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

Cyclometalated cinchophen ligands on iridium(III): towards water-soluble complexes with visible luminescence

Rebecca A. Smith , Emily C. Stokes , Emily E. Langdon-Jones , James A. Platts , Benson M. Kariuki , Andrew J. Hallett * and Simon J. A. Pope *
School of Chemistry, Main Building, Cardiff University, Cardiff CF10 3AT, UK. E-mail: popesj@cardiff.ac.uk; hallettaj@cardiff.ac.uk; Fax: +44 (0) 29-20874030; Tel: +44 (0) 29-20879316

Received 26th April 2013 , Accepted 29th May 2013

First published on 6th June 2013


Abstract

Eight cationic heteroleptic iridium(III) complexes, [Ir(epqc)2(N^N)]+, were prepared in high yield from a cyclometalated iridium bridged-chloride dimer bearing two ethyl-2-phenylquinoline-4-carboxylate (epqc) ligands. Two X-ray crystallographic studies were undertaken on selected complexes (where the ancillary ligand N^N = 4,4′-dimethyl-2,2′-bipyridine and 4,7-diphenyl-1,10-phenanthroline) each confirming the proposed formulations, showing an octahedral coordination at Ir(III). In general, the complexes are luminescent (620–630 nm) with moderately long lifetimes indicative of phosphorescence. Hydrolysis of the ethyl ester moieties of the epqc ligands gave the analogous cinchophen-based complexes, which were water-soluble and visibly luminescent (568–631 nm). The spectroscopic and redox characterisation of the complexes was complemented by DFT and TD-DFT calculations, supporting the assignment of dominant 3MLCT to the emissive character.


Introduction

There has been considerable attention and effort disbursed on iridium(III) complexes with cyclometalated ligands, due to their tuneable and generally efficient photoluminescence properties and subsequent performance in a variety of photophysical and electronic applications.1 Such complexes are capable of showing intense phosphorescence at room temperature;2 the heavy atom iridium centre mediating strong spin–orbit coupling and intersystem crossing (ISC), mixing the singlet and triplet excited states and generating high phosphorescent efficiencies. As a consequence, cyclometalated iridium complexes have found many applications in a variety of opto-electronically related applications, such as electrochemical cells,3 photovoltaics,4 and luminescence imaging.5 Their ability to perform in such roles relies upon an understanding of their excited state properties, which can be modulated by altering the cyclometalating and/or ancillary ligand associated with the iridium centre. The ability to tune luminescence emission wavelengths through variation of cyclometalating ligands and ancillary ligands (predominantly for cationic complexes) renders such complexes very useful, particularly in electrochemiluminescence (ECL).6 For example, increasing the π-conjugation of phenylpyridine by adding an aromatic ring (to give phenylquinoline) bathochromically shifts the 3π–π* and triplet metal-to-ligand charge transfer (3MLCT) emission due to lower lying π* orbitals.7 Consequently, there have been several reports of phenylquinoline derivatives as cyclometalating ligands with various ancillary ligands.8 However, many applications of Ir(III) complexes in this context, including biologically-related uses, require water solubility. Surprisingly then, there have only been a handful of reports of water-soluble iridium cyclometalated complexes, most commonly where the ancillary diimine ligands are functionalised with solubilising groups such as sugars,8j,l triazoles,9 polyethyleneglycol (PEG),10 bioconjugates11 and carboxylate groups,12 as well as the bis-cyclometalated bis-aqua complexes;13 reports of water-solubilising fuctionalisation at the cyclometalating ligand are extremely rare.14 The purpose of this paper is to present the synthesis and photophysical properties of a class of iridium complex that incorporate cyclometalated cinchophen-based ligands, providing a convenient route towards water-soluble complexes with exploitable photophysical properties; the structural, spectroscopic, electrochemical and photophysical studies are presented together with supporting DFT and TD-DFT calculations on the complexes.

Results and discussion

Synthesis and characterisation of the ligands and complexes

As an aside, it is noteworthy that many of the literature reports on functionalised phenylquinoline compounds describe biological activities15 with applications as anti-malarial, anti-inflammatory and antibacterial agents;16 ester-functionalised phenylquinoline compounds have shown promise as amyloid fibril17 and reverse transcriptase18 inhibitors. Ethyl-2-phenylquinoline-4-carboxylate (epqcH) was prepared by simply dissolving and heating 2-phenylquinoline-4-carboxylic acid (cinchophen) in ethanol with a few drops of conc. H2SO4. The precursor iridium chloro-bridged dimer [(epqc)2Ir(μ-Cl)2Ir(epqc)2] (2) was synthesised according to established literature conditions.19 Compound 2 readily reacted (Scheme 1) with a range of selected diimine-type ligands in 2-methoxyethanol; work-up and counter anion exchange yielded the crude monometallic complexes, [Ir(epqc)2(N^N)]PF6 {N^N = 2,2′-bipyridine (bpy) 3a; 4,4′-dimethyl-2,2′-bipyridine (dmbpy) 3b; 1,10-phenanthroline (phen) 3d; 4,7-diphenyl-1,10-phenanthroline (dip) 3e; dipyrido[3,2-a:2′,3′-c]phenazine (dppz) 3f; benzo[i]dipyrido[3,2-a:2′,3′-c]phenazine (dppn) 3g; naphtha[2,3-a]dipyrido[3,2-h:2′,3′-f]phenazine-5,18-dione (qdppz) 3h}. For the reaction of 2 and diethyl-2,2′-bipyridine-4,4′-dicarboxylate (debpy), the ethyl ester groups of bpy were selectively converted to 2-methoxyethyl esters in situ, as informed by 1H NMR spectroscopy and mass spectrometry therefore giving [Ir(epqc)2(dmbpc)]PF63c (dmbpc = di-2-methoxyethyl-2,2′-bipyridine-4,4′-dicarboxylate). It should be noted that the ethyl ester functionality of the cyclometalated cinchophen ligand was retained in all cases. Further purification for each complex was achieved using column chromatography (silica; MeOH–CH2Cl2, 1[thin space (1/6-em)]:[thin space (1/6-em)]9) with elution of the first red band and subsequent removal of solvent giving the complexes as pure red-coloured powders in good yields (>75%). The resultant complexes 3a–h were soluble in a range of common organic solvents.
Synthetic route to complexes [Ir(epqc)2(N^N)]PF63a–h and [Ir(pqca)2(N^N)]Cl 4a–g.
Scheme 1 Synthetic route to complexes [Ir(epqc)2(N^N)]PF63a–h and [Ir(pqca)2(N^N)]Cl 4a–g.

The conversion of complexes 3a–g to their corresponding free acids was achieved by stirring the esterified complexes in an equi-volume mixture of 1 M KOH and acetone under an inert atmosphere. Subsequent neutralisation with 1 M HCl, removal of solvent and extraction with methanol (allowing removal of KCl) led to the isolation of complexes 4a–g as their chloride salts, [Ir(pqca)2(N^N)]Cl. However, 1H NMR spectroscopy and mass spectrometry indicated that it was not possible to isolate 4h by this method, the reasons for which are currently unknown.

All new complexes were characterised using a range of spectroscopic techniques. Firstly, the 1H NMR spectra of complexes 3a–h are complicated in the aromatic region with overlapping resonances associated with the cyclometalated and diimine ligands; the retention of the ethyl ester functionality was observed in the aliphatic region. The 2-methoxyethyl groups of 3c appeared as broadened triplets at 4.46 and 3.68 ppm together with a singlet at 3.30 ppm. 31P–{1H} NMR spectroscopy confirmed the presence of the PF6 ion with a signature septet (1JPF) at ca. −145 ppm in all cases. Low and high resolution mass spectra were obtained for the complexes 3a–h and each confirmed the identity of the cationic, monomeric species of type [(epqc)2Ir(N^N)]+, revealing the parent cations of [M − PF6]+ in each case, with the appropriate isotopic distribution. Complex purity was confirmed by elemental analysis.

Upon hydrolysis of the ester groups, the 1H NMR spectra of the isolated analogues 4a–g all confirmed the absence of the ethyl, or 2-methoxyethyl groups in the case of 4c, yielding 2,2′-bipyridyl-4,4′-dicarboxylic acid (bpdc), as well as improved resolution of the aromatic resonances. The absence of a resonance in the 31P–{1H} NMR spectra indicated the exchange of PF6 with Cl counter ions in the deprotected species. The corresponding mass spectra were easily obtained: complexes 4a–g confirmed the identity of the cationic, monomeric species of type [Ir(pqca)2(N^N)]+ showing the characteristic cluster of peaks associated with the [M − Cl]+ parent ion. Solid-state IR spectra were also obtained on all complexes highlighting the conversion from ester (ca. 1720 cm−1) to carboxylate (ca. 1580 cm−1) via significant low energy shifts in ν(C[double bond, length as m-dash]O), and the absence of the PF6 counter ion stretch (ca. 835 cm−1) confirming exchange with chloride.

X-ray crystallography studies

Single crystals of 3b and 3e suitable for X-ray diffraction studies were isolated following vapour diffusion of Et2O into concentrated CHCl3 or MeCN solutions of the complexes, over a period of 48 h at −20 °C. The bond lengths and bond angles are reported in Table 1, and the associated data collection parameters are reported in Table S1, ESI.
Table 1 Selected bond lengths (Å) and angles (°) for complexes 3b and 3e
Bond length (Å)/angle (°) 3b Calculated value 3e Calculated value
a Non-bonded metrics and those involving planes/centroids were not included in the refinement, and thus do not have an e.s.d.
Ir(1)–N(1) 2.094(8) 2.107 2.078(6) 2.107
Ir(1)–N(2) 2.112(8) 2.107 2.107(7) 2.107
Ir(1)–N(3) 2.182(8) 2.220 2.190(6) 2.218
Ir(1)–N(4) 2.158(7) 2.220 2.153(7) 2.218
Ir(1)–C(1) 1.998(9) 1.997 1.989(8) 1.997
Ir(1)–C(19) 2.001(9) 1.997 2.005(8) 1.997
C(14)⋯N(3) 3.140a 3.127 3.009a 3.128
C(32)⋯N(4) 3.173a 3.128 3.232a 3.128
N(1)–Ir(1)–N(2) 170.7(3) 172.1 173.3(3) 172.8
N(1)–Ir(1)–C(1) 80.0(4) 80.0 79.4(5) 79.9
N(2)–Ir(1)–C(19) 80.3(3) 80.0 79.9(3) 79.9
C(1)–Ir(1)–C(19) 89.2(4) 90.1 88.6(3) 90.5


The structures obtained for the three complexes (Fig. 1a and 2a) confirmed the proposed formulations, and show that the iridium(III) center in these [Ir(epqc)2(N^N)]+ complexes adopts a distorted octahedral coordination geometry. trans Angles at the metal centers ranged from 168.0(3)° to 172.4(3)° for 3b and 168.0(3)° to 174.9(3)° for 3e. The diimine ligand is always coordinated trans to the cyclometalated phenyl rings; the complexes retain the cis-C,C and trans-N,N chelating disposition of the original chloro-bridged dimer as reported in related examples.7c,20 The trans-influence of the carbon donors rendered slightly longer Ir–N bond lengths for the diimine ligands {2.182(8) and 2.158(7) Å for 3b} than the epqc ligands {2.094(8) and 2.112(8) Å for 3b}. The bite angles of the epqc ligands (80.0(4)° and 80.3(4)° for 3b) were slightly larger than that of N^N ligands {75.1(3) for 3b}. Similar observations have been reported in related cyclometalated iridium(III) polypyridine systems [Ir(pq)2(N^N)]+.1e,7c,d,20 For 3e the structure showed that the phenyl substituents of the 4,7-diphenyl-1,10-phenanthroline ligand are twisted out of planarity from the phenanthroline unit by 55.1° and 50.2° respectively. In addition, as with related phenylquinoxaline complexes,7d,21 there is a distortion within the quinoline moiety caused by the steric interactions between the chelating ligands (Fig. 1b and 2b). Some of the inter-ligand C⋯C and C⋯N non-bonding contact distances (Table 1) are shorter than 3.4 Å (i.e. the sum of the van der Waals radii of the atoms). This results in the phenyl groups of the epqc ligand showing deformation angles of 14.1° and 19.1° for 3b and 17.5° and 28.6° for 3e with respect to the quinoline fragment of the ligand.


(a) Ortep representation of [Ir(epqc)2(dmbpy)]+3b (50% probability ellipsoids, solvent molecules, PF6− anion and hydrogen atoms have been omitted for clarity) and (b) showing non-bonding contact interactions between chelating ligands.
Fig. 1 (a) Ortep representation of [Ir(epqc)2(dmbpy)]+3b (50% probability ellipsoids, solvent molecules, PF6 anion and hydrogen atoms have been omitted for clarity) and (b) showing non-bonding contact interactions between chelating ligands.

(a) Ortep representation of [Ir(epqc)2(dip)]+3e (50% probability ellipsoids, solvent molecules, PF6− anion and hydrogen atoms have been omitted for clarity) and (b) showing non-bonding contact interactions between chelating ligands.
Fig. 2 (a) Ortep representation of [Ir(epqc)2(dip)]+3e (50% probability ellipsoids, solvent molecules, PF6 anion and hydrogen atoms have been omitted for clarity) and (b) showing non-bonding contact interactions between chelating ligands.

The bond lengths and angles of 3b and 3e were compared with the optimised values calculated from density functional theory (DFT) studies (also see DFT section and Table 1). In general, a reasonable agreement was obtained between the theoretical and experimentally observed bond lengths, although some small differences were found. The calculated Ir–Nbipyridine bond lengths, Ir–N(3) and Ir–N(4), are 0.038 and 0.062 Å longer than the experimental values for 3b. In the case of the cyclometalated Ir–C bonds, calculated values are very similar to the experimentally obtained data for Ir–C(1) in 3b and 3e; similarly the Ir–Nquinoline bonds, where the calculated and experimental values are comparable. Again, the calculated structures reveal C⋯C and C⋯N interactions between the ligands. The calculated deformation angles of the quinoline fragment are much lower for 3b and 3e.

Density functional theory (DFT) studies

DFT calculations (computed using the B3PW91 hybrid functional) were performed to investigate the frontier orbitals and provide qualitative descriptions of the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energy levels.

For the [Ir(epqc)2(N^N)]+ complexes, the energy levels of the HOMO are sufficiently different (ΔE > 0.2 eV) from the other MOs to be considered independent. In each case the HOMO was located on the metal 5d(Ir) centre and cyclometalated phenyl rings (Fig. S1, ESI), with little or no coverage of the ancillary diimine ligands. However, varying the diimine ligands does impart a subtle perturbation of the HOMO energy (EHOMO = −7.66 to −7.84 eV). The LUMO for the complexes is, in many cases, close enough in energy to be considered isoenergetic with other close-lying MOs (e.g. LUMO + 1, LUMO + 2). For the complexes where this is not the case (3c, 3f, 3g and 3h), the LUMOs are predominantly delocalised over the diimine ligands as with previous studies of related compounds.7d For 3a, 3b, 3d and 3e the orbitals show a mixture of diimine and phenylquinoline localisation. The diimine imparts a larger degree of variation in the energy levels of the LUMO energy (ELUMO = −4.71 to −5.47 eV), with 3h showing the lowest LUMO energy level (ELUMO = −5.47 eV) and, therefore, smallest bandgap (Ebandgap = 2.29 eV). These results suggest that the lowest energy absorption is predicted to comprise of significant MLCT character and that variation of the diimine ligand could lead to a small degree of tuneable optical properties within this series of complexes. The corresponding cinchophen complexes (4a–h) showed the same general localisation of the frontier orbitals, but revealed a drop in both the HOMO and LUMO energies by an average of 0.12 eV and 0.13 eV, respectively (Fig. S2, ESI).

Electrochemical studies

The electrochemical characteristics of the [Ir(epqc)2(N^N)]PF6 (3a–h) complexes were studied in de-oxygenated CH2Cl2. The HOMO energy levels (EHOMO) were determined from the ionisation potential of the first oxidation (Ir3+/4+) by direct correlation with the redox couple of FeCp20/1+. The cyclic voltammograms, measured at a platinum disc electrode (scan rate υ = 200 mV s−1, 1 × 10−3 M solutions, 0.1 M [NBu4][PF6] as a supporting electrolyte), each showed one non-fully reversible oxidation (Table S2, ESI), over the range +1.38 to +1.45 V. The extent of the irreversibility can be ascribed to the contribution of the cyclometalating ligands to the electron density of the HOMO,22 which in this case is calculated to be ca. 45% from DFT studies. The EHOMO values were determined using the relevant equations23 and the resultant values fall in the narrow range −5.72 to −5.83 eV (Table S3, ESI). Each complex also showed one or two partially reversible or irreversible reduction waves, assigned to ligand-centred processes involving both the diimine and phenylquinoline ligands, with complex 3h showing at least four reduction processes, some of which must be associated with the highly reducible anthraquinone fragment.

Electronic properties of the complexes

The UV-vis absorption spectra of complexes 3a–h were obtained as aerated MeCN solutions (5 × 10−5 M) (Fig. 3, Table 3). Strong absorption bands between 250 and 380 nm were assigned to spin allowed 1π–π* ligand-centered (LC) transitions arising from both the cyclometalated and diimine ligands within each complex. Weaker bands at 380–480 nm in the visible region were assigned to spin-allowed metal-to-ligand charge transfer bands (1MLCT) with the possibility of spin-forbidden 3MLCT transitions contributing to the weaker low-energy shoulder. For 3h it is also likely that ligand-centred transitions associated with the anthraquinone moiety also contribute in this wavelength region; previous studies have shown that such species can possess intra-ligand CT (formally n–π*) character, which is likely to contribute to the lower energy parts of the spectral profile. The variation in coordinated diimine ligand imparts only a very minor variation in the wavelength positioning of the visible absorption bands. As expected the absorption spectra of the corresponding carboxylic acid derivatives 4a–g (5 × 10−5 M MeOH) share many of the same common features. The principal observation from these spectra is that the visible MLCT-based bands appear as a defined transition ca. 430–455 nm.
UV-vis spectra of selected [Ir(epqc)2(N^N)]PF6 complexes in MeCN solutions (5 × 10−5 M).
Fig. 3 UV-vis spectra of selected [Ir(epqc)2(N^N)]PF6 complexes in MeCN solutions (5 × 10−5 M).

TD-DFT calculations (Table 2) in simulated MeCN suggest the assignment of the lowest lying absorption bands as having substantial MLCT character. For 3a the lowest energy absorption involves excitation from HOMO (Ir-5d + phenyl-π) to LUMO + 1 (quinoline-π*), predicted to lie at 528 nm (oscillator strength = 0.04 au), and so coincides reasonably closely with the lowest energy shoulder feature seen in Fig. 4. A set of stronger bands centred around 370 nm (387 nm, 0.15 au; 375 nm, 0.17 au; 359 nm, 0.10 au) is also predicted, again in good agreement with Fig. 4. These bands consist of varying combinations of Ir-5d and epqc-π orbitals excited into epqc-π* orbitals, but have no contribution from the ancillary bpy-π* orbitals. In comparison, TD-DFT simulation of 3c in MeCN results in broadly the same pattern of predicted bands at 520 nm (0.04 au), 406 nm (0.06 au), 389 nm (0.07 au) and 384 nm (0.09 au), although the higher energy bands are reduced in intensity relative to the low energy MLCT band. In this complex, two low energy bands involving excitation from HOMO (Ir 5d + phenyl-π) to LUMO and LUMO + 1 (bipy-π* and quinoline-π*, respectively) are predicted at 564 and 522 nm, but with very low intensity (<0.001 au), suggesting the possibility of MLCT/LLCT character.


Normalised luminescence emission spectra of selected [Ir(epqc)2(N^N)]PF6 complexes in MeCN solutions (5 × 10−5 M).
Fig. 4 Normalised luminescence emission spectra of selected [Ir(epqc)2(N^N)]PF6 complexes in MeCN solutions (5 × 10−5 M).
Table 2 Calculated excitation wavelengths for 3a and 4a from TD-DFT studies showing the dominant transitions
3a (ester) 4a (acid) 4a (carboxylate)
λ (nm) f Character λ (nm) f Character λ (nm) f Character
a Excitation wavelength. b Oscillator strength.
528 0.0422 HOMO → LUMO + 1 543 0.0348 HOMO → LUMO + 1 456 0.056 HOMO → LUMO
411 0.0538 HOMO − 1 → LUMO 420 0.0509 HOMO − 1 → LUMO 406 0.021 HOMO − 3 → LUMO
387 0.1495 HOMO − 2 → LUMO + 1 394 0.1245 HOMO − 2 → LUMO + 1 381 0.056 HOMO − 3 → LUMO + 1
385 0.0469 HOMO − 2 → LUMO 392 0.0421 HOMO − 2 → LUMO 366 0.115 HOMO − 6 → LUMO + 2
375 0.1743 HOMO − 3 → LUMO 383 0.1547 HOMO − 3 → LUMO 363 0.060 HOMO − 3 → LUMO + 3
359 0.1040 HOMO − 4 → LUMO 366 0.1732 HOMO − 4 → LUMO + 1 356 0.045 HOMO − 7 → LUMO + 1
HOMO − 4 → LUMO + 1
359 0.0815 HOMO − 4 → LUMO 364 0.0478 HOMO − 4 → LUMO 355 0.037 HOMO − 7 → LUMO + 2
HOMO − 4 → LUMO + 1 HOMO − 1 → LUMO + 2


Deprotection of the ester to yield complex 4a was modelled using both the neutral acid and deprotonated carboxylate form. In the neutral complex, calculations predict a red-shift of the absorption bands discussed above by ca. 15 nm, and alteration of the relative intensities of the absorptions, with predicted bands lying at 544 nm (oscillator strength of 0.04 au), 394 nm (0.12 au), 383 nm (0.15 au) and 365 nm (0.18 au). In contrast, simulation of the deprotonated carboxylate form of 4a indicates a substantial blue-shift of the absorption bands, with the lowest energy absorption with significant oscillator strength coming at 456 nm (0.6 au), and a more intense band at 365 nm (0.12 au). The latter approach is in much better agreement with the experimental data, such that modelling 4a as the bis-carboxylate form seems to be more appropriate.

Steady state luminescence measurements were conducted on aerated MeCN solutions, irradiating the 1MLCT wavelength (λex = 450 nm) absorption (Table 3, Fig. 4). The emission maxima for complexes 3a–g show very little variation (627–630 nm), are broad and featureless, and typically characteristic of MLCT-based transitions.24 This variation is less than that predicted by TD-DFT (see ESI), possibly suggesting a diminished diimine contribution to the excited state. Corresponding excitation spectra showed that the complexes could be excited up to a wavelength of ca. 520 nm. Time-resolved emission lifetime measurements revealed that the decays were single-exponential, in each case ca. 200 ns, typical of 3MLCT character. The complexes each exhibited modest quantum yields (Φ) in aerated MeCN, in line with related species. For 3h, the emission profile was very different, with a higher energy peak at 422 nm (τ < 5 ns), which was assigned to a ligand-centred transition arising from the anthraquinone chromophore, with no evidence of a comparable MLCT transition. The absence of 3MLCT emission is attributed to the quenching of that state by the anthraquinone-based ancillary ligand.

Table 3 Photophysical properties of [Ir(epqc)2(N^N)]PF6 (3a–h) complexesa
Complex λ abs (nm) λ em (nm) τ , (ns) Φ
a MeCN solution. b λ ex = 459 nm. c Not measured.
3a 470 (4500), 350 (22[thin space (1/6-em)]000), 292 (39[thin space (1/6-em)]500), 268 (45[thin space (1/6-em)]650) 627 219 0.022
3b 467 (4200), 350 (23[thin space (1/6-em)]500), 290 (46[thin space (1/6-em)]500), 265 (55[thin space (1/6-em)]300) 630 211 0.020
3c 464 (2850), 352 (18[thin space (1/6-em)]150), 291 (39[thin space (1/6-em)]500), 267 (47[thin space (1/6-em)]850) 620 189 0.020
3d 460 (4500), 356 (25[thin space (1/6-em)]750), 288 (52[thin space (1/6-em)]750), 269 (65[thin space (1/6-em)]350) 630 178 0.029
3e 465 (3550), 344 (23[thin space (1/6-em)]300), 290 (55[thin space (1/6-em)]950), 269 (59[thin space (1/6-em)]000) 628 181 0.019
3f 463 (3650), 364 (29[thin space (1/6-em)]700), 272 (82[thin space (1/6-em)]550) 628 210 0.019
3g 461 (5450), 416 (12[thin space (1/6-em)]100), 398 (13[thin space (1/6-em)]550), 327 (58[thin space (1/6-em)]600), 292 (51[thin space (1/6-em)]900), 266 (68[thin space (1/6-em)]200) 628 173 0.018
3h 460 (4550), 367 (29[thin space (1/6-em)]500), 278 (83[thin space (1/6-em)]750) 426 2.1 (57%), <1 (43%) c


For ease of comparison the corresponding emission and excitation of the deprotected complexes 4a–g were obtained in MeCN, MeOH and water. With the exception of 4f these species all showed a hypsochromic shift of ca. 50 nm in the 3MLCT emission maxima (Table 4). Similar measurements in water resulted in an emission peak at ca. 595 nm, revealing the solvent-sensitivity and dipolar nature of the excited state. Emission wavelengths in methanol were typically intermediate between those for water and acetonitrile (for example, Fig. 5). Relative to the esterified analogues (3a–g), the measured lifetimes were generally extended for complexes 4a–g reflecting the increased energy gap. However, in water the lifetime values also varied greatly as a function of the type of diimine ligand, with 4e (likely to be the most hydrophobic of the diimines in this study) displaying the longest lifetime (τ = 619 ns), suggesting greater shielding of the excited state from the surrounding solvent.


Normalised luminescence emission spectra of 4g in various solvents (5 × 10−5 M).
Fig. 5 Normalised luminescence emission spectra of 4g in various solvents (5 × 10−5 M).
Table 4 Photophysical properties of [Ir(pqca)2(N^N)]Cl (4a–g) complexesa
  CH3CN H2O CH3OH
Complex λ em/nm τ/ns λ em/nm τ/ns λ em/nm τ/ns
a λ ex = 459 nm.
4a 569 211 592 261 583 414
4b 572 320 594 223 586 383
4c 574 258 589 190 582 385
4d 574 439 590 201 584 446
4e 577 322 608 619 585 370
4f 631 174 630 58 630 82
4g 568 113 598 199 584 297


The longer emission wavelength of 4f in all solvents appears somewhat anomolous and could be due to a number of factors. With reference to the TD-DFT calculations, the protonation state of the cinchophen ligands influences the emission wavelength. Repeating the measurement in 0.1 M NaOH blue-shifted the emission peak to ca. 590 nm, in accordance with the other complexes in the cinchophen-based series. However, it is difficult to rationalise why only 4f (versus4g, for example) would retain a protonated carboxylic acid form. An alternative explanation considers the role of protonation at the dppz ligand. The closely related species [Ir(ppy)2(dppz)]+[thin space (1/6-em)]25 reveals λem = 630 nm in MeCN and is thus very comparable to 4f. However, H-bonding interactions with the phenazine nitrogens can dramatically influence the emission properties; in fact, [Ir(ppy)2(dppz)]+ is non-emissive in water, unlike 4f. The reported photophysics of [Ir(ppy)2(dppn)]+[thin space (1/6-em)]25 also compare very well (λem = 583 nm in MeOH) to 4g, and thus the differences in emission wavelength for 4f could be due to protonation interactions at dppz26 (of course, these would also be sensitive to the addition of 0.1 M NaOH).

Experimental section

All reactions were performed with the use of vacuum line and Schlenk techniques. Reagents were commercial grade and used without further purification. 1H and 13C–{1H} NMR spectra were recorded on an NMR-FT Bruker 400 or 250 MHz and 31P–{1H} NMR spectra on a Joel Eclipse 300 MHz spectrometer and recorded in CDCl3 or MeOD solutions. 1H and 13C–{1H} NMR chemical shifts (δ) were determined relative to internal tetramethylsilane, Si(CH3)4 and are given in ppm. Low-resolution mass spectra were obtained by the staff at Cardiff University. High-resolution mass spectra were carried out at the EPSRC National Mass Spectrometry Service at Swansea University, UK. UV-Vis studies were performed on a Jasco V-650 spectrophotometer fitted with a Jasco temperature control unit in MeCN or MeOH solutions (5 × 10−5 M) at 20 °C. Photophysical data were obtained on a JobinYvon–Horiba Fluorolog spectrometer fitted with a JY TBX picoseconds photodetection module in MeCN, MeOH or H2O solutions. Emission spectra were uncorrected and excitation spectra were instrument corrected. The pulsed source was a Nano-LED configured for 372 or 459 nm output operating at 500 kHz. Luminescence lifetime profiles were obtained using the JobinYvon–Horiba FluoroHub single photon counting module and the data fits yielded the lifetime values using the provided DAS6 deconvolution software. Electrochemical studies were carried out using a Parstat 2273 potentiostat in conjunction with a three-electrode cell. The auxiliary electrode was a platinum wire and the working electrode a platinum (1.0 mm diameter) disc. The reference was a silver wire separated from the test solution by a fine porosity frit and an agar bridge saturated with KCl. Solutions (10 ml CH2Cl2) were 1.0 × 10−3 mol dm−3 in the test compound and 0.1 mol dm−3 in [NBun4][PF6] as the supporting electrolyte. Under these conditions, E0, for the one-electron oxidation of [Fe(η-C5H5)2] added to the test solutions as an internal calibrant, is +0.46 V in CH2Cl2.27 Unless specified, all electrochemical values are at υ = 200 mV s−1. Microanalyses were performed by London Metropolitan University, UK.

Data collection and processing

Diffraction data for 3b and 3e were collected on a Nonius KappaCCD using graphite-monochromated Mo-Kα radiation (λ = 0.71073 Å) at 150 K. Software package Apex 2 (v2.1) was used for the data integration, scaling and absorption correction.

Structure analysis and refinement

The structure was solved by direct methods using SHELXS-97 and was completed by iterative cycles of ΔF-syntheses and full-matrix least squares refinement. All non-H atoms were refined anisotropically and difference Fourier syntheses were employed in positioning idealised hydrogen atoms and were allowed to ride on their parent C-atoms. All refinements were against F2 and used SHELX-97.28 CCDC reference numbers 907280 and 907282 contain the supplementary crystallographic data for this paper.

DFT studies

DFT geometry optimisation and orbital calculations were performed on the Gaussian 03 program.29 Geometry optimisations were carried out without constraints using the B3PW91 functional. The LANL2DZ30 basis set was used for the Ir centers, and was invoked with pseudo-potentials for the core electrons, a 6-31G(d,p)31 basis set for all coordinating atoms with a 6-31G32 basis set for all remaining atoms. All optimisations were followed by frequency calculations to ascertain the nature of the stationary point (minimum or saddle point). TD-DFT studies were performed in Gaussian0933 using the same functional, but with 6-31G(d) on all non-metal atoms, and also included a simulated MeCN or H2O environment using the polarised continuum model (PCM) approach.34 For prediction of absorption spectra, the geometry used to calculate orbital and other properties was used without modification. For prediction of emission, however, the triplet state was allowed to relax to its optimal geometry using unrestricted B3PW91 in the gas phase, prior to solvated TD-DFT.

Synthesis

The ligands ethyl 2-phenylquinoline-4-carboxylate (epqcH) 1,35 dipyrido[3,2-a:2′,3′-c]phenazine (dppz),36 benzo[i]dipyrido[3,2-a:2′,3′-c]phenazine (dppn)37 and naphtha[2,3-a]dipyrido[3,2-h:2′,3′-f]phenazine-5,18-dione (qdppz)38 were prepared according to reported procedures.

[Ir(epqc)2(μ-Cl)2Ir(epqc)2] 2 was prepared by variation of the standard literature procedures for other bridged-chloride dimers.19 Thus, IrCl3·3H2O (0.266 g, 0.75 mmol) and epqcH (0.520 g, 2.02 mmol) in 2-methoxyethanol–water (3[thin space (1/6-em)]:[thin space (1/6-em)]1, 8 ml) were heated at reflux for 48 h. Water (25 ml) was added to give a dark purple precipitate, which was filtered and dried in vacuo. The product was used in subsequent reaction without further purification.24

[Ir(epqc)2(bpy)]PF63a. [(epqc)2Ir(μ-Cl)2Ir(epqc)2] 2 (0.061 g, 0.04 mmol) and 2,2′-bipyridine (0.013 g, 0.08 mmol) in 2-methoxyethanol (5 ml) were heated at 120 °C for 16 h. The solvent was then removed in vacuo and the crude product dissolved in MeCN (4 ml). KPF6 (1.05 g, 5.705 mmol) in water (2 ml) was added and the solution stirred for 10 min. Water (20 ml) was added and the product extracted with CH2Cl2 (2 × 20 ml). The combined organic phases were washed with water (30 ml) and brine (30 ml) before being dried over MgSO4. The solution was filtered and the solvent removed in vacuo. The crude product was then purified by column chromatography (silica, CH2Cl2). After elution of unreacted organics with CH2Cl2 the product was eluted as the first red fraction with CH2Cl2MeOH (9[thin space (1/6-em)]:[thin space (1/6-em)]1). The product was concentrated in volume (to ca. 3 ml) and precipitated by the slow addition of Et2O (5 ml) and dried in vacuo. Yield = 0.055 g, 76%. 1H NMR (250 MHz, CDCl3) δH = 8.51 (2H, s), 8.42 (2H, d, 3JHH = 5.8 Hz), 8.23 (2H, d, 3JHH = 7.5 Hz), 8.05–7.8 (6H, m), 7.70–7.35 (6H, m), 7.21 (2H, app. t, 3JHH = 7.6 Hz), 7.02 (2H, app. t, 3JHH = 7.5 Hz), 6.86 (2H, app. t, 3JHH = 7.6 Hz), 6.53 (2H, d, 3JHH = 7.6 Hz), 4.50 (4H, m), 1.52 (6H, m) ppm. 13C–{1H} NMR (75 MHz, CDCl3) δC = 169.6, 166.4, 156.9, 148.7, 147.4, 145.0, 140.2, 138.3, 131.6, 130.4, 130.1, 129.1, 128.1, 127.8, 126.9, 125.5, 125.1, 124.1, 123.5, 120.6, 116.6, 62.2, 14.4 ppm. UV-vis (MeCN): λmax (ε/dm3 mol−1 cm−1) 470 (4500), 350 (22[thin space (1/6-em)]000), 292 (39[thin space (1/6-em)]500), 268 (45[thin space (1/6-em)]650) nm. Elemental analysis: Calcd (%) for C46H36N4O4IrPF6: C, 52.82, H, 3.47, N, 5.36; Found: C, 52.90, H, 3.48, N, 5.36. ES MS found m/z 901.3, calculated m/z 901.0 for [M − PF6]+. HR MS found m/z 901.2370, calculated m/z 901.2363 for [C46H36N4O4191Ir]+. IR (solid): ν 1722 (CO), 837 (PF6) cm−1.
[Ir(epqc)2(dmbpy)]PF63b. Prepared similarly from 2 (0.069 g, 0.037 mmol) and 4-4′-dimethyl-2,2′-bipyridine (dmbpy) (0.0136 g, 0.074 mmol). Yield = 0.058 g, 85%. 1H NMR (250 MHz, CDCl3) δH = 8.51 (2H, s), 8.47 (2H, d, 3JHH = 7.9 Hz), 8.01–7.81 (4H, m), 7.72 (2H, d, 3JHH = 7.9 Hz), 7.52–7.31 (4H, m), 7.2–7.0 (4H, m), 6.95 (2H, app. t, 3JHH = 8.5 Hz). 6.83 (2H, app. t, 3JHH = 7.3 Hz), 6.50 (2H, d, 3JHH = 7.3 Hz), 4.62 (4H, q, 3JHH = 7.1 Hz), 2.47 (6H, s), 1.56 (6H, t, 3JHH = 7.2 Hz) ppm. 13C–{1H} NMR (75 MHz, CDCl3) δC = 169.7, 165.2, 155.2, 152.7, 148.9, 146.6, 145.1, 138.7, 134.8, 131.5, 129.8, 129.5, 128.8, 127.6, 127.2, 126.4, 125.3, 123.3, 122.4, 120.2, 118.6, 65.9, 21.3, 15.4 ppm. UV-vis (MeCN): λmax (ε/dm3 mol−1 cm−1) 467 (4200), 350 (23[thin space (1/6-em)]500), 290 (46[thin space (1/6-em)]500), 265 (55[thin space (1/6-em)]300) nm. Elemental analysis: Calcd (%) for C48H40N4O4IrPF6: C, 53.68, H, 3.75, N, 5.22; Found: C, 53.86, H, 3.79, N, 5.26. ES MS found m/z 929.3, calculated m/z 929.3 for [M − PF6]+. HR MS found m/z 927.2650, calculated m/z 927.2650 for [C48H40N4O4191Ir]+. IR (solid): ν 1720 (CO), 829 (PF6) cm−1.
[Ir(epqc)2(dmbpc)]PF63c. Prepared similarly from 2 (0.069 g, 0.049 mmol) and diethyl-2,2′-bipyridine-4,4′-dicarboxylate (0.028 g, 0.098 mmol). Yield = 0.072 g, 74%. 1H NMR (250 MHz, CDCl3) δH = 8.67 (2H, d, 3JHH = 7.5 Hz), 8.58 (2H, d, 3JHH = 7.6 Hz), 8.43–8.34 (2H, m), 8.16–8.05 (4H, m), 7.55 (2H, d, 3JHH = 7.5 Hz), 7.45 (2H, app. t, 3JHH = 8.7 Hz), 7.35–7.19 (4H, m), 7.08 (2H, app. t, 3JHH = 8.7 Hz), 6.88 (2H, app. t, 3JHH = 7.6 Hz), 6.51 (2H, d, 3JHH = 7.6 Hz), 4.52 (4H, q, 3JHH = 7.2 Hz), 4.46 (4H, br t), 3.68 (4H, br t), 3.30 (6H, s), 1.48 (6H, t, 3JHH = 7.2 Hz) ppm. 13C–{1H} NMR (75 MHz, CDCl3) δC = 169.6, 166.4, 162.9, 156.7, 150.1, 148.4, 147.1, 144.9, 140.7, 139.0, 134.8, 131.6, 130.0, 129.0, 127.7, 127.1, 124.9, 124.0, 123.2, 120.5, 118.8, 116.6, 70.3, 65.6, 62.9, 61.9, 15.2 ppm. UV-vis (MeCN): λmax (ε/dm3 mol−1 cm−1) 464 (2900), 352 (18[thin space (1/6-em)]200), 291 (39[thin space (1/6-em)]500), 267 (47[thin space (1/6-em)]900) nm. Elemental analysis: Calcd (%) for C54H48N4O10IrPF6: C, 51.88, H, 3.87, N, 4.48; Found: C, 51.74, H, 3.84, N, 4.46. ES MS found m/z 1105.3, calculated m/z 1105.2 for [M − PF6]+. HR MS found m/z 1105.2977, calculated m/z 1103.2977 for [C54H48N4O10191Ir]+. IR (solid): ν 1724 (CO), 835 (PF6) cm−1.
[Ir(epqc)2(phen)]PF63d. Prepared similarly from 2 (0.064 g, 0.041 mmol) and 1,10-phenanthroline monohydrate (0.015 g, 0.082 mmol). Yield = 0.060 g, 82%. 1H NMR (250 MHz, CDCl3) δH = 8.66 (2H, d, 3JHH = 7.6 Hz), 8.52–8.36 (4H, m), 8.23 (2H, d, 3JHH = 7.6 Hz), 8.16 (2H, d, 3JHH = 7.5 Hz), 7.92–7.77 (4H, m), 7.70–7.51 (4H, m), 7.33 (2H, d, 3JHH = 8.2 Hz), 6.90 (2H, app. t, 3JHH = 7.6 Hz), 6.80 (2H, app. t, 3JHH = 7.6 Hz), 6.63 (2H, d, 3JHH = 7.5 Hz), 4.58 (4H, q, 3JHH = 7.3 Hz), 1.53 (6H, t, 3JHH = 7.3 Hz) ppm. 13C–{1H} NMR (75 MHz, CDCl3) δC = 169.6, 164.9, 156.9, 149.6, 148.9, 147.9, 144.6, 138.6, 136.9, 135.0, 130.7, 130.0, 129.7, 128.8, 127.7, 127.4, 126.6, 125.2, 124.3, 123.5, 120.3, 118.6, 62.0, 14.1 ppm. UV-vis (MeCN): λmax (ε/dm3 mol−1 cm−1) 460 (4500), 356 (25[thin space (1/6-em)]800), 288 (52[thin space (1/6-em)]800), 269 (65[thin space (1/6-em)]400) nm. Elemental analysis: Calcd (%) for C48H36N4O4IrPF6: C, 53.88, H, 3.39, N, 5.24; Found: C, 53.75, H, 3.51, N, 5.28. ES MS found m/z 925.2, calculated m/z 925.2 for [M − PF6]+. HR MS found m/z 923.2345, calculated m/z 923.2337 for [C48H36N4O4191Ir]+. IR (solid): ν 1722 (CO), 833 (PF6) cm−1.
[Ir(epqc)2(dip)]PF63e. Prepared similarly from 2 (0.060 g, 0.037 mmol) and 4,7-diphenyl-1,10-phenanthroline (0.025 g, 0.074 mmol). Yield = 0.060 g, 78%. 1H NMR (250 MHz, CDCl3) δH = 8.61 (2H, s), 8.51 (2H, d, 3JHH = 5.6 Hz), 8.48 (2H, d, 3JHH = 8.2 Hz), 8.16 (4H, d, 3JHH = 7.5 Hz), 7.81 (2H, s), 7.78 (2H, d, 3JHH = 5.6 Hz), 7.51–7.34 (10H, m), 7.38–7.28 (4H, m), 6.9–6.75 (4H, m), 6.61 (2H, d, 3JHH = 7.5 Hz), 4.60 (4H, q, 3JHH = 7.1 Hz), 1.55 (6H, t, 3JHH = 7.1 Hz) ppm. 13C–{1H} NMR (75 MHz, CDCl3) δC = 170.0, 165.2, 156.7, 151.3, 148.6, 147.2, 145.3, 138.8, 135.2, 131.6, 130.9, 130.1, 129.9, 129.3, 128.7, 128.2, 127.9, 126.9, 126.7, 126.0, 125.6, 124.8, 124.4, 123.8, 120.5, 119.1, 63.0, 14.4 ppm. UV-vis (MeCN): λmax (ε/dm3 mol−1 cm−1) 465 (3600), 344 (23[thin space (1/6-em)]300), 290 (55[thin space (1/6-em)]900), 269 (59[thin space (1/6-em)]000) nm. Elemental analysis: Calcd (%) for C60H44N4O4IrPF6: C, 58.96, H, 3.63, N, 4.58; Found: C, 58.90, H, 3.57, N, 4.54. ES MS found m/z 1077.3, calculated m/z 1077.2 for [M − PF6]+. HR MS found m/z 1075.2947, calculated m/z 1075.2963 for [C60H44N4O4191Ir]+. IR (solid): ν 1721 (CO), 835 (PF6) cm−1.
[Ir(epqc)2(dppz)]PF63f. Prepared similarly from 2 (0.072 g, 0.046 mmol) and dppz (0.028 g, 0.099 mmol). Yield = 0.081 g, 75%. 1H NMR (250 MHz, CDCl3) δH = 9.46 (1H, d, 3JHH = 8.0 Hz), 9.39 (1H, d, 3JHH = 8.1 Hz), 9.14 (1H, br s), 8.62–8.55 (2H, m), 8.46 (1H, d, 3JHH = 4.8 Hz), 8.30 (1H, app. t, 3JHH = 4.1 Hz), 8.20–7.90 (4H, m), 7.85 (2H, app. t, 3JHH = 6.7 Hz), 7.80–7.75 (2H, m), 7.71–7.61 (2H, m), 7.49–7.42 (2H, m), 7.30–7.23 (4H, m), 7.16–7.06 (2H, m), 6.80 (2H, app. t, 3JHH = 7.7 Hz), 6.56 (1H, d, 3JHH = 7.6 Hz), 4.46 (4H, q, 3JHH = 7.1 Hz), 1.55 (6H, t, 3JHH = 7.1 Hz) ppm. 13C–{1H} NMR (75 MHz, CDCl3) δC = 169.7, 164.8, 156.3, 152.1, 149.6, 148.7, 148.3, 145.2, 142.8, 138.9, 138.6, 136.0, 135.1, 132.4, 131.5, 130.4, 129.5, 128.6, 127.9, 127.3, 126.8, 124.5, 123.8, 118.9, 62.9, 14.2 ppm. UV-vis (MeCN): λmax (ε/dm3 mol−1 cm−1) 463 (3600), 364 (29[thin space (1/6-em)]700), 272 (82[thin space (1/6-em)]600) nm. Elemental analysis: Calcd (%) for C54H38N6O4IrPF6: C, 55.34, H, 3.27, N, 7.17; Found: C, 55.42, H, 3.33, N, 7.08. ES MS found m/z 1027.3, calculated m/z 1027.3 for [M − PF6]+. HR MS found m/z 1025.2561, calculated m/z 1025.2555 for [C54H38N6O4191Ir]+. IR (solid): ν 1724 (CO), 830 (PF6) cm−1.
[Ir(epqc)2(dppn)]PF63g. Prepared similarly from 2 (0.064 g, 0.041 mmol) and dppn (0.028 g, 0.084 mmol). Yield = 0.078 g, 78%. 1H NMR (250 MHz, CDCl3) δH = 9.43 (1H, d, 3JHH = 8.2 Hz), 8.92 (1H, br s), 8.76 (1H, s), 8.66 (2H, app. t, 3JHH = 4.8 Hz), 8.55 (1H, d, 3JHH = 4.8 Hz), 8.48–8.41 (2H, m), 8.38–8.31 (2H, m), 8.22 (1H, d, 3JHH = 6.7 Hz), 8.14–8.06 (2H, m), 8.03–7.94 (2H, m), 7.55–7.16 (7H, m), 7.10–6.95 (3H, m), 6.92–6.80 (3H, m), 6.71–6.54 (2H, m), 4.50 (4H, q, 3JHH = 7.1 Hz), 1.48 (6H, t, 3JHH = 7.1 Hz) ppm. 13C–{1H} NMR (75 MHz, CDCl3) δC = 170.2, 165.0, 150.5, 149.8, 148.7, 148.5, 145.2, 138.9, 138.4, 136.5, 136.0, 135.2, 133.7, 131.8, 131.3, 130.0, 129.0, 128.8, 128.2, 128.1, 127.6, 127.1, 126.7, 125.3, 124.7, 124.0, 119.6, 63.0, 14.4 ppm. UV-vis (MeCN): λmax (ε/dm3 mol−1 cm−1) 461 (5400), 416 (12[thin space (1/6-em)]100), 398 (13[thin space (1/6-em)]500), 327 (58[thin space (1/6-em)]600), 292 (51[thin space (1/6-em)]900), 266 (68[thin space (1/6-em)]200) nm. Elemental analysis: Calcd (%) for C58H40N6O4IrPF6: C, 57.00, H, 3.30, N, 6.88; Found: C, 57.01, H, 3.34, N, 6.76. ES MS found m/z 1077.2, calculated m/z 1077.3 for [M − PF6]+. HR MS found m/z 1075.2715, calculated m/z 1075.2711 for [C58H40N6O4191Ir]+. IR (solid): ν 1719 (CO), 828 (PF6) cm−1.
[Ir(epqc)2(qdppz)]PF63h. Prepared similarly from 2 (0.068 g, 0.044 mmol) and qdppz (0.037 g, 0.090 mmol). Yield = 0.094 g, 83%. 1H NMR (250 MHz, CDCl3) δH = 9.38 (2H, d, 3JHH = 5.1 Hz), 8.64 (2H, s), 8.61 (2H, d, 3JHH = 5.0 Hz), 8.34 (2H, dd, 3JHH = 8.6 and 4.1 Hz), 8.13 (2H, app. t, 3JHH = 6.90 Hz), 8.07–7.92 (3H, m), 7.70 (2H, app. t, 3JHH = 8.9 Hz), 7.56 (1H, d, 3JHH = 7.1 Hz), 7.44 (1H, app. t, 3JHH = 7.1 Hz), 7.38 (2H, app. t, 3JHH = 7.9 Hz), 7.28 (1H, app. t, 3JHH = 7.4 Hz), 7.19 (2H, app. t, 3JHH = 7.6 Hz), 7.15–7.02 (2H, m), 6.94 (2H, app. t, 3JHH = 7.5 Hz), 6.86 (2H, app. t, 3JHH = 7.3 Hz), 6.62 (2H, app. t, 3JHH = 5.74 Hz), 4.50 (4H, q, 3JHH = 7.2 Hz), 1.42 (6H, t, 3JHH = 7.2 Hz) ppm. 13C–{1H} NMR (75 MHz, CDCl3) δC = 166.4, 156.7, 148.9, 138.5, 136.5, 130.2, 130.0, 129.0, 127.9, 127.6, 125.5, 124.1, 120.4, 120.3, 116.5, 62.1, 14.4 ppm. UV-vis (MeCN): λmax (ε/dm3 mol−1 cm−1) 460 (4550), 367 (29[thin space (1/6-em)]500), 278 (83[thin space (1/6-em)]750), 212 (85[thin space (1/6-em)]650) nm. Elemental analysis: Calcd (%) for C62H40N6O6IrPF6: C, 57.19, H, 3.10, N, 6.45; Found: C, 57.12, H, 3.18, N, 6.40. ES MS found m/z 1157.3, calculated m/z 1157.2 for [M − PF6]+. HR MS found m/z 1155.2587, calculated m/z 1155.2610 for [C62H40N6O6191Ir]+. IR (solid): ν 1722 (CO), 1610 (CO), 832 (PF6) cm−1.
[Ir(pqca)2(bpy)]Cl 4a. Complex 3a (0.036 g, 0.020 mmol) was dissolved in acetone (5 ml) and KOH (1 M soln, 5 ml) added. The mixture was stirred at room temp. for 14 h. The solvent was removed in vacuo and water (20 ml) added, followed by neutralisation with HCl (1 M soln). Water was removed in vacuo and product was dissolved in MeOH (10 ml). The solution was filtered to remove salts and the solvent removed in vacuo. Yield = 0.022 g, 65%. 1H NMR (400 MHz, MeOD) δH = 8.37 (2H, s), 8.20–8.15 (4H, m), 8.12 (2H, d, 3JHH = 7.7 Hz), 8.06 (2H, d, 3JHH = 7.7 Hz), 7.89 (2H, app. t, 3JHH = 8.0 Hz), 7.44 (2H, app. t, 3JHH = 7.3 Hz), 7.30 (2H, d, 3JHH = 7.6 Hz), 7.23 (2H, app. t, 3JHH = 7.5 Hz), 7.03 (2H, app. t, 3JHH = 7.7 Hz), 6.87 (2H, app. t, 3JHH = 7.7 Hz), 6.69 (2H, app. t, 3JHH = 7.4 Hz), 6.53 (2H, d, 3JHH = 7.7 Hz) ppm. UV-vis (MeOH): λmax (ε/dm3 mol−1 cm−1) 453, (1200), 333 (11[thin space (1/6-em)]200), 285 (12[thin space (1/6-em)]350) nm. ES MS found m/z 845.1, calculated m/z 844.9 for [M − Cl]+. HR MS found m/z 843.1710, calculated m/z 843.1711 for [C42H28N4O4191Ir]+. IR (solid): ν 1578 (CO) cm−1.
[Ir(pqca)2(dmbpy)]Cl 4b. Prepared similarly from 3b (0.050 g, 0.088 mmol). Yield = 0.041 g, 80%. 1H NMR (250 MHz, MeOD) δH = 8.67 (2H, s), 8.70 (2H, d, 3JHH = 5.8 Hz), 8.46 (2H, d, 3JHH = 8.2 Hz), 8.29 (2H, d, 3JHH = 8.2 Hz), 8.19 (2H, d, 3JHH = 7.6 Hz), 7.98 (2H, s), 7.95 (2H, app. t, 3JHH = 6.2 Hz), 7.73 (2H, app. t, 3JHH = 7.0 Hz), 7.44 (2H, d, 3JHH = 8.4 Hz), 7.27 (2H, app. t, 3JHH = 7.6 Hz), 6.82 (2H, app. t, 3JHH = 7.1 Hz), 6.67 (2H, d, 3JHH = 7.6 Hz), 2.57 (6H, s) ppm. 13C–{1H} NMR (75 MHz, MeOD) δC = 171.1, 155.5, 1519, 151.2, 147.9, 147.0, 146.2, 134.4, 130.2, 129.6, 129.3, 128.6, 127.6, 127.4, 126.9, 126.2, 125.0, 124.5, 124.3, 122.6, 114.3, 19.8 ppm. UV-vis (MeOH): λmax (ε/dm3 mol−1 cm−1) 450 (1600), 341 (6000), 286 (16[thin space (1/6-em)]800), 267 (18[thin space (1/6-em)]600) nm. ES MS found m/z 873.2, calculated m/z 873.0 for [M − Cl]+. HR MS found m/z 871.2018, calculated m/z 871.2024 for [C44H32N4O4191Ir]+. IR (solid): ν 1578 (CO) cm−1.
[Ir(pqca)2(bpdc)]Cl 4c. Prepared similarly from 3c (0.055 g, 0.053 mmol). Yield = 0.044 g, 91%. 1H NMR (250 MHz, MeOD) δH = 8.68 (2H, br s), 8.38 (2H, s), 8.27 (2H, d, 3JHH = 5.7 Hz), 8.19 (2H, app. t, 3JHH = 9.1 Hz), 7.92 (2H, s), 7.90 (2H, d, 3JHH = 5.7 Hz), 7.43–7.31 (4H, m), 7.17 (2H, app. t, 3JHH = 7.7 Hz), 6.98 (2H, app. t, 3JHH = 7.9 Hz), 6.80 (2H, app. t, 3JHH = 7.1 Hz), 6.53 (2H, d, 3JHH = 7.2 Hz) ppm. 13C–{1H} NMR (75 MHz, MeOD) δC = 172.2, 170.4, 160.4, 157.8, 151.0, 150.7, 148.1, 147.2, 146.3, 139.3, 135.4, 134.7, 130.3, 129.9, 129.6, 128.8, 128.6, 127.5, 127.0, 126.7, 126.3, 125.5, 124.4, 122.8, 116.5, 114.5 ppm. UV-vis (MeOH): λmax (ε/dm3 mol−1 cm−1) 449 (1300), 338 (5000), 287 (15[thin space (1/6-em)]700), 262 (20[thin space (1/6-em)]400) nm. ES MS found m/z 933.1, calculated m/z 932.9 for [M − Cl]+. HR MS found m/z 931.1507, calculated m/z 931.1508 for [C44H28N4O8191Ir]+. IR (solid): ν 1578 (CO) cm−1.
[Ir(pqca)2(phen)]Cl 4d. Prepared similarly from 3d (0.040 g, 0.043 mmol). Yield = 0.026 g, 70%. 1H NMR (400 MHz, MeOD) δH = 8.81 (2H, s), 8.62 (2H, d, 3JHH = 5.8 Hz), 8.46 (2H, app. t, 3JHH = 8.2 Hz), 8.29 (2H, d, 3JHH = 8.2 Hz), 8.19 (2H, d, 3JHH = 7.6 Hz), 7.98 (2H, s), 7.95 (2H, app. t, 3JHH = 6.2 Hz), 7.73 (2H, app. t, 3JHH = 7.0 Hz), 7.44 (2H, d, 3JHH = 8.4 Hz), 7.27 (2H, app. t, 3JHH = 7.6 Hz), 6.91 (2H, app. t, 3JHH = 7.4 Hz), 6.82 (2H, app. t, 3JHH = 7.1 Hz), 6.67 (2H, d, 3JHH = 7.6 Hz) ppm. UV-vis (MeOH): λmax (ε/dm3 mol−1 cm−1) 445 (2600), 339 (11[thin space (1/6-em)]000), 266 (37[thin space (1/6-em)]400) nm. ES MS found m/z 869.1, calculated m/z 868.9 for [M − Cl]+. HR MS found m/z 867.1704, calculated m/z 867.1711 for [C44H28N4O4191Ir]+. IR (solid): ν 1597 (CO) cm−1.
[Ir(pqca)2(dip)]Cl 4e. Prepared similarly from 3e (0.030 g, 0.024 mmol). Yield 0.018 g, 76%. 1H NMR (400 MHz, MeOD) δH = 8.84 (2H, s), 8.69 (2H, d, 3JHH = 7.6 Hz), 8.53 (2H, d, 3JHH = 7.8 Hz), 8.31 (2H, d, 3JHH = 7.6 Hz), 7.93 (4H, app. t, 3JHH = 7.2 Hz), 7.59–7.49 (12H, m), 7.35 (2H, app. t, 3JHH = 6.2 Hz), 7.28 (2H, app. t, 3JHH = 6.8 Hz), 6.98–6.88 (4H, m), 6.71 (2H, d, 3JHH = 7.8 Hz) ppm. 13C–{1H} NMR (75 MHz, MeOD) δC = 172.2, 170.1, 151.0, 150.9, 150.7, 148.1, 147.2, 146.3, 135.4, 134.7, 130.3, 130.0, 129.7, 129.6, 128.8, 128.6, 128.5, 127.5, 127.4, 127.0, 126.7, 126.3, 125.5, 124.4, 122.8, 116.5, 114.5 ppm. UV-vis (MeOH): λmax (ε/dm3 mol−1 cm−1) 448 (1700), 402 (1400), 336 (7900), 288 (22[thin space (1/6-em)]400), 266 (22[thin space (1/6-em)]600) nm. ES MS found m/z 1021.2, calculated m/z 1021.1 for [M − Cl]+. HR MS found m/z 1019.2333, calculated m/z 1019.2337 for [C56H36N4O4191Ir]+. IR (solid): ν 1576 (CO) cm−1.
[Ir(pqca)2(dppz)]Cl 4f. Prepared similarly from 3f (0.052 g, 0.044 mmol). Yield 0.035 g, 79%. 1H NMR (400 MHz, MeOD) δH = 9.48 (2H, d, 3JHH = 7.6 Hz), 8.74 (2H, s), 8.62 (2H, d, 3JHH = 5.0 Hz), 8.38–8.30 (2H, m), 8.22 (2H, d, 3JHH = 7.9 Hz), 8.13–8.05 (3H, m), 7.98 (2H, app t), 7.87 (2H, br d), 7.75 (1H, app. t), 7.59 (1H, app. t), 7.50–7.40 (3H, m), 7.21–7.12 (2H, m), 6.88–6.80 (2H, m), 6.61 (2H, d, 3JHH = 7.6 Hz) ppm. UV-vis (MeOH): λmax (ε/dm3 mol−1 cm−1) 435 (1800), 335 (9000), 289 (18[thin space (1/6-em)]400), 265 (23[thin space (1/6-em)]400) nm. HR MS found m/z 969.1914, calculated m/z 969.1929 for [C50H30N6O4191Ir]+. IR (solid): ν 1575 (CO) cm−1.
[Ir(pqca)2(dppn)]Cl 4g. Prepared similarly from 3g (0.048 g, 0.039 mmol). Yield 0.029 g, 71%. 1H NMR (400 MHz, MeOD) δH = 9.49 (2H, d, 3JHH = 7.6 Hz), 8.80 (2H, s), 8.61 (2H, dd, JHH = 5.3 and 1.4 Hz), 8.34–8.28 (4H, m), 8.20–7.85 (4H, overlapping m), 7.51–7.38 (5H, m), 7.33 (2H, d, 3JHH = 7.6 Hz), 7.18–7.05 (3H, m), 6.81–6.70 (4H, m), 6.60 (2H, d, 3JHH = 7.7 Hz) ppm. UV-vis (MeOH): λmax (ε/dm3 mol−1 cm−1) 429 (2600), 330 (17[thin space (1/6-em)]700), 287 (24[thin space (1/6-em)]200), 263 (33[thin space (1/6-em)]800) nm. HR MS found m/z 1019.2089, calculated m/z 1019.2085 for [C54H32N6O4191Ir]+. IR (solid): ν 1578 (CO) cm−1.

Conclusions

Water-soluble, luminescent iridium(III) complexes can be conveniently synthesised through the use of cinchophen-based ligands. The Ir(III) coordination chemistry of ethyl-2-phenylquinoline-4-carboxylate can be achieved using traditional methods and the resultant cyclometalated complexes are tolerant to the subsequent deprotection strategy. The new complexes of the form [Ir(pqca)2(N^N)]+ are luminescent via an excited state that is best described as possessing substantial 3MLCT character; the CT nature was reflected in the solvent-sensitive properties of the complexes. TD-DFT calculations also revealed that careful modelling of the protonated form of the carboxylic acid/carboxylate is necessary to allow a more precise approximation of the electronic characteristics of the complexes.

Acknowledgements

We thank Cardiff University and EPSRC for financial support, and the EPSRC National Mass Spectrometry Service Centre at Swansea University.

Notes and references

  1. (a) R. D. Costa, E. Ortí, H. J. Bolink, F. Monti, G. Accorsi and N. Armaroli, Angew. Chem., Int. Ed., 2012, 51, 8178 CrossRef CAS; (b) A. F. Rausch, H. H. H. Homeier and H. Yersin, Top. Organomet. Chem., 2010, 29, 193 CrossRef CAS; (c) N. M. Shavaleev, F. Monti, R. D. Costa, R. Scopelliti, H. J. Bolink, E. Ortí, G. Accorsi, N. Armaroli, E. Baranoff, M. Grätzel and Md. K. Nazeeruddin, Inorg. Chem., 2012, 51, 2263 CrossRef CAS; (d) A. J. Hallett, N. White, W. Wu, X. Cui, P. N. Horton, S. J. Coles, J. Zhao and S. J. A. Pope, Chem. Commun., 2012, 48, 10838 RSC; (e) J. D. Routledge, A. J. Hallett, J. A. Platts, P. N. Horton and S. J. A. Pope, Eur. J. Inorg. Chem., 2012, 4065 CrossRef CAS; (f) A. J. Hallett, B. M. Kariuki and S. J. A. Pope, Dalton Trans., 2011, 40, 9474 RSC.
  2. U. Scherf and D. Neher, Polyfluorenes, Springer, Berlin, 2008 Search PubMed.
  3. (a) I.-S. Shin, H.-C. Lim, J.-W. Oh, J.-K. Lee, T. H. Kim and H. Kim, Electrochem. Commun., 2011, 13, 64 CrossRef CAS; (b) H.-C. Su, Y.-H. Lin, C.-H. Chang, H.-W. Lin, C.-C. Wu, F.-C. Fang, H.-F. Chen and K.-T. Wong, J. Mater. Chem., 2010, 20, 5521 RSC; (c) L. He, L. Duan, J. Qiao, G. Dong, L. Wang and Y. Qiu, Chem. Mater., 2010, 22, 3535 CrossRef CAS; (d) C. Rothe, C.-J. Chiang, V. Jankus, K. Abdullah, X. Zeng, R. Jitchati, A. S. Batsanov, M. R. Bryce and A. P. Monkman, Adv. Funct. Mater., 2009, 19, 2038 CrossRef CAS; (e) S. Graber, K. Doyle, M. Neuburger, C. E. Housecroft, E. C. Constable, R. D. Costa, E. Orti, D. Repetto and H. J. Bolink, J. Am. Chem. Soc., 2008, 130, 14944 CrossRef CAS; (f) R. D. Costa, E. Ortí and H. J. Bolink, Pure Appl. Chem., 2011, 83, 2115 CrossRef CAS.
  4. (a) E. Baranoff, J.-H. Yum, I. Jung, R. Vulcano, M. Grätzel and Md. K. Nazeeruddin, Chem.–Asian J., 2010, 5, 496 CrossRef CAS; (b) E. Baranoff, J.-H. Yum, M. Grätzel and Md. K. Nazeeruddin, J. Organomet. Chem., 2009, 694, 2661 CrossRef CAS.
  5. (a) V. Fernández-Moreira, F. L. Thorp-Greenwood and M. P. Coogan, Chem. Commun., 2010, 46, 186 RSC; (b) F. L. Thorp-Greenwood, R. G. Balasingham and M. P. Coogan, J. Organomet. Chem., 2012, 714, 12 CrossRef CAS; (c) K. K.-W. Lo, S. P.-Y. Li and K. Y. Zhang, New J. Chem., 2011, 35, 265 RSC; (d) P.-K. Lee, W. H.-T. Law, H.-W. Liu and K. K.-W. Lo, Inorg. Chem., 2011, 50, 8570 CrossRef CAS; (e) F. L. Thorp-Greenwood, Organometallics, 2012, 31, 5686 CrossRef CAS.
  6. (a) B. D. Muegge and M. M. Richter, Anal. Chem., 2004, 76, 73 CrossRef CAS; (b) J. I. Kim, I.-S. Shin, H. Kim and J.-K. Lee, J. Am. Chem. Soc., 2005, 127, 1614 CrossRef CAS; (c) M. M. Richter, Chem. Rev., 2004, 104, 3003 CrossRef CAS; (d) W. Miao, Chem. Rev., 2008, 108, 2506 CrossRef CAS; (e) F. Kessler, R. D. Costa, D. Di Censo, R. Scopelliti, E. Ortí, H. J. Bolink, S. Meier, W. Sarfert, M. Grätzel, Md. K. Nazeeruddin and E. Baranoff, Dalton Trans., 2012, 41, 180 RSC.
  7. (a) F.-M. Hwang, H.-Y. Chen, P.-S. Chen, C.-S. Liu, Y. Chi, C.-F. Shu, F.-I. Wu, P.-T. Chou, S.-M. Peng and G.-H. Lee, Inorg. Chem., 2005, 44, 1344 CrossRef CAS; (b) C. Dragonetti, L. Falciola, P. Mussini, S. Righetto, D. Roberto, R. Ugo, A. Valore, F. De Angelis, S. Fantacci, A. Sgamellotti, M. Ramon and M. Muccini, Inorg. Chem., 2007, 46, 8533 CrossRef CAS; (c) P.-K. Lee, H.-W. Liu, S.-M. Yiu, M.-W. Louie and K. K.-W. Lo, Dalton Trans., 2011, 40, 2180 RSC; (d) E. E. Langdon-Jones, J. D. Routledge, A. J. Hallett, D. Crole, B. D. Ward, J. A. Platts and S. J. A. Pope, Inorg. Chem., 2013, 52, 448 CrossRef CAS.
  8. Selected examples (a) W.-H. Zhang, X.-H. Zhang, A. L. Tan, M. A. Yong, D. J. Young and T. S. A. Hor, Organometallics, 2012, 31, 553 CrossRef CAS; (b) K. K.-W. Lo, J. S.-W. Chan, L.-H. Lui and C.-K. Chung, Organometallics, 2004, 23, 3108 CrossRef CAS; (c) R. Tao, J. Qiao, G. Zhang, L. Duan, L. Wang and Y. Qiu, J. Phys. Chem. C, 2012, 116, 11658 CrossRef CAS; (d) W.-J. Xu, S.-J. Liu, T.-C. Ma, Q. Zhao, A. Pertegás, D. Tordera, H. J. Bolink, S.-H. Ye, X.-M. Liu, S. Sun and W. Huang, J. Mater. Chem., 2011, 21, 13999 RSC; (e) C. Dragonetti, A. Valore, A. Colombo, S. Righetto and V. Trifiletti, Inorg. Chim. Acta, 2012, 388, 163 CrossRef CAS; (f) B. Tong, J.-Y. Qiang, Y.-Q. Xu, Q. Mei, T. Duan, Q. Chen and Q.-F. Zhang, Inorg. Chem. Commun., 2011, 14, 1937 CrossRef CAS; (g) P.-K. Lee, W. H.-T. Law, H.-W. Liu and K. K.-W. Lo, Inorg. Chem., 2011, 50, 8570 CrossRef CAS; (h) Q. Zhao, F. Li, S. Liu, M. Yu, Z. Liu, T. Yi and C. Huang, Inorg. Chem., 2008, 47, 9256 CrossRef CAS; (i) K. K.-W. Lo, C.-K. Chung, T. K.-M. Lee, L.-H. Lui, K. H.-K. Tsang and N. Zhu, Inorg. Chem., 2003, 42, 6886 CrossRef CAS; (j) M.-J. Li, P. Jiao, W. He, C. Yi, C.-W. Li, X. Chen, G.-N. Chen and M. Yang, Eur. J. Inorg. Chem., 2011, 197 CrossRef CAS; (k) M. Schmittel, Q. Shu and M. E. Cinar, Dalton Trans., 2012, 41, 6064 RSC; (l) M.-J. Li, P. Jiao, M. Lin, W. He, G.-N. Chen and X. Chen, Analyst, 2011, 136, 205 RSC.
  9. S. Zanarini, M. Felici, G. Valenti, M. Marcaccio, L. Prodi, S. Bonacchi, P. Contreras-Carballada, R. M. Williams, M. C. Feiters, R. J. M. Noltr, L. De Cola and F. Paolucci, Chem.–Eur. J., 2011, 17, 4640 CrossRef CAS.
  10. (a) S. P.-Y. Li, H.-W. Liu, K. Y. Zhang and K. K.-W. Lo, Chem.–Eur. J., 2010, 16, 8329 CrossRef CAS; (b) H. Yang, L. Li, L. Wan, Z. Zhou and S. Yang, Inorg. Chem. Commun., 2010, 13, 1387 CrossRef CAS.
  11. N. B. Sankaran, A. Z. Rys, R. Nassif, M. K. Nayak, K. Metera, B. Chen, H. S. Bazzi and H. F. Sleiman, Macromolecules, 2010, 43, 5530 CrossRef CAS.
  12. (a) W. Jiang, Y. Gao, Y. Sun, F. Ding, Y. Xu, Z. Bian, F. Li, J. Bian and C. Huang, Inorg. Chem., 2010, 49, 3252 CrossRef CAS; (b) C. Li, J. Lin, Y. Guo and S. Zhang, Chem. Commun., 2011, 47, 4442 RSC.
  13. N. D. McDaniel, F. J. Coughlin, L. L. Tinker and S. Bernhard, J. Am. Chem. Soc., 2008, 130, 210 CrossRef CAS.
  14. (a) A. Habibagahi, Y. Mébarki, Y. Sultan, G. A. Yap and R. J. Crutchley, ACS Appl. Mater. Interfaces, 2009, 1, 1785 CrossRef CAS; (b) K. K.-W. Lo, C.-K. Chung and N. Zhu, Chem.–Eur. J., 2003, 9, 475 CrossRef CAS; (c) K. Konishi, H. Yamaguchi and A. Harada, Chem. Lett., 2006, 720 CrossRef CAS.
  15. (a) J. P. Michael, Nat. Prod. Rep., 1997, 14, 605 RSC; (b) M. Balasubramanian and J. G. Keay, in Comprehensive Heterocyclic Chemistry II, ed. A. R. Katritzky, C. W. Rees and E. F. V. Scriven, Pergamon Press, Oxford, New York, 1996, vol. 5, p. 245 Search PubMed; (c) A. Burger, Medicinal Chemistry, Interscience, New York, NY, 2nd edn, 1960, p. 349 Search PubMed; (d) J. C. Krantz and C. J. Carr, The Pharmacological Principle of Medical Practice, Williams and Wilkins, Baltimore, MD, London, 6th edn, 1965, p. 826 Search PubMed.
  16. (a) R. D. Larsen, E. G. Corley, A. O. King, J. D. Carrol, P. Davis, T. R. Verhoeven, P. J. Reider, M. Labelle, J. Y. Gauthier, Y. B. Xiang and R. J. Zamboni, J. Org. Chem., 1996, 61, 3398 CrossRef CAS; (b) Y.-L. Chen, K.-C. Fang, J.-Y. Sheu, S.-L. Hsu and C.-C. Tzeng, J. Med. Chem., 2001, 44, 2374 CrossRef CAS; (c) G. Roma, M. D. Braccio, G. Grossi, F. Mattioli and M. Ghia, Eur. J. Med. Chem., 2000, 35, 1021 CrossRef CAS; (d) D. Dubé, M. Blouin, C. Brideau, C.-C. Chan, S. Desmarais, D. Ethier, J.-P. Falgueyret, R. W. Friesen, M. Girard, Y. Girard, J. Guay, D. Riendeau, P. Tagari and R. N. Young, Bioorg. Med. Chem. Lett., 1998, 8, 1255 CrossRef; (e) M. P. Maguire, K. R. Sheets, K. McVety, A. P. Spada and A. Zilberstein, J. Med. Chem., 1994, 37, 2129 CrossRef CAS; (f) S.-Y. Tanaka, M. Yasuda and A. Baba, J. Org. Chem., 2006, 71, 800 CrossRef CAS.
  17. P. W. Baures, S. A. Peterson and J. W. Kelly, Bioorg. Med. Chem., 1998, 6, 1389 CrossRef CAS.
  18. S. Vilar, L. Santana and E. Uriarte, J. Med. Chem., 2006, 49, 1118 CrossRef CAS.
  19. M. Nonoyama, Bull. Chem. Soc. Jpn., 1974, 47, 767 CrossRef CAS.
  20. K. Y. Zhang, S. P.-Y. Li, N. Zhu, I. W.-S. Or, M. S.-H. Cheung, Y.-W. Lam and K. K.-W. Lo, Inorg. Chem., 2010, 49, 2530 CrossRef CAS.
  21. H.-Y. Chen, C.-H. Yang, Y. Chi, Y.-M. Cheng, Y.-S. Yeh, P.-T. Chou, H.-Y. Hsieh, C.-S. Liu, S.-M. Peng and G.-H. Lee, Can. J. Chem., 2006, 84, 309 CrossRef CAS.
  22. (a) F. Neve, A. Crispini, S. Campagna and S. Serroni, Inorg. Chem., 1999, 38, 2250 CrossRef CAS; (b) S. Ladouceur, D. Fortin and E. Zysman-Colman, Inorg. Chem., 2010, 49, 5625 CrossRef CAS.
  23. (a) A. Mishra, P. K. Nayak, D. Ray, M. P. Patankar, K. L. Narasimhan and N. Periasamy, Tetrahedron Lett., 2006, 47, 4715 CrossRef CAS; (b) B. W. D. Andrade, S. Datta, S. R. Forrest, P. Djurovich, E. Polikarpov and M. E. Thompson, Org. Electron., 2005, 6, 11 CrossRef.
  24. J. E. Jones, R. L. Jenkins, R. S. Hicks, A. J. Hallett and S. J. A. Pope, Dalton Trans., 2012, 41, 10372 RSC.
  25. K. K.-K. Lo, C.-K. Chung and N. Zhu, Chem.–Eur. J., 2006, 12, 1500 CrossRef CAS.
  26. A referee is thanked for stimulating and guiding comments.
  27. N. G. Connelly and W. E. Geiger, Chem. Rev., 1996, 96, 877 CrossRef CAS.
  28. SHELXL-PC Package, Bruker Analytical X-ray Systems, Madison, WI, 1988 Search PubMed.
  29. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery Jr., T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian 03, Revision E.01, Gaussian, Inc., Wallingford, CT, 2004 Search PubMed.
  30. (a) P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 270 CrossRef CAS; (b) P. J. Hay and W. R. Wadt, J. Chem. Phys., 1985, 82, 299 CrossRef CAS.
  31. (a) M. M. Francl, W. J. Pietro, W. J. Hehre, J. S. Binkley, D. J. DeFrees, J. A. Pople and M. S. Gordon, J. Chem. Phys., 1982, 77, 3654 CrossRef CAS; (b) P. C. Hariharan and J. A. Pople, Theor. Chem. Acc., 1973, 28, 213 CrossRef CAS.
  32. R. Ditchfield, W. J. Hehre and J. A. Pople, J. Chem. Phys., 1971, 54, 724 CrossRef CAS.
  33. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, T. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, GAUSSIAN 09 (Revision C.01), Gaussian, Inc., Wallingford, CT, 2010 Search PubMed.
  34. J. Tomasi, B. Mennucci and R. Cammi, Chem. Rev., 2005, 105, 2999 CrossRef CAS and references therein.
  35. L. Monti and G. Franchi, Gazz. Chim. Ital., 1951, 81, 544 CAS.
  36. C. M. Dupureur and J. K. Barton, Inorg. Chem., 1997, 36, 33 CrossRef CAS.
  37. (a) V. W. W. Yam, K. K. W. Lo, K. K. Cheung and R. Y. C. Kong, J. Chem. Soc., Chem. Commun., 1995, 1191 RSC; (b) M. Shilpa, P. Nagababu, Y. P. Kumar, J. N. L. Latha, M. R. Reddy, K. S. Karthikeyan, N. Md. Gabra and S. Satyanarayana, J. Fluoresc., 2011, 21, 1155 CrossRef CAS.
  38. (a) R. López, B. Loeb, T. Boussie and T. J. Meyer, Tetrahedron Lett., 1996, 37, 5437 CrossRef; (b) A. Ambroise and B. G. Maiya, Inorg. Chem., 2000, 39, 4256 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Data collection parameters for the crystallographic studies, electrochemical data for 3a–h, pictorial representations of the calculated frontier orbitals for 3a–h and 4a–h and Cartesian coordinates obtained from DFT calculations. CCDC 907280 and 907282. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c3dt51098k

This journal is © The Royal Society of Chemistry 2013
Click here to see how this site uses Cookies. View our privacy policy here.