature of the substituents, on both the benzimidazole and the phenyl moieties. In reduction, most of the complexes display a reversible reduction wave, with the exception of four of the complexes, whose cyclometallating ligands are substituted by chlorine atoms (IrL22) and by CF3 group on the phenyl ring (IrL6,8,92). The introduction of electron withdrawing groups (Cl and CF3) on the benzimidazole moiety shifts the reduction to less negative potential (IrL22, Ered = −2.20 V and IrL42, Ered = −2.34 V) in comparison with IrL12. In contrast, the introduction of electron donating groups (OMe) solely, on either the benzimidazole and/or the phenyl moieties, does not induce a decrease of the reduction potential (IrL3,5,72, Ered = −2.38 V) with respect to IrL12. In the case of the position isomers substituted both by CF3 and by OMe groups on the cyclometallating ligand, the influence of the electron-donating group prevails on the reduction potential, albeit the reduction peaks of complexes IrL6,9,82 are irreversible. Such a behaviour has been previously described.59 In oxidation, all the complexes display a reversible peak whose potential is dependent on the substituent. As expected, the electron withdrawing CF3 group and chlorine atoms lead to a more positive Eox in comparison with IrL12. The influence is greater when the group is on the phenyl rather than on the benzimidazole. This agrees with the fact that the HOMO is localized on the Ir-phenyl moiety.47,72,73 Similarly, the electron donating OMe group, both on the benzimidazole and the phenyl, leads to a decrease in Eox (vs.IrL12). However, one can notice that the substitution by CF3 groups on both “sides” of the cyclometallating ligand has a synergetic effect on Eox (IrL42, Eox = 0.70 V; and IrL82, Eox = 0.94 V); the synergy is less effective in the case of OMe (IrL32, Eox = 0.57 V; and IrL72, Eox = 0.55 V) as the potentials with one or two MeO groups are almost identical. The presence of both OMe and CF3 groups on the cyclometallating ligand demonstrates the prevalence of the electron withdrawing group over the electron donating group, as shown by the Eox of 0.80 V and 0.71 V of IrL92 and IrL102, respectively. One should notice that the incorporation of two OMe moieties for the complex IrL72 induces a less positive oxidation potential than for the other complexes. It should be also noticed that, as expected, all complexes possessing such a donating group have a HOMO partly localized on the OMe moiety. The cyclometallating ligand 2-phenylbenzimidazole, used instead of the most encountered 2-phenylpyridine (ppy), has a substantial effect on the electrochemical properties of complexes in comparison with [Ir(ppy)2pic].
Table 2 Redox potentials of complexes IrLn2pic: E (V) vs. Ag/AgNO3 (0.01 M) in deaerated CH3CN
Complex |
E
Red (V) |
E
Ox (V) |
irr denotes irreversible reduction peak. |
IrL12
|
−2.38 |
0.61 |
IrL22
|
−2.20irr |
0.71 |
IrL32
|
−2.38 |
0.57 |
IrL42
|
−2.34 |
0.70 |
IrL52
|
−2.38 |
0.60 |
IrL62
|
−2.42irr |
0.84 |
IrL72
|
−2.38 |
0.55 |
IrL82
|
−2.21irr |
0.94 |
IrL92
|
−2.41irr |
0.80 |
IrL102
|
−2.34 |
0.71 |
Photophysical properties
Absorption spectroscopy.
The absorption spectra of the complexes have been registered in dilute solution of CH2Cl2 at room temperature. They are displayed in Fig. 2 and data are gathered in Table 3 (individual absorption spectra are presented in Fig. S2†). The intense bands at around 300 nm can be ascribed to ligand-centred (LC) π–π* transitions from the cyclometallating and ancillary ligands. Broad and relatively weak absorption bands observed in the longer wavelength region, over 350 nm, are attributed to the overlap of metal-to-ligand charge transfer (MLCT) and ligand-to-ligand charge transfer (LLCT) and those over roughly 460 nm to direct absorption from singlet ground state to triplet excited state (Fig. S3†), as a consequence of the strong spin orbit coupling effect exerted by the Ir(III) core.7,26,74–76 For instance, the less intense lowest-lying bands, displayed as weak tails in the absorption spectra, roughly around 430 nm, are ascribed to spin-forbidden triplet transitions. As expected, a focus on the CT absorption band wavelengths evidences the influence of the substituents: for instance, electron withdrawing groups (CF3 and Cl) induce a hypsochromic shift and, in contrast, the electron donating methoxy group induces a bathochromic shift. It seems that the influence of the substituent on the absorption band energies does not greatly depend on the position on the cyclometallating ligand. To assign the observed absorption bands, TD-DFT computations were conducted on the relaxed ground state geometries. The simulated spectra, along with the band assignments, have been compiled in the ESI (Table S2 and Fig. S32†). It should be noted that the simulated spectra match well the experimental trends.
 |
| Fig. 2 Absorption spectra of the ten complexes in CH2Cl2 at room temperature. | |
Table 3 Photophysical properties
|
Photoluminescence in dilute solution of CH2Cl2, 298 Ka,b |
Photoluminescence at 77 Kc |
Complex |
Absorption λ [nm] (ε × 103 [M−1 cm−1]) |
λ [nm] |
Φ (air) |
τ [μs] (air) |
k
r × 105 [s−1] |
k
nr × 105 [s−1] |
k[O2]d × 109 [M−1 s−1] |
E
00
(eV) |
λ [nm] |
τ [μs] |
Ir(ppy)3 in CH2Cl2 was used as a reference.
In deaerated solution unless otherwise mentioned.
Recorded in butyronitrile.
With [O2] = 2.2 mM in dichloromethane.
Energy of the emitting excited state.
The most intense band.
|
IrL12
|
298 (31.9), 310 (31.8), 345 (11.9), 375 (8.8), 402 (7.3), 429 (4.8), 480 (1.0) |
503, 537f |
0.09 (0.02) |
0.32 (0.08) |
2.8 |
28.4 |
5.0 |
2.47 |
486f, 521, 567 |
3.31 |
IrL22
|
309 (28.3), 322 (35.4), 350 (11.8), 382 (8.2), 413 (7.7), 441 (5.1), 488 (0.8) |
507f, 542, 587sh |
0.35 (0.03) |
1.10 (0.12) |
3.5 |
5.9 |
4.4 |
2.45 |
490f, 530, 573, 627sh |
2.94 |
IrL32
|
304 (31.4), 316 (36.7), 347 (15.8), 374 (11.1), 403 (8.4), 431 (5.1), 493 (0.3) |
516, 552f, 600sh |
0.11 (0.01) |
0.89 (0.10) |
1.2 |
10.0 |
5.1 |
2.40 |
500f, 537, 584, 640sh |
4.10 |
IrL42
|
297 (28.8), 309 (29.7), 343 (10.7), 376 (8.4), 408 (6.7), 442 (3.5), 484 (0.7) |
502f, 532, 580sh |
0.42 (0.03) |
1.21 (0.12) |
3.5 |
4.8 |
4.9 |
2.47 |
488f, 521, 563, 616sh |
3.49 |
IrL52
|
299 (40.1), 311 (40.8), 336 (20.0), 369 (14.7), 410 (8.4), 468 (1.1) |
563 |
0.04 (0.01) |
0.32 (0.08) |
1.2 |
30.0 |
4.3 |
2.20 |
471, 507f, 547 |
4.58 |
IrL62
|
302 (36.3), 314 (33.9), 357 (12.5), 387 (9.9), 408 (7.7), 449 (3.7), 496 (1.0) |
514f, 554, 605sh |
0.40 (0.03) |
1.02 (0.19) |
3.9 |
5.9 |
5.5 |
2.41 |
505f, 544, 589 |
4.40 |
IrL72
|
302 (30.8), 313 (33.3), 341 (19.9), 362 (15.9), 389 (10.7), 410 (8.4), 457 (0.8) |
500sh, 560f |
0.30 (0.04) |
1.07 (0.19) |
2.8 |
6.5 |
2.8 |
2.48 |
486f, 526, 569, 621sh |
6.75 |
IrL82
|
298 (29.4), 312 (28.7), 358 (8.8), 387 (7.6), 412 (5.8), 462 (2.3), 491 (0.6) |
516f, 554, 600sh |
0.43 (0.05) |
1.91 (0.23) |
2.2 |
3.0 |
1.8 |
2.40 |
500f, 541, 584 |
4.31 |
IrL92
|
309 (31.9), 322 (35.0), 356 (14.9), 384 (11.2), 412 (7.8), 446 (3.7), 491 (0.4) |
485, 519f, 554sh |
0.17 (0.02) |
0.78 (0.10) |
2.2 |
10.6 |
4.4 |
2.56 |
470f, 506, 544 |
3.35 |
IrL102
|
299 (29.7), 313 (30.8), 337 (15.8), 367 (11.7), 395 (7.9), 416 (6.3), 472 (0.6) |
536f, 578, 627sh |
0.24 (0.03) |
0.91 (0.08) |
2.6 |
8.3 |
3.5 |
2.31 |
519f, 562, 612 |
4.04 |
The primary transition is a mixture of MLCT and LLCT (L = phenylbenzimidazole) in all complexes except for IrL62 and IrL82. Additionally, all complexes exhibit a weak initial transition (with a small oscillator strength) that corresponds to ML′CT and LL′CT (charge towards the picolinate moieties), except for IrL72 and IrL102 which exhibit two rather strong transitions that are close in energy. IrL52 is the only complex that exhibits two very weak transitions corresponding to ML′CT and LL′CT before the primary transition (MLCT and LLCT). Moreover, this strong transition is approximately 0.20 eV higher in energy than in the other complexes. Complex IrL72 also exhibits such a pattern. These differences compared to other complexes may explain a distinct excited state energy order for these complexes.
Emission spectroscopy.
The emission spectra of the complexes have been recorded in both deaerated and air-equilibrated dilute solution of CH2Cl2 at room temperature and at 77 K in butyronitrile rigid matrix. The room temperature spectra are displayed in Fig. 3. The photophysical data are gathered in Table 3 and the individual spectra are to be found in ESI† as well as those at 77 K. At r.t., the complexes display structured emission spectra (with the exception of IrL52 and IrL72 whose emission spectra are broad) in the visible range, with lifetimes of μs, with large Stokes shifts, and sensitivity to the presence of oxygen (k[O2] ranging from 1.8 M−1 s−1 to 5.5 M−1 s−1). Therefore the emission can be ascribed to phosphorescence as expected for this family of iridium(III) complexes. The photoluminescence quantum yields range from 0.04 to 0.43 in deaerated CH2Cl2. The photophysical properties of (ppy)2Irpic77 are worth comparing with those of IrL12. (ppy)2Irpic displays an emission at 505 nm (Φ = 0.15, τ = 514 ns, deaerated) in CH2Cl2 which is comparable with IrL12, but the use of 2-phenylbenzimidazole instead of 2-phenylpyridine as cyclometallating ligand seems to induce a broadening of the emission band with a concomitant more pronounced vibronic progression.77 Albeit, in both complexes the nature of the emitting excited state is predominantly 3IL, in view of the slight rigidochromism (∼10 nm) observed at low temperature.
 |
| Fig. 3 Emission spectra of the ten complexes in dilute solution of CH2Cl2 at room temperature. | |
The introduction of electron withdrawing group solely (i.e. CF3 and Cl) either on the benzimidazole (IrL22 and IrL42) or on the cyclometallating ring (IrL62) does not induce a pronounced hypsochromic shift in comparison with IrL12. As a consequence, IrL42 and IrL22 display emission spectra in the same range as IrL12, whereas IrL62 displays a slight bathochromic shift (E00 = 2.20 eV). Along these four complexes the shape of the spectra changes by getting more structured, which is accompanied by an increase of kr (i.e. radiative rate constant) from 2.8 × 105 s−1 to 3.9 × 105 s−1 for IrL12 and IrL62 respectively. On the other hand, the introduction of electron donating group solely on either the benzimidazole (IrL32) or the cyclometallating ring (IrL52) induces a slight bathochromic shift in comparison with IrL12, with E00 of 2.47 eV, 2.40 eV and 2.20 eV for IrL12, IrL32, and IrL52 respectively. The kr is 1.2 × 105 s−1 for both IrL32 and IrL52, smaller than kr (2.8 × 105 s−1) observed for IrL12 and, in addition, IrL52 displays a broad emission.
The substitution by electron withdrawing (IrL82) or by electron donating (IrL72) groups on both “sides” of the 2-phenylbenzimidazole does induce slight changes in the emission energy. The presence of two CF3 groups in IrL82 draws a hypsochromic shift of the emission (in accordance with the E00 of 2.40 eV) and the radiative rate constant decreases slightly (2.2 × 105 s−1), in comparison with IrL12. On the other hand, a similar substitution by MeO groups for IrL72 has quite a dramatic consequence on the emission spectrum shape, which becomes structureless and displays a bathochromic shift with E00 of 2.21 eV. One can notice that kr are similar for these two complexes, 2.8 × 105 s−1. Finally, the last two position isomers, IrL92 and IrL102, substituted by antagonist functional groups in different positions, CF3 and OMe, exhibit distinct emission spectra. Both spectra are structured with similar radiative constants, 2.2 × 105 s−1 and 2.6 × 105 s−1, respectively, for IrL92 and IrL102, that are less than that observed for IrL12 (kr = 2.8 × 105 s−1), particularly for IrL92. The impact of the position of the substituent is dramatic when looking at the emission energy: while IrL92 displays a hypsochromic shift, with E00 of 2.56 eV, the IrL102 spectrum is red shifted, with E00 of 2.31 eV, both with respect to IrL12 (2.47 eV).
The 3MLCT or 3LC nature of the phosphorescence-emitting excited states can be experimentally assessed by the photophysical properties.37,78 For instance, a typical 3MLCT emission like that of fac-Ir(ppy)3 will display a radiative constant of around 2 × 105 s−1 (τ ∼2 μs), and for the case of a pure 3LC emission the radiative constant will be smaller, like that of (thpy)2Ir(acac)79 (thpy = 2-(2-pyridyl)benzothiophene; acac = acetylacetonate) which displays a kr of 0.2 × 105 s−1 (τ ∼5.3 μs). Within the series, the radiative constants range from 1.2 × 105 s−1 to 3.9 × 105 s−1, which seems to indicate that most of the complexes display an emission emanating from the radiative deactivation of a lowest 3MLCT state, with the exception of IrL32 and IrL52, whose radiative constants are significantly lower than 2 × 105 s−1. However, kr is not the only parameter that allows one to characterize the nature of the lowest excited state and other experimental parameters have to be considered: (i) the shape of the emission spectrum which is structureless and broad for an emission emanating from a 3MLCT, (ii) the rigidochromic effect at low temperature, and (iii) the linear relationship80 for 3MLCT emission between the emission energy and ΔE1/2 = (Eox − Ered) eV. Thus, with regard to the shape of the spectrum, only IrL52 and IrL72 display a broad and structureless emission which is characteristic of a 3MLCT emission. The rigidochromism is a property of the transition metal complexes with an emission emanating from the radiative deactivation of 3CT excited state, typically 3MLCT, displaying a blue-shifted emission when the solution environment becomes rigid, by lowering the temperature in solution.33 The superimposed spectra at r.t. and in benzonitrile at 77 K are presented in Fig. S4.† Looking at the spectra, two features are observed, being a marked rigidochromism and a slight one.
Complexes IrL52 and IrL62 are very representative and the superimposed spectra are shown in Fig. 4. It is somewhat striking that the emission spectra at r.t. in CH2Cl2 and at 77 K in butyronitrile are almost superimposed for IrL62. Within the series, only IrL52 and IrL72 exhibit broad emission spectra at r.t., displaying a strong rigidochromic effect and a change in the spectra shape. Finally, a linear relation between the emission energy and ΔE1/2 is not verified within the series, which rules out an emission emanating from the pure 3MLCT for this family of complexes. To conclude, while the excited states (3IL, 3ILCT and 3MLCT) are known to be very close in iridium(III) complexes, it seems that, in the light of the experimental results, the present series of complexes displays an emission at r.t. with a strong proportion of 3IL excited state, with the exception of IrL52 and IrL72, whose emission at r.t. has strong 3MLCT character. As the primary transition is a mixture of MLCT and LLCT (see above), the complexes rapidly undergo intersystem crossing and the 3MLCT and 3ILCT are populated and a subsequent internal conversion leads to the population of the 3IL excited state.
 |
| Fig. 4 Superimposition of the emission spectra at room temperature (full line) in CH2Cl2 and at 77 K (dashed line) in butyronitrile of complexes IrL52 (left) and IrL62 (right). | |
To gain a deeper understanding of the phosphorescence properties, we performed optimizations of the first triplet excited state and simulations of the luminescence resulting from this state. As shown in Fig. S33 of the ESI,† our simulated spectra match the recorded spectra at room temperature. The relative intensities and peak positions are well reproduced in our simulations, and we accurately replicate the spacing between the peaks in cases where there is sufficient vibronic coupling. Additionally, our simulations capture the bell-curve shape observed in the luminescence spectra of IrL52 and IrL72, further validating our model. The correlation between the simulated and observed luminescence at room temperature is shown in Fig. S34.† This high level of agreement allows us to confidently localize the electrons in the excited state (spin density) and estimate the expected colour using a CIE (x,y) horseshoe diagram. Most of the complexes have a spin density localized over the metal and phenylbenzimidazole moieties (Fig. S35†). However, for IrL52 and, surprisingly, IrL92, the spin density is localized on the metal and the picolinate moieties, with a larger spin density on the picolinate for IrL52 than for IrL92. These findings explain the distinctive behaviour of complex IrL52 compared to the others. Finally, we can compare the predicted colours from our simulations with the observed ones, as shown in Fig. 5. As one can see, our simulations reproduce well the observed colour.
 |
| Fig. 5 Chromaticity diagram CIE-1931: experimental (top) and simulated (bottom) color. | |
Conclusion
We described a series of ten original 2-phenylbenzimidazole-based iridium(III) complexes with picolinate ancillary ligand, which have been characterized by NMR, HRMS and elemental analysis. Their luminescence properties have been studied in dilute solution at room temperature and in butyronitrile at low temperature, both in steady state and in time resolved spectroscopy. We demonstrated that the choice of the substituent on the cyclometallating ligand allows one to finely tune the emission energy of the complexes by manipulating the electronic properties (i.e. Hammett parameter), which also has an influence on the electrochemical properties. Moreover, the nature of the lowest-lying excited state(s) is affected by the substituents and their position, and the emission emanates from the radiative deactivation of 3CT or 3IL excited states. In particular IrL52 can be recognized as a “genuine” 3MLCT emitter whereas IrL62 displays the opposite behaviour being a “true” 3IL emitter, which has been demonstrated by both experimental techniques and state-of-the-art computational methods.
Experimental
Synthesis of the complexes
The crude μ-dichloridodimers were synthetized from HL1–HL10 and IrCl3·nH2O as reported in our previous work.59 In a round-bottom flask under argon, the selected μ-dichloridodimer, 2-picolinic acid and Na2CO3 (1
:
3
:
3) were dissolved in a deaerated 3
:
1 mixture of 2-ethoxyethanol/water and heated at 100 °C overnight. At r.t., water was added, and the precipitate was filtered off, washed with water and diethylether, and dried. The precipitates were purified over pre-treated SiO2 with Et3N using CH2Cl2/CH3OH as eluent.
IrL12
.
Crude [Ir(L1)2(μ-Cl)]2 (0.041 g, 0.06 mmol), 2-picolinic acid (0.022 g, 0.18 mmol), Na2CO3 (0.19 g, 0.18 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH 9
:
1). Product isolated as yellow powder (41 mg, 89%). 1H NMR (500 MHz, CD2Cl2) δ 8.15–8.11 (m, 1H; Hα), 7.99 (dd, J = 8.0, 0.6 Hz, 1H; H10), 7.88–7.86 (m, 2H; Hβ, δ), 7.79 (d, J = 9.0 Hz, 1H; H5′), 7.77 (d, J = 9.1 Hz, 1H; H5), 7.70 (d, J = 8.4 Hz, 1H; H13′), 7.68 (d, J = 7.9 Hz, 1H; H13), 7.32 (td, J = 7.0, 1.5 Hz, 1H; Hγ), 7.31–7.27 (m, 1H, H12), 7.27–7.23 (m, 1H; H11), 7.21 (ddd, J = 8.3, 7.3, 1.0 Hz, 1H, H12′), 6.96 (dt, J = 15.2, 7.6 Hz, 2H, H4′, 4), 6.86 (ddd, J = 8.2, 7.4, 0.9 Hz, 1H; H11′), 6.70 (tdd, J = 7.5, 3.4, 1.2 Hz, 2H; H3′, 3), 6.51 (dd, J = 7.7, 0.9 Hz, 1H; H2′), 6.22 (dd, J = 7.7, 0.9 Hz, 1H; H2), 5.80–5.67 (m, 3H; H10, 2(CH(CH3)2)), 1.86 (ddd, J = 9.8, 6.9, 4.9 Hz, 12H; 2(CH(CH3)2)). 13C NMR (126 MHz, CD2Cl2) δ 173.4, 164.1, 163.1, 154.0, 152.6, 150.5, 149.8, 141.3, 141.2, 137.8, 136.0, 135.54, 135.46, 134.7, 134.2, 133.9, 129.9, 129.6, 128.1, 127.8, 125.7, 125.4, 124.3, 123.7, 123.1, 122.6, 121.7, 121.1, 117.9, 115.2, 114.3, 113.7, 50.3, 50.3, 22.1, 22.1, 21.9, 21.8. HRMS (ESI) found m/z 786.24064, calcd m/z 786.24166 for C38H35IrN5O2 [M-H]+. Elemental analysis calcd (%) for C38H34IrN5O2·CH3OH, C, 57.34; H, 4.69; N, 8.57, found C, 57.09; H, 4.74; N, 8.51.
IrL22
.
Crude [Ir(L2)2(μ-Cl)]2 (0.050 g, 0.06 mmol), 2-picolinic acid (0.022 g, 0.18 mmol), Na2CO3 (0.019 g, 0.18 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). Purification over SiO2 was not possible due to the low solubility of the complex. The complex was washed with H2O, diethylether and methanol, then recrystallized in a CH2Cl2/pentane mixture. The product was isolated as a pale-yellow powder (52 mg, 95%). 1H NMR (500 MHz, CD2Cl2) δ 8.19 (dd, J = 7.8, 0.5 Hz, 1H; Hα), 8.14 (s, 1H; H10), 7.95 (td, J = 7.7, 1.6 Hz, 1H; Hβ), 7.85–7.82 (m, 1H; Hδ), 7.80–7.78 (m, 3H, H13, 13′, H5), 7.77 (d, J = 7.9 Hz, 1H, H5′), 7.40 (ddd, J = 7.5, 5.4, 1.5 Hz, 1H; Hγ), 7.03–6.946 (m, 2H; H4′, 4), 6.78–6.72 (m, 2H; H3′, 3), 6.51 (dd, J = 7.7, 0.8 Hz, 1H; H2′), 6.20 (dd, J = 7.7, 0.8 Hz, 1H; H2), 5.69 (dh, J = 14.1, 6.9 Hz, 2H; 2(CH(CH3)2)), 5.58 (s, 1H, H10′), 1.87–1.78 (m, 12H, 2(CH(CH3)2)). 13C NMR (126 MHz, CD2Cl2) δ 173.2, 166.3, 165.2, 153.9, 152.7, 150.9, 149.9, 140.6, 140.4, 138.4, 135.6, 135.5, 134.7, 134.6, 133.4, 133.1, 130.7, 130.3, 128.6, 128.3, 128.2, 127.9, 127.1, 126.4, 126.2, 125.9, 122.1, 121.5, 118.8, 116.3, 115.3, 115.0, 54.0, 50.8, 50.8, 22.1, 22.0, 21.9, 21.8. HRMS (ESI) found m/z 922.08386, calcd m/z 922.08545 for C38H31Cl4IrN5O2 [M – H]+. Elemental analysis calcd (%) for C38H30Cl4IrN5O2, C, 49.46; H, 3.28; N, 7.59, found C, 49.54; H, 3.51; N, 7.62.
IrL32
.
Crude [Ir(L3)2(μ-Cl)]2 (0.054 g, 0.06 mmol), 2-picolinic acid (0.024 g, 0.19 mmol), Na2CO3 (0.020 g, 0.19 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH) progressive increase of CH3OH from 1% to 6.5%. Product isolated as a yellow powder (46 mg, 78%). 1H NMR (500 MHz, CD2Cl2) δ 8.39 (s, 1H; H10), 8.17 (d, J = 7.7 Hz, 1H; Hα), 7.95–7.87 (m, 2H; Hβ, δ) 7.87–7.75 (m, 4H; H13′, 13, H5′, 5), 7.54 (d, J = 8.4 Hz, 1H; H12), 7.45 (d, J = 8.6 Hz, 1H; H12), 7.41–7.35 (m, 1H; Hγ), 7.05–6.95 (m, 2H; H4′, 4), 6.75 (dd, J = 14.0, 6.9 Hz, 2H; H3′, 3), 6.55 (d, J = 7.6 Hz, 1H; H2), 6.23 (d, J = 7.6 Hz, 1H; H2), 5.82 (s, 1H; H10′) 5.81–5.71 (m, 2H; 2(CH(CH3)2)), 2.02–1.73 (m, J = 6.6 Hz, 12H; 2(CH(CH3)2)). 13C NMR (126 MHz, CD2Cl2) δ 173.1, 166.4, 165.3, 154.0, 152.9, 151.2, 149.9, 140.9, 140.6, 138.3, 136.2, 136.0, 135.7, 135.2, 134.7, 130.7, 130.3, 128.24, 128.19, 128.16, 126.6, 126.4, 126.3, 126.12, 126.07, 126.0, 126.0, 125.8, 125.7, 123.9, 123.7, 122.0, 121.5, 120.00, 119.97, 119.9, 119.5, 119.4, 119.4, 115.52, 115.49, 115.45, 115.42, 114.7, 114.3, 112.71, 112.68, 112.64, 112.60, 50.81, 50.77, 22.15, 22.07, 22.0, 21.9. 19F NMR (470 MHz, CD2Cl2) δ −61.2, −61.5. HRMS (ESI) found m/z 922.21578, calcd m/z 922.21645 for C40H33F6IrN5O2 [M]+. Elemental analysis calcd (%) for C40H32F6IrN5O2, C, 52.16; H, 3.51; N, 7.61, found C, 51.92; H, 3.65; N, 7.75.
IrL42
.
Crude [Ir(L4)2(μ-Cl)]2 (0.067 g, 0.09 mmol), 2-picolinic acid (0.032 g, 0.27 mmol), Na2CO3 (0.028 g, 0.27 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH) progressive increase of CH3OH until 5%. Product isolated as a yellow-orange powder (57 mg, 76%). 1H NMR (500 MHz, CD2Cl2) δ 8.15–8.11 (m, 1H; Hα), 7.90–7.83 (m, 3H; H10, Hβ, δ), 7.73 (d, J = 8.1 Hz, 1H; H5), 7.71 (d, J = 8.1 Hz, 1H; H5′), 7.32 (dd, J = 5.6, 1.6 Hz, 1H, Hγ), 7.15 (d, J = 2.3 Hz, 1H; H13′), 7.12 (d, J = 2.2 Hz, 1H, H13), 6.98–6.88 (m, 3H; H11, H4′, 4,), 6.71–6.65 (m, 2H; H3′, 3), 6.50 (dd, J = 9.0, 2.4 Hz, 1H; H11′), 6.48 (dd, J = 7.7, 1.0 Hz, 1H; H2′), 6.21 (dd, J = 7.7, 0.9 Hz, 1H; H2), 5.69 (pd, J = 13.2, 6.3 Hz, 2H; 2(CH(CH3)2)), 5.56 (d, J = 9.0 Hz, 1H; H10′), 3.86 (s, 3H; –OCH3), 3.83 (s, 3H; –OCH3), 1.88–1.79 (m, 12H; 2(CH(CH3)2)). 13C NMR (126 MHz, CD2Cl2) δ 173.35, 163.58, 162.46, 156.79, 156.33, 154.00, 152.02, 149.79, 137.76, 136.33, 135.89, 135.77, 135.72, 135.28, 134.88, 134.62, 129.51, 129.19, 128.00, 127.73, 125.18, 124.95, 121.57, 120.98, 118.38, 115.68, 112.60, 112.19, 98.50, 98.19, 56.51, 56.48, 50.07, 21.85, 21.84, 21.72, 21.63. HRMS (ESI) found m/z 846.26208, calcd m/z 846.26281 for C40H39IrN5O4 [M − H]+. Elemental analysis calcd (%) for C40H38IrN5O4·H2O, C, 55.67; H, 4.67; N, 8.12, found C, 55.81; H, 4.75; N 7.98.
IrL52
.
Crude [Ir(L5)2(μ-Cl)]2 (0.049 g, 0.06 mmol), 2-picolinic acid (0.022 g, 0.18 mmol), Na2CO3 (0.019 g, 0.18 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH/Et3N) progressive increase of CH3OH from 1% to 6.5%. Product isolated as a bright yellow powder (29 mg, 54%). 1H NMR (500 MHz, CD2Cl2) δ 8.17 (dd, J = 7.8, 0.5 Hz, 1H; Hα), 8.02 (d, J = 8.0 Hz, 1H; H10), 7.91 (td, J = 7.7, 1.5 Hz, 1H; Hβ), 7.88–7.78 (m, 3H; Hδ, H5,5′), 7.76 (d, J = 8.4 Hz, 1H; H13), 7.73 (d, J = 8.3 Hz, 1H; H13′), 7.40–7.35 (m, 2H; Hγ,H12), 7.30 (ddd, J = 16.7, 8.3, 1.0 Hz, 2H; H12′, H11), 7.23 (dd, J = 8.3, 1.3 Hz, 1H; H4′), 7.18 (dd, J = 8.3, 1.3 Hz, 1H; H4), 6.94 (t, J = 7.8 Hz, 1H; H11′), 6.62 (d, J = 1.4 Hz, 1H; H2′), 6.30 (d, J = 1.4 Hz, 1H; H2), 5.78–6.54 (m, 3H; 2(CH(CH3)2), H10), 1.90 (dd, J = 9.5, 7.0 Hz, 3H; CH(CH3)2), 1.83 (t, J = 6.9 Hz, 3H; CH(CH3)2).13C NMR (126 MHz, CD2Cl2) δ 173.4, 162.8, 161.9, 153.7, 152.2, 150.5, 149.9, 141.0, 140.9, 139.73, 139.72, 139.44, 139.43, 138.41, 134.1, 133.9, 130.80, 130.78, 130.75, 130.6, 130.3, 130.13, 130.10, 130.08, 128.4, 128.0, 125.53, 125.49, 125.3, 125.2, 125.1, 124.4, 124.1, 123.6, 123.3, 123.2, 118.9, 118.84, 118.81, 118.30, 118.27, 118.24, 118.1, 115.4, 114.6, 114.0, 50.78, 50.77, 22.00, 21.96, 21.9, 21.8. 19F NMR (470 MHz, CD2Cl2) δ −63.3, −63.5. HRMS (ESI) found m/z 922.21583, calcd m/z 922.21645 for C40H33F6IrN5O2 [M-H]+. Elemental analysis calcd (%) for C40H32F6IrN5O2·H2O, C, 51.17; H, 3.65; N, 7.46, found C, 51.28; H, 3.76; N, 7.31.
IrL62
.
Crude [Ir(L6)2(μ-Cl)]2 (0.052 g, 0.07 mmol), 2-picolinic acid (0.025 g, 0.20 mmol), Na2CO3 (0.022 g, 0.020 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH) progressive increase of CH3OH until 5%. Product isolated as an orange-yellowish powder (29 mg, 49%). 1H NMR (500 MHz, CD2Cl2) δ 8.14 (d, J = 7.4 Hz, 1H; Hα), 8.00–7.94 (m, 1H; Hβ), 7.90 (d, J = 5.3 Hz, 1H; H10), 7.87 (td, J = 7.7, 1.5 Hz, 1H; Hδ), 7.73 (d, J = 8.8 Hz, 1H. H5), 7.71 (d, J = 8.8 Hz, 1H; H5′), 7.67–7.62 (m, 2H; H13,13′), 7.36–7.32 (m, 1H; Hγ), 7.28–7.22 (m, 2H; H12, 11), 7.17 (t, J = 7.8 Hz, 1H; H12′), 6.85 (t, J = 7.3 Hz, 1H; H11′), 6.56 (dd, J = 8.7, 2.7 Hz, 1H; H4′), 6.51 (dd, J = 8.7, 2.7 Hz, 1H; H4), 5.94 (d, J = 2.6 Hz, 1H; H2′), 5.70 (d, J = 2.7 Hz, 1H; H2), 5.69 (d, J = 2.8 Hz, 1H; H10′), 5.68–5.56 (m, 2H; 2(CH(CH3)2)), 3.40 (s, 1H; Ph-OCH3), 3.32 (s, 1H; Ph-OCH3), 1.93–1.71 (m, 12H; 2(CH(CH3)2)). 13C NMR (126 MHz, CD2Cl2) δ 173.4, 164.2, 163.1, 160.7, 160.4, 155.3, 153.9, 152.9, 149.8, 141.3, 141.2, 137.8, 134.0, 133.6, 128.5, 128.2, 128.0, 127.7, 126.9, 126.6, 124.2, 123.5, 122.7, 122.2, 119.6, 118.9, 117.6, 114.9, 113.9, 113.3, 108.0, 107.1, 54.9, 54.8, 50.1, 50.0, 21.9, 21.8, 21.7. HRMS (ESI) found m/z 846.26255, calcd m/z 846.26281 for C40H39IrN5O4 [M − H]+. Elemental analysis calcd (%) for C40H38IrN5O4, C, 56.85; H, 4.54; N, 8.29, found C, 56.71; H, 4.78; N, 8.44.
IrL72
.
Crude [Ir(L7)2(μ-Cl)]2 (0.111 g, 0.11 mmol), 2-picolinic acid (0.042 g, 0.34 mmol), Na2CO3 (0.036 g, 0.34 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH) progressive increase from 1% to 5% of CH3OH. Product isolated as a bright yellow powder (90 mg, 74%). 1H NMR (500 MHz, CD2Cl2) δ 8.41 (s, 1H; H10), 8.21 (d, J = 7.7 Hz, 1H; Hα), 7.98–7.91 (m, 2H; Hβ, δ), 7.91–7.81 (m, 4H; H5,5′, H13,13′), 7.62 (d, J = 8.8 Hz, 1H; H12), 7.52 (d, J = 8.8 Hz, 1H; H12′), 7.47–7.40 (m, 1H; Hγ), 7.28 (d, J = 8.3 Hz, 1H; H4′), 7.22 (d, J = 8.3 Hz, 1H; H4), 6.68 (s, 1H; H2′), 6.29 (s, 1H; H2), 5.86 (s, 1H; H10′), 5.74 (hept, J = 13.8, 6.9 Hz, 2H; 2(CH(CH3)2)), 2.04–1.75 (m, 12H; 2(CH(CH3)2)). 13C NMR (126 MHz, CD2Cl2) δ 173.1, 165.1, 164.1, 153.6, 152.3, 151.0, 150.0, 140.6, 140.2, 138.9, 138.57, 138.56, 136.1, 135.9, 131.8, 131.5, 131.3, 131.2, 131.04, 131.02, 131.99, 130.96, 130.7, 130.5, 130.08, 130.05, 130.02, 130.00, 128.5, 128.4, 127.3, 127.0, 126.7, 126.4, 126.2, 125.8, 125.6, 125.3, 125.1, 123.7, 123.5, 123.1, 123.0, 120.95, 120.92, 120.90, 120.87, 120.39, 120.36, 120.34, 119.18, 119.15, 119.12, 118.67, 118.65, 118.62, 118.59, 115.82, 115.78, 115.75, 115.71, 115.27, 114.7, 112.91, 112.88, 112.84, 112.80, 51.30, 51.27, 22.0, 21.89, 21.86. 19F NMR (470 MHz, CD2Cl2) δ −61.5, −61.8, −63.4, −63.6. HRMS (ESI) found m/z 1058.19085, calcd m/z 1058.19124 for C42H31F12IrN5O2 [M – H]+. Elemental analysis calcd (%) for C42H30F12IrN5O2, C, 47.72; H, 2.89; N, 6.63, found C, 47.55; H, 2.75; N, 6.87.
IrL82
.
Crude [Ir(L8)2(μ-Cl)]2 (0.054 g, 0.07 mmol), 2-picolinic acid (0.024 g, 0.20 mmol), Na2CO3 (0.021 g, 0.20 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH) progressive increase of CH3OH until 5%. Product isolated as an orange powder (40 mg, 67%). 1H NMR (500 MHz, CD2Cl2) δ 8.13 (d, J = 7.6, 1H; Hα), 7.86 (ddd, J = 8.9, 7.4, 5.4 Hz, 3H; Hβ, δ,H10), 7.66 (t, J = 8.3 Hz, 2H; H5, 5′), 7.36–7.30 (m, 1H; Hγ), 7.11 (d, J = 2.2 Hz, 1H; H13), 7.08 (d, J = 2.2 Hz, 1H; H13′), 6.88 (dd, J = 9.0, 2.3 Hz, 1H; H11), 6.53 (dd, J = 8.7, 2.7 Hz, 1H; H11′), 6.51–6.46 (m, 2H, H4′,4), 5.92 (d, J = 2.6 Hz, 1H; H2′), 5.69 (t, J = 5.8 Hz, 1H; H2), 5.66–5.56 (m, 2H; 2(CH(CH3)2)), 5.54 (d, J = 9.0 Hz, 1H; H10′) 3.85 (s, 3H; -OCH3), 3.82 (s, 3H; –OCH3), 3.42 (s, 3H; Ph-OCH3), 3.35 (s, 3H; Ph-OCH3), 1.87–1.75 (m, 12H; 2(CH(CH3)2)). 13C NMR (126 MHz, CD2Cl2) δ 173.4, 163.7, 162.6, 160.4, 160.2, 156.5, 156.0, 154.7, 153.9, 152.2, 149.9, 137.7, 135.8, 134.7, 134.4, 128.9, 128.6, 128.0, 127.7, 126.4, 126.2, 119.5, 118.9, 118.0, 115.2, 112.1, 111.7, 107.7, 106.8, 98.5, 98.2, 56.50, 56.47, 55.0, 54.8, 49.9, 49.8, 21.75, 21.74, 21.6, 21.5. HRMS (ESI) found m/z 906.28393, calcd m/z 906.28397 for C42H43IrN5O6 [M-H]+. Elemental analysis calcd (%) for C42H42IrN5O6·CH3OH, C, 55.11; H, 4.95; N, 7.47, found C, 55.23; H, 4.71; N, 7.58.
IrL92
.
Crude [Ir(L9)2(μ-Cl)]2 (0.044 g, 0.05 mmol), 2-picolinic acid (0.018 g, 0.14 mmol), Na2CO3 (0.015 g, 0.14 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH) progressive increase from 1% to 10% of CH3OH. Product isolated as an orange powder (37 mg, 63%). 1H NMR (500 MHz, CD2Cl2) δ 8.36 (s, J = 11.6 Hz, 1H; H10), 8.17 (d, J = 7.6 Hz, 1H; Hα), 7.94 (d, J = 5.3 Hz, 1H; Hδ), 7.91 (td, J = 7.7, 1.5 Hz, 1H; Hβ), 7.82–7.73 (m, 4H; H5,5′, H13,13′), 7.50 (dd, J = 8.7, 1.3 Hz, 1H; H12), 7.44–7.36 (m, 2H; Hγ, H12′), 6.61 (dd, J = 8.8, 2.6 Hz, 1H; H4′), 6.56 (dd, J = 8.8, 2.6 Hz, 1H; H4), 6.00 (d, J = 2.6 Hz, 1H; H2′), 5.79 (s, 1H; H10′), 5.75–5.62 (m, 3H; 2(CH(CH3)2), H2), 3.45 (s, 3H), 3.37 (s, 3H), 1.88–1.81 (m, 12H). 13C NMR (126 MHz, CD2Cl2) δ 173.1, 166.4, 165.3, 161.2, 161.0, 155.6, 153.9, 153.5, 149.9, 140.9, 140.7, 138.3, 136.1, 135.8, 128.3, 128.2, 128.1, 127.7, 127.6, 127.4, 127.3, 126.6, 126.4, 126.1, 126.0, 125.91, 125.87, 125.7, 125.5, 125.2, 124.0, 123.7, 121.8, 121.6, 119.9, 119.58, 119.56, 119.53, 119.07, 119.05, 119.02, 115.02, 114.99, 114.3, 113.9, 112.12, 112.09, 108.3, 107.4, 55.0, 54.9, 50.6, 50.5, 22.01, 21.97, 21.8, 21.7. 19F NMR (470 MHz, CD2Cl2) δ −61.2, −61.5. HRMS (ESI) found m/z 982.23679, calcd m/z 982.23760 for C42H37F6IrN5O4 [M-H]+. Elemental analysis calcd (%) for C42H36F6IrN5O4, C, 51.42; H, 3.71; N, 7.14, found C, 51.56; H, 3.84; N, 7.25.
IrL102
.
Crude [Ir(L10)2(μ-Cl)]2 (0.043 g, 0.05 mmol), 2-picolinic acid (0.018 g, 0.15 mmol), Na2CO3 (0.016 g, 0.15 mmol) and 2-ethoxyethanol/water 3
:
1 (10 mL). SiO2 (CH2Cl2/CH3OH) progressive increase from 1% to 5% of CH3OH. Product isolated as a bright yellow powder (27 mg, 57%). 1H NMR (500 MHz, CD2Cl2) δ 8.17 (d, J = 7.3 Hz, 1H; Hα), 7.94–7.86 (m, 2H; H10, Hβ), 7.83–7.73 (m, 3H; Hδ, H5,5′), 7.36 (ddd, J = 7.5, 5.4, 1.5 Hz, 1H; Hγ), 7.21 (dd, J = 8.3, 1.3 Hz, 1H; H4′), 7.18–7.11 (m, 3H; H18, 13, H4), 6.95 (dd, J = 9.1, 2.3 Hz, 1H; H11), 6.60 (d, J = 1.4 Hz, 1H; H2′), 6.56 (dd, J = 9.0, 2.3 Hz, 1H; H11′), 6.31 (d, J = 1.3 Hz, 1H; H2), 5.70–5.61 (m, 2H; 2(CH(CH3)2)), 5.60 (d, J = 9.0 Hz, 1H; H10), 3.88 (s, J = 7.6 Hz, 3H; –OCH3), 3.85 (s, 3H; –OCH3), 1.88 (dd, J = 14.7, 7.0 Hz, 6H 2(CH(CH3)2)), 1.81 (t, J = 7.1 Hz, 6H 2(CH(CH3)2)). 13C NMR (126 MHz, CD2Cl2) δ 173.3, 162.2, 161.1, 157.4, 156.9, 153.7, 151.6, 149.9, 149.7, 140.0, 139.8, 139.7, 138.3, 135.5, 135.4, 135.0, 134.7, 130.62, 130.59, 130.56, 130.53, 130.4, 130.1, 130.04, 130.01, 129.98, 129.95, 129.89, 129.7, 128.3, 128.0, 125.6, 125.4, 127.0, 124.6, 123.4, 123.2, 118.78, 118.75, 118.72, 118.68, 118.22, 118.19, 118.16, 115.9, 113.7, 113.3, 98.4, 98.0, 56.5, 56.4, 50.5, 21.8, 21.70, 21.66, 21.60. 19F NMR (470 MHz, CD2Cl2) δ −63.2, −63.3. HRMS (ESI) found m/z 982.23673, calcd m/z 982.23760 for C42H37F6IrN5O4 [M-H]+. Elemental analysis calcd (%) for C42H36F6IrN5O4, C, 51.42; H, 3.71; N, 7.14, found C, 51.49; H, 3.37; N, 6.99.
Computational details
Density functional theory (DFT) simulations have been performed using the Gaussian16 package.81 Based on previous theoretical investigations conducted by some of us,82–85 we considered the B3PW91 functional86–88 in addition to the LanL2Dz basis set, which includes a pseudopotential to describe core electrons for large atoms together with polarization functions on C (d; 0.587), N (d; 0.736), O (d; 0.961), F (d; 1.577) Cl (d; 0.75) and Ir (f; 0.938).89–93 The polarizable continuum model (PCM)94,95 was used to take into account solvent effects (CH2Cl2). For computational savings, the –OMe and –NiPr fragments were replaced by –OH and –NH groups. Geometry relaxations of the singlet (ground state) and triplet (excited state) states were performed and carefully checked by the calculation of vibrational frequencies. The general adiabatic shift approach (AS)96 was considered for estimating vibrational contributions to the computation of emission spectra.
All the simulations of phosphorescence spectra were performed within the Franck–Condon approximation. The vibronic calculations were achieved enforcing a sum-overstates (time-independent) approach which implies a truncation of the summation over an infinite number of states. To limit the number of integrals to be taken into account, a class-based prescreening has been employed based on the work of Santoro and coworkers and as implemented in Gaussian.97–99
In the present work, the following parameters were enforced:
Cmax1 = 70, Cmax2 = 70, NmaxI = 100 × 108 |
The highest class state (maxbands tag) considered was 9.
Post-treatments were done using the Gaussview and VMS packages.100–102 Horseshoe plots were realized using an in house Python code.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
The authors thank the CNRS and Université de Grenoble Alpes for their support. This work benefited from state aid managed by the National Research Agency under the “Investments for the future” and frp, the “Investissements d'avenir” program bearing the reference ANR-15-IDEX-02. E. M.-V. was supported by the CONACYT program 707986. This work was partially supported by the CBH-EUR-GS (ANR-17-EURE-0003). For the purpose of Open Access, a CC-BY 3.0 public copyright licence (https://creativecommons.org/licenses/by/3.0/) has been applied by the authors to the present document and will be applied to all subsequent versions up to the Author Accepted Manuscript arising from this submission. The NanoBio ICMG (UAR 2607) is acknowledged for providing facilities for mass spectrometry (A. Durand, L. Fort, and R. Gueret), and single-crystal X-ray diffraction (N. Altounian).
References
-
D. J. Gaspar and E. Polikarpov, OLED Fundamentals: Materials, Devices, and Processing of Organic Light-Emitting Diodes, CRC Press, 1st edn, 2015 Search PubMed.
- A. Sandström, P. Matyba and L. Edman, Appl. Phys. Lett., 2010, 96, 053303 CrossRef.
-
J. Kalinowski, V. Fattori, M. Cocchi and J. A. G. Williams, Light-emitting devices based on organometallic platinum complexes as emitters, Springer International Publishing, Berlin, Heidelberg Germany, 2011, vol. 255 Search PubMed.
-
H. Yersin, Topics in Current Chemistry, 2004, pp. 1–26 Search PubMed.
-
H. Yersin, Highly Efficient OLEDs with Phosphorescent Materials, Wiley, 2007 Search PubMed.
- M. A. Baldo, M. E. Thompson and S. R. Forrest, Nature, 2000, 403, 750–753 CrossRef CAS PubMed.
-
A. F. Rausch, H. H. Homeier and H. Yersin, Organometallic Pt(II) and Ir(III) Triplet Emitters for OLED Applications and the Role of Spin–Orbit Coupling: A Study Based on High-Resolution Optical Spectroscopy, Springer, Berlin, Heidelberg Germany, 2010 Search PubMed.
- C. F. R. Mackenzie, L. Zhang, D. B. Cordes, A. M. Z. Slawin, I. D. W. Samuel and E. Zysman-Colman, Adv. Opt. Mater., 2023, 11, 1–14 Search PubMed.
- Y. Zhang and J. Qiao, iScience, 2021, 24, 102858 CrossRef CAS PubMed.
- M. Mauro, Chem. Commun., 2021, 57, 5857–5870 RSC.
-
L. F. Gildea and J. A. G. Williams, in Organic Light-Emitting Diodes (OLEDs), ed. A. Buckley, Woodhead Publishing, Sawston, UK, 2013, pp. 77–113 Search PubMed.
- K. P. S. Zanoni, R. L. Coppo, R. C. Amaral and N. Y. Murakami Iha, Dalton Trans., 2015, 44, 14559–14573 RSC.
- C. E. Housecroft and E. C. Constable, Coord. Chem. Rev., 2017, 350, 155–177 CrossRef CAS.
- E. Baranoff, J.-H. Yum, I. Jung, R. Vulcano, M. Grätzel and M. K. Nazeeruddin, Chem. – Asian J., 2010, 5, 496–499 CrossRef CAS PubMed.
- H. Guo, S. Ji, W. W. Wu, J. Shao and J. Zhao, Analyst, 2010, 135, 2832–2840 RSC.
- S. Medina-Rodríguez, S. A. Denisov, Y. Cudré, L. Male, M. Marín-Suárez, A. Fernández-Gutiérrez, J. F. Fernández-Sánchez, A. Tron, G. Jonusauskas, N. D. McClenaghan and E. Baranoff, Analyst, 2016, 141, 3090–3097 RSC.
- J.-L. Fillaut, J. Andriès, R. D. Marwaha, P.-H. Lanoë, O. Lohio, L. Toupet and J. A. Gareth Williams, J. Organomet. Chem., 2008, 693, 228–234 CrossRef CAS.
- P.-H. Lanoë, J.-L. Fillaut, V. Guerchais, H. Le Bozec and J. A. Gareth Williams, Eur. J. Inorg. Chem., 2011, 2011, 1255–1259 CrossRef.
- E. Ortega-Forte, S. Hernández-García, G. Vigueras, P. Henarejos-Escudero, N. Cutillas, J. Ruiz and F. Gandía-Herrero, Cell. Mol. Life Sci., 2022, 79, 510 CrossRef CAS PubMed.
- J. Shen, T. W. Rees, L. Ji and H. Chao, Coord. Chem. Rev., 2021, 443, 214016 CrossRef CAS.
- C. Caporale and M. Massi, Coord. Chem. Rev., 2018, 363, 71–91 CrossRef CAS.
- C. Y.-S. Chung and V. W.-W. Yam, Chemistry, 2014, 20, 13016–13027 CrossRef CAS PubMed.
- T. Tu, W. Fang, X. Bao, X. Li and K. H. Dötz, Angew. Chem., Int. Ed., 2011, 50, 6601–6605 CrossRef CAS PubMed.
- K. Li, T. Zou, Y. Chen, X. Guan and C. Che, Chem. – Eur. J., 2015, 7441–7453 CrossRef CAS PubMed.
- C. K. Prier, D. A. Rankic and D. W. C. MacMillan, Chem. Rev., 2013, 113, 5322–5363 CrossRef CAS PubMed.
- H. L. B. S. Di Bella, C. Dragonetti, M. Pizzotti, D. Roberto, F. Tessore, R. Ugo, M. G. Humphrey, M. P. Cifuentes, M. Samoc, L. Murphy, J. A. G. Williams, Z. Liu, Z. Bian, C. Huang, N. C. Fletcher, M. C. Lagunas and V. Guerchais, Top. Organomet. Chem., 2010, 37, 179 Search PubMed.
-
A. Barbieri, F. Barigelletti, E. C.-C. Cheng, L. Flamigni, T. Gunnlaugsson, R. A. G. Kirgan, D. Kumaresan, J. P. Leonard, C. B. Nolan, D. P. Rillema, C. Sabatini, R. H. Schmehl, K. Shankar, F. Stomeo, B. P. Sullivan, S. Vaidya, B. Ventura and J. A. G. Williams, Photochemistry and Photophysics of Coordination Compounds II, in Topics in current Chemistry, Springer, 2007 Search PubMed.
- D. H. Volman, G. S. Hammond, D. C. Neckers, M. Maestri, V. Balzani, C. Deuschel-Cornioley and A. Von Zelewsky, Adv. Photochem., 1992, 1 DOI:10.1002/9780470133484.ch1.
-
Z. Bian, M. P. Cifuentes, S. Di Bella, N. C. Fletcher, C. Dragonetti, V. Guerchais, C. Huang, M. G. Humphrey, M. C. Lagunas, J. A. G. Le Bozec, H. Williams, Z. Liu, L. Murphy, M. Pizzotti, R. M. S. Dominique, A. F. Tessore and R. Ugo, Molecular Organometallic Materials for Optics, Springer, 2011, vol. 33 Search PubMed.
- G. Lu, L. Gu, S. Li, H.-B. Han, X. Wang, C. Zhou, J.-J. Lu and L. Zhou, New J. Chem., 2023, 18603–18609 Search PubMed.
- T. Sajoto, P. I. Djurovich, A. B. Tamayo, J. Oxgaard, W. A. Goddard and M. E. Thompson, J. Am. Chem. Soc., 2009, 131, 9813–9822 CrossRef CAS PubMed.
- P. J. Hay, J. Phys. Chem. A, 2002, 106, 1634–1641 CrossRef CAS.
- A. J. Lees, Comments Inorg. Chem., 1995, 17, 319–346 CrossRef CAS.
- R. D. Costa, E. Ortí, H. J. Bolink, S. Graber, S. Schaffner, M. Neuburger, C. E. Housecroft and E. C. Constable, Adv. Funct. Mater., 2009, 19, 3456–3463 CrossRef CAS.
- A. F. Henwood and E. Zysman-Colman, Top. Curr. Chem., 2016, 374, 36 CrossRef PubMed.
-
L. Flamigni, A. Barbieri, C. Sabatini, B. Ventura and F. Barigelletti, Photochemistry and Photophysics of Coordination Compounds II, 2007, pp. 143–203 Search PubMed.
- P. A. Scattergood, A. M. Ranieri, L. Charalambou, A. Comia, D. A. W. Ross, C. R. Rice, S. J. O. Hardman, J.-L. Heully, I. M. Dixon, M. Massi, F. Alary and P. I. P. Elliott, Inorg. Chem., 2020, 59, 1785–1180 CrossRef CAS PubMed.
- Z.-J. Gao, T.-H. Yeh, J.-J. Xu, C.-C. Lee, A. Chowdhury, B.-C. Wang, S.-W. Liu and C.-H. Chen, ACS Omega, 2020, 5, 10553–10561 CrossRef CAS PubMed.
- D. Chen, W.-Z. He, H.-S. Liao, Y.-X. Hu, D.-D. Xie, B.-Y. Wang, H.-J. Chi, Y.-L. Lv, X. Zhu and X. Li, Org. Electron., 2023, 113, 106715 CrossRef CAS.
- Z. Ge, T. Hayakawa, S. Ando, M. Ueda, T. Akiike, H. Miyamoto, T. Kajita and M. Kakimoto, Chem. Mater., 2008, 20, 2532–2537 CrossRef CAS.
- J. H. Zhao, Y. X. Hu, H. Y. Lu, Y. L. Lü and X. Li, Org. Electron., 2017, 41, 56–72 CrossRef CAS.
- J. F. Lemonnier, L. Guénée, C. Beuchat, T. A. Wesolowski, P. Mukherjee, D. H. Waldeck, K. A. Gogick, S. Petoud and C. Piguet, J. Am. Chem. Soc., 2011, 133, 16219–16234 CrossRef CAS PubMed.
- N. M. Shavaleev, S. V. Eliseeva, R. Scopelliti and J. C. G. Bünzli, Chem. – Eur. J., 2009, 15, 10790–10802 CrossRef CAS PubMed.
- J.-H. Zhao, Y.-X. Hu, Y. Dong, X. Xia, H.-J. Chi, G.-Y. Xiao, X. Li and D.-Y. Zhang, New J. Chem., 2017, 41, 1973–1979 RSC.
- W. S. Huang, J. T. Lin, C. H. Chien, Y. T. Tao, S. S. Sun and Y. S. Wen, Chem. Mater., 2004, 16, 2480–2488 CrossRef CAS.
- J. Jayabharathi, K. Jayamoorthy and V. Thanikachalam, J. Organomet. Chem., 2014, 761, 74–83 CrossRef CAS.
- Y. Jiao, M. Li, N. Wang, T. Lu, L. Zhou, Y. Huang, Z. Lu, D. Luo and X. Pu, J. Mater. Chem. C, 2016, 4, 4269–4277 RSC.
- H. Cao, H. Sun, Y. Yin, X. Wen, G. Shan, Z. Su, R. Zhong, W. Xie, P. Li and D. Zhu, J. Mater. Chem. C, 2014, 2, 2150–2159 RSC.
- X. Wei, J. Peng, J. Cheng, M. Xie, Z. Lu, C. Li and Y. Cao, Adv. Funct. Mater., 2007, 17, 3319–3325 CrossRef CAS.
- G. Li, Y. Wu, G. Shan, W. Che, D. Zhu, B. Song, L. Yan, Z. Su and M. R. Bryce, Chem. Commun., 2014, 50, 6977–6980 RSC.
- X. Ouyang, D. Chen, S. Zeng, X. Zhang, S. Su and Z. Ge, J. Mater. Chem., 2012, 22, 23005–23011 RSC.
- H. T. Cao, L. Ding, J. Yu, G. G. Shan, T. Wang, H. Z. Sun, Y. Gao, W. F. Xie and Z. M. Su, Dyes Pigm., 2019, 160, 119–127 CrossRef CAS.
- J. X. Cai, T. L. Ye, X. F. Fan, C. M. Han, H. Xu, L. L. Wang, D. G. Ma, Y. Lin and P. F. Yan, J. Mater. Chem., 2011, 21, 15405–15416 RSC.
- L. Han, D. Zhang, Y. Zhou, Y. Yang, H. Y. Woo, F. Bai and R. Yang, Dyes Pigm., 2013, 99, 1010–1015 CrossRef CAS.
- T. Yu, F. Yang, X. Chen, W. Su and Y. Zhao, New J. Chem., 2017, 41, 2046–2054 RSC.
- H.-T. T. Mao, C.-X. X. Zang, G.-G. G. Shan, H.-Z. Z. Sun, W.-F. F. Xie and Z.-M. M. Su, Inorg. Chem., 2017, 56, 9979–9987 CrossRef CAS PubMed.
- S. V. Tatarin, P. Kalle, I. V. Taydakov, E. A. Varaksina, V. M. Korshunov and S. I. Bezzubov, Dalton Trans., 2021, 50, 6889–6900 RSC.
- E. Martínez-Vollbert, C. Philouze, I. Gautier-Luneau, Y. Moreau, P.-H. Lanoë and F. Loiseau, Phys. Chem. Chem. Phys., 2021, 23, 24789–24800 RSC.
- E. Martìnez-Vollbert, C. Ciambrone, W. Lafargue-Dit-Hauret, C. Latouche, F. Loiseau and P.-H. Lanoë, Inorg. Chem., 2022, 61, 3033–3049 CrossRef PubMed.
- T.-R. Chen, P.-C. Liu, H.-P. Lee, F.-S. Wu and K. H.-C. Chen, Eur. J. Inorg. Chem., 2017, 2017, 2023–2031 CrossRef CAS.
- D. G. Congrave, Y. Ting Hsu, A. S. Batsanov, A. Beeby and M. R. Bryce, Organometallics, 2017, 36, 981–993 CrossRef CAS.
- F. O. Garces, K. Dedeian, N. L. Keder and R. J. Watts, Acta Crystallogr., Sect. C: Cryst. Struct. Commun., 1993, 49, 1117 CrossRef.
- M. Nonoyama, Bull. Chem. Soc. Jpn., 1974, 47, 767 CrossRef CAS.
- K. J. Suhr, L. D. Bastatas, Y. Shen, L. A. Mitchell, G. A. Frazier, D. W. Taylor, J. D. Slinker and B. J. Holliday, Dalton Trans., 2016, 45, 17807–17823 RSC.
- K. J. Suhr, L. D. Bastatas, Y. Shen, L. A. Mitchell, B. J. Holliday and J. D. Slinker, ACS Appl. Mater. Interfaces, 2016, 8, 8888–8892 CrossRef CAS PubMed.
- M. Martínez-Alonso, J. Cerdá, C. Momblona, A. Pertegás, J. M. Junquera-Hernández, A. Heras, A. M. Rodríguez, G. Espino, H. Bolink and E. Ortí, Inorg. Chem., 2017, 56, 10298–10310 CrossRef PubMed.
- K. A. Phillips, T. M. Stonelake, K. Chen, Y. Hou, J. Zhao, S. J. Coles, P. N. Horton, S. J. Keane, E. C. Stokes, I. A. Fallis, A. J. Hallett, S. P. O'Kell, J. M. Beames and S. J. A. Pope, Chem. – Eur. J., 2018, 24, 8577–8588 CrossRef CAS PubMed.
-
G. A. Jeffrey, An Introduction to Hydrogen Bonding, Oxford University Press, Oxford, 1997 Search PubMed.
- T. Steiner, Angew. Chem., Int. Ed., 2002, 41, 48–76 CrossRef CAS.
- E. Baranoff and B. F. E. Curchod, Dalton Trans., 2015, 44, 8318–8329 RSC.
- S. Neukermans, F. Vorobjov, T. Kenis, R. De Wolf, J. Hereijgers and T. Breugelmans, Electrochim. Acta, 2020, 332, 135484 CrossRef CAS.
- J. Frey, B. F. E. Curchod, R. Scopelliti, I. Tavernelli, U. Rothlisberger, M. K. Nazeeruddin and E. Baranoff, Dalton Trans., 2014, 43, 5667–5679 RSC.
-
J. C. Deaton and F. N. Castellano, Iridium(III) in Optoelectronic and Photonics Applications, 2017, pp. 1–69 Search PubMed.
-
L. Silvestroni, G. Accorsi, N. Armaroli, V. Balzani, G. Bergamini, S. Campagna, F. Cardinali, C. Chiorboli, M. T. Indelli, N. A. P. Kane-Maguire, A. Listorti, F. Nastasi, F. Puntoriero and F. Scandola, Photochemistry and Photophysics of Coordination Compounds I, 2007, vol. 281 Search PubMed.
- Y. You and S. Y. Park, Dalton Trans., 2009, 1267–1282 RSC.
- G. A. Crosby, Acc. Chem. Res., 1975, 8, 231–238 CrossRef CAS.
- E. Baranoff, B. F. E. Curchod, F. Monti, F. Steimer, G. Accorsi, I. Tavernelli, U. Rothlisberger, R. Scopelliti, M. Grätzel and M. K. Nazeeruddin, Inorg. Chem., 2012, 51, 799–811 CrossRef CAS PubMed.
-
A. Barbieri, F. Barigelletti, E. C.-C. Cheng, L. Flamigni, T. Gunnlaugsson, R. A. Kirgan, D. Kumaresan, J. P. Leonard, C. B. Nolan, D. P. Rillema, C. Sabatini, R. H. Schmehl, K. Shankar, F. Stomeo, B. P. Sullivan, S. Vaidya, B. Ventura, J. A. G. Williams and V. W.-W. Yam, Topics in Current Chemistry: Photochemistry and Photophysics of Coordination Compounds II, Springer Berlin Heidelberg, 1980, vol. 7 Search PubMed.
- S. Lamansky, P. Djurovich, D. Murphy, F. Abdel-Razzaq, R. Kwong, I. Tsyba, M. Bortz, B. Mui, R. Bau and M. E. Thompson, Inorg. Chem., 2001, 40, 1704–1711 CrossRef CAS PubMed.
- A. Juris, V. Balzani, F. Barigelletti, S. Campagna, P. Belser and A. von Zelewsky, Coord. Chem. Rev., 1988, 84, 85–277 CrossRef CAS.
-
M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Jr. Montgomery, J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16 (Revision B.01), Gaussian Inc., Wallingford CT, 2016 Search PubMed.
- F. Vazart and C. Latouche, Theor. Chem. Acc., 2015, 134, 144 Search PubMed.
- C. Latouche, F. Palazzetti, D. Skouteris and V. Barone, J. Chem. Theory Comput., 2014, 10, 4565–4573 CrossRef CAS PubMed.
- C. Latouche, A. Baiardi and V. Barone, J. Phys. Chem. B, 2015, 119, 7253–7257 CrossRef CAS PubMed.
- C. Latouche, D. Skouteris, F. Palazzetti and V. Barone, J. Chem. Theory Comput., 2015, 11, 3281–3289 CrossRef CAS PubMed.
- J. P. Perdew, K. Burke and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 16533–16539 CrossRef CAS PubMed.
- J. P. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 33, 8822–8824 CrossRef PubMed.
- A. D. Becke, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS.
-
T. H. Dunning and P. J. Hay, Generalized gradient approximation for the exchange-correlation hole of a many-electron system, ed. H. F. Schaefer, Springer US, Boston, MA, 1977, pp. 1–27 Search PubMed.
- W. R. Wadt and P. J. Hay, J. Chem. Phys., 1985, 82, 284–298 CrossRef CAS.
- P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 299–310 CrossRef CAS.
- P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 270–283 CrossRef CAS.
- R. Schira and C. Latouche, Dalton Trans., 2021, 50, 746–753 RSC.
-
R. C. Benedetta Mennucci, Continuum Solvation Models in Chemical Physics: From Theory to Applications, Wiley, 2008 Search PubMed.
- B. Mennucci, J. Tomasi, R. Cammi, J. R. Cheeseman, M. J. Frisch, F. J. Devlin, S. Gabriel and P. J. Stephens, J. Phys. Chem. A, 2002, 106, 6102–6113 CrossRef CAS.
- J. Bloino, M. Biczysko, F. Santoro and V. Barone, J. Chem. Theory Comput., 2010, 6, 1256–1274 CrossRef CAS.
- F. Santoro, R. Improta, A. Lami, J. Bloino and V. Barone, J. Chem. Phys., 2007, 126, 084509 CrossRef PubMed.
- F. Santoro, A. Lami, R. Improta and V. Barone, J. Chem. Phys., 2007, 126, 184102 CrossRef PubMed.
- F. Santoro, A. Lami, R. Improta, J. Bloino and V. Barone, J. Chem. Phys., 2008, 128, 224311 CrossRef PubMed.
- D. Licari, A. Baiardi, M. Biczysko, F. Egidi, C. Latouche and V. Barone, J. Comput. Chem., 2015, 36, 321–334 CrossRef CAS PubMed.
- V. Barone, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2016, 6, 86–110 CAS.
-
R. Dennington, T. A. Keith and J. M. Millam, GaussView version 6, 2019 Search PubMed.
Footnotes |
† Electronic supplementary information (ESI) available. CCDC 2246643–2246645. For ESI and crystallographic data in CIF or other electronic format see DOI: https://doi.org/10.1039/d3dt03498d |
‡ Please contact this author regarding the theoretical aspects of the article. |
|
This journal is © The Royal Society of Chemistry 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.