DOI:
10.1039/C6RA16319J
(Paper)
RSC Adv., 2016,
6, 87389-87399
Synthesis, structures and photophysical properties of luminescent cyanoruthenate(II) complexes with hydroxylated bipyridine and phenanthroline ligands†
Received
24th June 2016
, Accepted 20th August 2016
First published on 5th September 2016
Abstract
Treatment of (PPh4)2[RuII(PPh3)2(CN)4] (1) with bpyOH and phenOH in DMF afforded two mononuclear compounds fac-(PPh4)[RuII(bpyOH)(PPh3)(CN)3] (2) and fac-(PPh4)[RuII(phenOH)(PPh3)(CN)3] (3), respectively (bpyOH = 6-hydroxy-2,2′-bipyridine; phenOH = 2-hydroxy-1,10-phenanthroline). These complexes have been characterized by various spectroscopic techniques. The structures of 1 and 2 have also been determined by X-ray crystallography. Their photophysical and electrochemical properties have been investigated. Compound 3 displays intense emission with much higher quantum efficiency (Φem = 19.4%) in solution, compared with other related ruthenate(II) diimine complexes. In addition, their solvatochromism, pH effects, and ion perturbation have also been investigated. The different photoluminescent behaviors between these complexes and the previously reported complex [RuII(phen)(PPh3)(CN)3]− suggest that the introduction of the hydroxyl group into the diimine ligand would significantly affect their emission properties. Through spectrophotometric titrations, the ground state pKa and excited state pKa* values, as well as the dynamic pH response ranges of these complexes have been determined. Their photophysical response towards various metal ions have also been studied.
Introduction
The utilization of luminescent transition metal complexes in the design of chemosensors and bioprobes has attracted considerable interest owing to their advantageous photophysical properties such as long emission lifetime, high photostability and large Stokes shifts.1,2 Luminescent transition metal complexes have been extensively studied as potential sensors/probes for O2 and CO2,3,4 pH,5 chloride,6 temperature,7 various anions and cations.8 In particular, Ru(II) complexes bearing various α,α′-diimine ligands have received much attention owing to their outstanding chemical and photochemical stability as well as their rich electrochemical and photophysical properties.9 [RuII(bpy)(CN)4]2−, one of the best luminescent ruthenium(II) compounds was firstly reported in 1986.10 Its photophysical properties have been thoroughly investigated in aqueous media and in acidic solutions.11 This complex shows highly environmentally sensitive metal-to-ligand charge-transfer (MLCT) absorption and emission due to the interaction between the cyanide ligands and the surrounding microenvironment including solvent molecules, protons and metal ions, this property has been utilized in the development of luminescent chemosensors, probes, and stimuli-responsive devices.12 However, the relatively low luminescent quantum yields of the tetracyanoruthenate(II) complexes limit their further development in various applications.13 Recently, we have prepared a series of derivatives of cyanoruthenate(II) diimine complexes [RuII(N^N)(PR3)(CN)3]−, with improved luminescent properties through the substitution of a cyanide ligand by a better π-accepting phosphine ligand.14 Similar to [RuII(N^N)(CN)4]2−, remarkable solvatochromic effect was observed in the absorption and emission spectra of [RuII(N^N)(PPh3)(CN)3]−.14
Herein, we attempt to modify the photophysical properties of the [RuII(N^N)(PPh3)(CN)3]− chromophore by introducing a hydroxyl group to the α-position of the diimine ligand. We anticipate that the electronic and H-bonding effects of the hydroxyl group would have a significant effect on the properties of these complexes. The keto–enol tautomerism of hydroxyl-substituted diimine ligands could also be utilized in the development of pH/metal ion sensors.15 In this context, two new tricyanoruthenate(II) complexes with hydroxyl-substituted diimine ligands have been prepared. The photophysical and electrochemical properties of these complexes have been investigated. In addition, the effects of the microenvironment on the photophysical properties such as solvatochromism, pH effects as well as cationic effect on the photophysical properties of these complexes have also been studied.
Results and discussion
Synthesis and characterization
Reaction of [RuII(PPh3)3Cl2] with excess KCN in MeOH afforded the complex trans-[RuII(PPh3)2(CN)4]2−, which was isolated as PPh4 salt (1) with high yield. The IR spectrum of 1 shows the characteristic ν(C
N) stretching bands at 2059 and 2100 cm−1. The electrospray ionization mass spectrum (ESI/MS) of 1 in MeOH (−ve mode) shows two predominant peaks at m/z 365.0 and 1069.0, which are assigned to [M]2− and [M + PPh4]−, respectively. Treatment of 1 with bpyOH or phenOH in DMF afforded fac-(PPh4)[RuII(bpyOH)(CN)3(PPh3)] (2) or fac-(PPh4)[RuII(phenOH)(CN)3(PPh3)] (3), respectively (Scheme 1). These complexes have been characterized by 1H NMR, 31P{1H} NMR, IR spectroscopy, mass spectrometry and elemental analysis. The crystal structures of 1 and 2 have also been determined by X-ray crystallography.
 |
| Scheme 1 Synthetic route to 2 and 3. | |
In the IR spectra, 2 and 3 are characterized by strong ν(C
N) stretches at 2093, 2057, 2045 cm−1 and 2089, 2082, 2075 cm−1, respectively. The ESI-mass spectra of 2 and 3 in methanol (−ve mode) show a predominant parent peak [M]− at m/z 614 and 638, respectively. In addition, 2 and 3 show 1H NMR and 31P{1H} NMR signals with chemical shifts and integral ratios consistent with their predicted structures. The NMR signals for phenyl protons in PPh3 and PPh4+ and aromatic protons in diimine ligands are comparable with those in the other related complexes.16 The singlet at δ 10.15 and 10.69 ppm are observed for 2 and 3, respectively which are ascribed to the exchangeable OH proton in the diimine ligands. This is further confirmed by the observation of disappearance of 1H NMR signal with chemical shift > 10 ppm when few drops of D2O were added. In their 31P NMR spectra, two singlet at δ 41.02 and 22.32 ppm in 2 and δ 41.84 and 22.32 ppm in 3 are observed, which are ascribed to the phosphorus nuclei in PPh3 and PPh4+, respectively.
X-ray crystal structures determination
The crystal structures of 1 and 2 have been determined by X-ray crystallography. The crystal data, data collection and structure refinement details are summarized in Table S1† and selected bond parameters about these two compounds are listed in Table S2.† As shown in Fig. 1, the ruthenium center of 1 adopts a distorted octahedral geometry and is coordinated by four cyanides in equatorial positions and two triphenylphosphine ligands in the axial positions. The average Ru–C and Ru–P bonds are 2.047(2) and 2.334(6) Å, respectively.
 |
| Fig. 1 The ORTEP drawing of the anionic structure of 1. | |
The ORTEP drawing of 2 is shown in Fig. 2a. The asymmetric unit of 2 contains two octahedral [RuII(bpyOH)(PPh3)(CN)3]− anions whose charge are balanced by two PPh4+ cations. The ruthenium center adopts a distorted octahedral geometry and is coordinated by two N atoms from a bpyOH ligand and two C atoms from two cyanides in the equatorial plane. The axial positions are occupied by a triphenylphosphine and a cyanide ligand. The Ru–P bond distance is 2.365(4) Å, which is comparable with those reported for related compounds.16 The bond length of Ru–CN trans to phosphine is 2.054(14) Å, which is longer than the two cis-Ru–CN bonds [2.001(15), 1.979(15) Å] in the equatorial positions. This is attributable to the stronger trans influence and π-accepting ability of the phosphine ligands compared with the diimine ligands. The Ru–N bond lengths are in the range of 2.116(12) to 2.139(12) Å, which are consistent with those observed in related Ru(II) complexes with bipyridine ligands.17 The hydroxyl group in bpyOH ligand shows a high degree of disorder, which is probably due to the presence of a pseudo-symmetry plane passing through the principle axis. The calculated structure of 2 is shown in Fig. 2b. The calculated structure parameters of 2 are compared with the X-ray crystallography data in Table S2.† The results obtained from the DFT B3LYP method with LANL2DZ/6-31G (d) basis set correspond well with the experimental data, as reflected from the deviation of bond lengths between the calculated and experimental data of ≤0.05 Å.
 |
| Fig. 2 (a) The ORTEP drawing and (b) optimized structure of anionic structure of 2. | |
Electrochemical properties
The electrochemical properties of these cyanoruthenate(II) complexes have been studied by cyclic voltammetry (CV) in deoxygenated MeCN solution (0.1 M nBuNPF6). The CV of 1 shows a reversible redox couple centered at E1/2 = 0.08 V (ΔEp = 84 mV, ipc/ipa = 0.9) (versus Cp2Fe+/0, Cp = cyclopentadienyl), which is tentatively assigned to the metal centered Ru3+/2+ couple. In addition, 2 and 3 show an reversible metal-centered Ru3+/2+ oxidation couple at E1/2 = 0.45 (ΔEp = 80 mV, ipc/ipa = 0.9) and 0.41 V (ΔEp = 76 mV, ipc/ipa = 0.9) and an irreversible ligand-based reduction couple at Epc = −2.50 and −2.54 V.
UV/vis absorption and emission properties
The UV/vis absorption properties of all complexes have been investigated (Fig. 3, Table 1). In CH3CN solution, 2 and 3 exhibit intense absorptions in the range of 250–350 nm with molar extinction coefficients on the order 104 dm3 mol−1 cm−1, which are ascribed to ligand-centered (LC) ππ* absorptions of phosphine and diimine ligands. In the visible region, both complexes show moderately intense absorptions with molar absorptivities on the order of 103 dm3 mol−1 cm−1, which is not observed in 1. According to the previous spectroscopic studies of related complexes,12,14 the lowest-energy absorptions of these complexes are tentatively assigned to 1MLCT [dπ(Ru) → π*(N^N)] transition. In order to gain further insights into the electronic structure of 2, DFT calculations have been performed. As shown in the contour plots of the calculated frontier molecular orbitals (MOs) of 2, the HOMO of 2 shows a mixing of dπ(Ru) and π(CN) orbitals (Fig. 4a), while the LUMO is mainly composed of π*(bpyOH) orbital (Fig. 4b). This result is consistent with the 1MLCT assignment for the lowest-energy transition and is suggestive of a mixing of LLCT character in the lowest-energy transition for 2 and 3.
 |
| Fig. 3 Overlaid UV/Vis spectra for 1–3 in CH3CN solution. | |
Table 1 UV/Vis absorption data of 2 and 3 in various solvent media
Complex |
Medium |
Absorption λabs/nm (ε/M−1 cm−1) |
2 |
MeOH |
269 (11 890), 276 (10 970), 332 (5770), 406 (1850) |
EtOH |
268 (5800), 276 (5470), 307 (5800), 414 (1010) |
CH2Cl2 |
269 (11 670), 276 (10 190), 314 (10 720), 441 (1670) |
DMSO |
316 (46 260), 447 (7330) |
CH3CN |
260 sh (9790), 268 (9740), 276 (9540), 307 (12 680), 330 sh (5820), 433 (2200) |
DMF |
315 (54 260), 446 (10 300) |
3 |
MeOH |
268 (20 080), 276 (17 620), 415 (3030) |
EtOH |
268 (14 780), 275 (12 950), 414 (2010) |
CH2Cl2 |
269 (35 350), 447 (4570) |
DMSO |
270 (35 000), 381 sh (3880), 446 (4560) |
CH3CN |
268 (29 020), 274 sh (26 170), 295 sh (10 290), 383 (3000), 433 (3580) |
DMF |
268 (28 740), 381 (3070), 446 (3640) |
 |
| Fig. 4 (a) Highest occupied and (b) lowest unoccupied molecular orbitals from energy-minimized calculated (Gaussian) anionic structure of 2. | |
Upon optical excitation, 2 and 3 exhibit solvent-dependent greenish yellow to orange emission in various degassed solutions at 298 K with sub-microsecond emission lifetimes (Fig. 5a and Table 2). According to previous spectroscopic studies of related ruthenium(II) diimine complexes,11,13 these emissions are ascribed to the 3MLCT [dπ(Ru) → π*(N^N)] excited state origin. The quantum yield of 2 increases from acetone (1.5%), CH2Cl2 (3.2%) to MeOH (5.8%), while 3 exhibits a different trend in these solvents [acetone (6.0%), MeOH (7.7%), and CH2Cl2 (19.4%)]. It is notable that 3 displays the highest emission quantum yield in CH2Cl2 solution; moreover, the quantum yield of 3 is much higher than those reported for other related cyanoruthenate(II) diimine complexes.14 The much higher quantum efficiency of 3 is probably due to two important factors: (1) a strong π-accepting ligand PPh3 replacing the CN ligand that will stabilize the dπ(Ru) orbital; (2) a strong electron-donating −OH group on the diimine ligand that will significantly raise the π* level of the diimine ligand, resulting in the remarkable increase in the MLCT emission energy as compared with related compounds.14 Both of these factors will subsequently raise energy of the 3MLCT emission state, leading to the higher quantum efficiency through effective impediment of the thermal deactivation pathway.
 |
| Fig. 5 (a) Overlaid emission spectra of 2 and 3 in CH2Cl2 solution at 298 K and (b) in EtOH/MeOH (4 : 1 v/v) glassy medium at 77 K. | |
Table 2 Emission data of 2 and 3 in various solvent media
|
Medium (T/K) |
Emission λem/nm (τo/μs)a |
ϕem × 103b |
Excitation at 400 nm. Luminescence quantum yield on excitation at 436 nm. In EtOH/MeOH (4 : 1, v/v). |
2 |
CH2Cl2 (298) |
610 (0.50) |
31.5 |
MeOH (298) |
578 (0.58) |
58.2 |
Acetone (298) |
629 (0.23) |
15.4 |
Glassc (77) |
535, 511 (11.17) |
|
3 |
CH2Cl2 (298) |
596 (5.72) |
194.2 |
MeOH (298) |
569 (3.85) |
77.2 |
Acetone (298) |
617 (1.63) |
60.5 |
Glassc (77) |
546, 521 (56.90) |
|
In EtOH/MeOH (4
:
1 v/v) glassy medium at 77 K (Fig. 5b, Table 2), the emissions of these complexes are hypsochromically shifted with respect to those recorded at room temperature. 2 and 3 show broad emission with peak maximum at ca. 511 nm (τ = 11.17 μs) and 521 nm (τ = 56.90 μs), respectively. These luminescence profiles feature the characteristic vibrational progressions with spacing ∼870 cm−1, which are similar to those reported in other [Ru(bpy)3]2+-type emitter.18 In view of their structured emission and the longer excited state lifetime, these emissions are derived from the 3MLCT [dπ(Ru) → π*(N^N)] excited state origin probably with some mixing of 3LC character. This is due to the rigidochromic effect, which is commonly observed in MLCT emitters.19
Solvent effect
Similar to the reported cyanoruthenate(II) diimine complexes,10,13 2 and 3 also display strong solvatochromism, owing to their solvent-sensitive MLCT transition and the interactions between the solvent molecules and −CN/−OH moieties of the complexes (Table 1, Fig. 6).12 In the absorption spectra of 2 in different solvents, the lowest-energy MLCT absorption shifts from acetone (456 nm) to MeOH (406 nm) with the energy difference of ∼2700 cm−1 (Fig. 6a). It is noteworthy that the solvatochromic shift of MLCT transitions of 2 in acetone, DMF, CH2Cl2 and MeOH is in line with the Gutmann acceptor number of solvents. The absorption peaks shift to the lower energy region with decreasing acceptor numbers of the solvent molecules. However, deviation of the absorptions in MeCN and DMSO from the correlation is clearly observed (Fig. 7a). This is attributable to the presence of hydrogen-bonding between these solvent molecules and −OH group in bpyOH. In the bpyOH ligand, the sp2 hybridized hydroxyl group has a lone pair electron and a δ+ H atom, which could act as H-bonding acceptor (A) or H-bonding donor (D) with respect to the solvent molecules. These two contrasting interactions could result in two opposite effects on the π* orbit of the diimine ligand: interaction of the lone pair with solvent will stabilize the π* orbits of the ligand; whereas H-bonding interaction via O–H⋯A(solvent) will raise the π* level. In MeCN and DMSO, the hydroxyl group effectively acts as a D that will raise the energy of π*(bpyOH) orbitals, which will lead to a blue-shift in absorption energies from those expected from their acceptor number. However, the hydroxyl moiety simultaneously acts as D and A in protic MeOH/EtOH, hence, these contrasting interactions may cancel out each other. As a result, the solvatochromic effects in MeOH/EtOH are in agreement with their Gutmann acceptor numbers.
 |
| Fig. 6 Overlaid UV/vis spectra of 2 in various solvent media. | |
 |
| Fig. 7 (a) The plot of the lowest MLCT absorption energy (Eabs) and (b) the plot of emission energy (Eem) of 2 versus the acceptor number of the solvents. | |
Similar to the trend observed in the absorption spectra, the emission of 2 is also solvent-dependent, the-peak shifts from acetone (628 nm) to MeOH (573 nm) (Table 2 and Fig. 8) with the energy difference of about 1520 cm−1, which is much less pronounced than that in the absorption spectra. This may be due to the reduced electron density in the cyanide ligands in the MLCT excited state, resulting in lower basicity of the ligand. On the other hand, the increased basicity on the excited state, compared to the ground state, is typical for complexes with a pendant basic site. The MLCT state of 2 obviously involves bpyOH ligand rather than PPh3 or CN− ligands, therefore the pendant −OH site of bpyOH has an increased affinity for protons. The net effect is that the hydroxyl group in bpyOH acts as a better A than D. This assumption is corroborated by the emission energies of 2 in MeCN and DMSO, which are essentially in line with their respective acceptor numbers, since the increased basicity of hydroxyl group will reduce its interaction with these H-bonding acceptor solvents. Moreover, the H-bonding effect between hydroxyl group and the protic solvents MeOH/EtOH in the ground state is also disequilibrated. The deviation of the 3MLCT emission of 2 in MeOH/EtOH also arises from the hydroxyl group acting as a better A than D in its excited state (Fig. 7b).
 |
| Fig. 8 Overlaid emission spectra of 2 in various solvent media. | |
Similar to the trend observed in the absorption and emission spectra of 2 in various solvent media, 3 shows the lowest-energy absorptions from acetone (445 nm) to MeOH (408 nm) with the energy difference ∼2030 cm−1 (Fig. S5†) and the emissions from acetone (616 nm) to MeOH (568 nm) with the energy difference ∼1370 cm−1 (Fig. S6†). Compared with the energy difference of 2 and 3 in various solvent media, both absorption and emission properties of 3 are found to be less sensitive to change of solvent media. The solvatochromism of [RuII(phen)(PPh3)(CN)3]− has been investigated in different solvents14 and such a variation trend is not observed in this system, which suggest that the introduction of hydroxyl group into the diimine ligand would significantly affect their emission properties.
pH effect
Previous studies show that the photophysical properties of cyanoruthenate complexes are sensitive to the pH of the microenvironment.10,12 Upon protonation of cyanide ligands, the MLCT [dπ(Ru) → π*(N^N)] energy would increase due to the better stabilization of dπ(Ru) orbitals in the presence of better π-accepting CNH ligands.20 As a result, blue-shift of the lowest-energy absorption and emission would be expected in acidic medium. In contrast, protonation at the −OH group of the diimine ligand will produce two opposite effects on the 1MLCT energy, as the effect of −OH group protonation on the complexes is two-fold: (1) withdrawal of the electron density from the diimine ligand will reduces its π-donating character and will stabilize the dπ(Ru) orbit, and (2) substantial stabilization of the diimine ligand LUMO. The net effects of these two processes produces an overall decrease in the MLCT [dπ(Ru) → π*(N^N)] energy. Moreover, diimine–OH moiety could be deprotonated under basic media. Though complicated response towards the environmental pH values may be expected due to the presence of two distinct and spatially separate sites, 2 and 3 exhibit simple response towards the external pH value change with no obvious oscillating behaviour as reflected in the absorption and emission spectra of 2 and 3 in the pH range of 2–13.
Due to solubility limitation of 2 and 3 in pure water, MeCN/H2O (1
:
4, v/v) solution was employed as the solvent for pH titration studies. The absorption spectral change of 2 with pH of the media is shown in Fig. 9a. The region pertaining to 1MLCT and 1LC transitions in 2 remain nearly constant over the pH range of 2–8. However, on further increase in pH from 8–12, there are gradual spectral changes in absorption, consistent with a deprotonation step. In this pH range, the 1MLCT maximum is red-shifted with a stabilization of the energy level by ca. 510 cm−1. The absorption band at ca. 304 nm decreases and the band at 356 nm increases with an isosbestic point at 339 nm. The evolved absorption band at 356 nm is tentatively assigned to a ligand-centered process involving the deprotonated bpyOH ligand, which occurs at relatively low energy possibly due to the presence of a more conjugated π-system. The presence of a well-defined isosbestic point suggests the existence of two detectable species in equilibrium, i.e. the deprotonated and protonated forms. The fact that the change in the 1LC absorption maximum arise solely from deprotonation at −OH group allows determination of an apparent pKa value. Plot of absorbance at 304 nm versus pH, gives the expected sigmoidal shape. According to eqn (1), the pKa value of 2 is determined to be 10.5 ± 0.2.
|
 | (1) |
 |
| Fig. 9 (a) Absorption and (b) emission spectral change of 2 (1.0 × 10−4 mol L−1) in pH 2–13 in MeCN/H2O (1 : 4, v:v). Inset shows the variation of (a) absorption at λabs = 304 nm and (b) emission intensity at λem = 565 nm. | |
The emission spectral change of 2 in MeCN/H2O solution as a function of pH has also been studied (Fig. 9b). At pH < 8, the emission of 2 does not exhibit appreciable pH dependence, whereas the luminescence intensity is progressively quenched as the pH is increased from 8 to 12. The plot of emission intensity vs. pH is sigmoidal with an inflexion point at pH 10.2, which corresponds to 50% of quenching (pHi). The pHi value is essentially identical with the measured pKa value by UV/vis absorption method. The monotonic trend of the sigmoidal curve also offers a clear proof of the existence of one excited-state protonation/deprotonation process of hydroxyl group in 2. In principle, the excited state pKa* needs to be corrected for the differences in the excited-state lifetime of protonated (τ) and deprotonated form (τ′) according to eqn (2).
|
pKa* = pHi + log τ/τ′
| (2) |
As the deprotonated species is essentially non-luminescent, the excited state pKa* for 2 could be determined by an empirical method based on Forster's cycle with the eqn (3), where νB and νBH are energies in cm−1 of the lowest-energy 1MLCT band maxima in deprotonated and protonated states.21,22
|
pKa* = pKa + (0.625/T)(νBH − νB)
| (3) |
For 2, we have pKa = 10.5 and νBH − νB = 510 cm−1, which gives pKa* ≈ 11.6. These values corroborate the obvious increased in basicity of the excited state as compared to that of the ground state. The UV/vis spectra of 2 shows that the 1LC transition remains unchanged at pH 2–8, which indicates that there is negligible interactions of −OH and −CN groups with H+ in this pH range. At higher pH values, deprotonation of −OH group occurs and the negative charge is delocalized onto the proximal pyridyl N atom. In this form, the ligand undergoes virtually a tautomerization to give a deprotonated amide ligand that is a better π-donor and will weaken the ligand-field strength around the metal center and destabilize the filled dπ(Ru) orbital. The luminescent intensity decreases significantly, possibly due to rapid radiationless decay resulting from thermal equilibrium between MLCT and d–d excited states. Similar emission quenching upon deprotonation of the appended phenolic OH groups was reported for related systems.23
Fig. S7a and S7b† show the ground state absorption and emission spectra for 3, respectively, at various pH values, and the trend is similar to 2. However, spectrophotometric titrations show that the pKa of 3 is lower than that of 2, the UV/vis spectra remain nearly constant over the pH range 2–7, whereas the deprotonation step is clearly observed over the pH range of 7–11. When the pH is increased from 7 to 11, the absorption band at 268 nm decreases and the band at 290 nm increases with an isosbestic point at 282 nm. Plot of pH as a function of the absorption at 290 nm gives a sigmoidal curve, as in the case of 2. The pKa value of 3 was determined to be 9.2 ± 0.2 according to eqn (1). The uncorrected emission spectrum of 3 has a maximum at 560 nm, which does not show an appreciable dependence on the pH values at pH < 7 (Fig. 10b). On increasing the pH from 7 to 11, the luminescence intensity is progressively quenched, but the emission wavelength remains change. The pKa* of the excited state of 3 is also determined by the approximate method according to eqn (3) with a value of 11.7 ± 0.2. In order to confirm that the pH effect of compounds 2 and 3 arise from the protonation/deprotonation equilibrium of the pendent hydroxyl group on the diimine ligands, the pH effect of [RuII(phen)(PPh3)(CN)3]− towards its absorption and emission spectra has also been carried out. As shown in Fig. S8 and S9,† both its absorption and emission spectra remain nearly unchanged in a wide pH range from 3–12, indicating that the pH effects do not originate from the cyano ligands.
 |
| Fig. 10 (a) The photograph on the colour changes taken under daylight (above) and the luminescence colour changes taken under UV illumination (λex = 365 nm) on 2 (5.0 × 10−5 mol L−1) (below) upon addition of 10 equiv. of various cations; (b) The absorption and (c) emission spectra of 2 upon addition of 10 equiv. of different metal ions in MeOH. | |
Effects of cations
The absorption and emission spectra of complexes towards different metal ions have been recorded and shown in Fig. 10. Since the behaviours of 2 and 3 towards the examined ions are quite similar, discussion will be mainly focused on 2. The UV/vis spectra of 2 display obvious changes upon addition of Fe3+ and Cu2+, while there no significant changes are observed for Mg2+, Ca2+, Na+, K+, Ni2+, Mn2+, and Co2+. Upon addition of Fe3+, the absorbance of the system increases remarkably and the absorption position also change significantly with the appearance of a strong absorption band at λabs = 367 nm. Upon addition of Cu2+, the absorbance of solution increases remarkably in the UV region. However, there is no significant change in the visible region. The significant change in higher-energy regions may indicate that the stronger interactions of −OH group with Cu2+ and Fe3+ ions.
Dramatic changes of the absorption and emission spectra in 2 in MeOH solution are also observed when ZnCl2 is added. The MLCT absorption of 2 moves from 405 nm to 371 nm in the presence of two mole equivalent of Zn2+. This constitutes a considerable blue-shift of about 2260 cm−1 in the MLCT absorption maximum upon complex formation. However, the LC (ππ*) transition observed at 303 nm undergoes a much less pronounced shift to 294 nm. These spectral changes may result from replacement of solvent molecules around the nitrogen atoms of cyanide ligands upon coordination to Zn2+, while the interaction between the −OH group and Zn2+ is negligible possibly due to the significant steric repulsion. The emission intensity and wavelength of 2 remain nearly unchanged upon addition of 10 mole equivalents of cations such as Na+ and K+, while there is a blue-shift of less than 5 nm with Ca2+ and Mg2+; The more significant blue-shift by addition of Zn2+ is most possibly due to that Zn2+ is a more stronger Lewis acid with stronger RuCN–Zn interactions. On the other hand, the emission intensity is nearly fully quenched by cations such as Fe3+, Cu2+, Co2+, Mn2+ and Ni2+, possibly due to photo-induced electron transfer (PET) effect from the unpaired electrons in these ions.8,24
Titration experiment
In order to further investigate the interaction of Zn2+ with 2, titration of 2 with Zn2+ has been carried out by UV/vis absorption and emission methods in MeOH (Fig. 11). Upon addition of Zn2+, the absorption band at ca. 405 nm decreases and the band at 371 nm increases. An isosbestic point is observed at relatively low mole ratios of Zn2+/2 from 0 to 0.5. Further increase of [Zn2+] results in an additional slight blue-shift in MLCT transition without any isosbestic points. The donor–acceptor interaction of Zn2+ with cyano ligands leads to an increase in the π-acceptor ability of the cyano ligands, hence, the MLCT band assigned as spin-allowed t2g6 → t2g5π* transition is shifted to higher energy.25 The emission of 2 with its maximum at λem = 578 nm is readily quenched by Zn2+, with concomitant increase in luminescence increase at λem = 522 nm. The emission profiles also exhibit a temporary isoemissive point at 544 nm with the mole ratio of Zn2+/2 up to 0.5. The emission intensity of λem = 522 nm is further enhanced by increasing mole ratio of Zn2+/2 from 0.5 to 1.0, and then the emission intensity remains nearly unchanged even when the mole ratio of Zn2+/2 is increased to 1.5. Both the UV/vis and emission spectra exhibit a similar response towards Zn2+. These spectral changes upon titration with Zn2+ can be attributed to the successive formation two new species in this process. An initial 1
:
2 stoichiometric adduct is formed at the mole ratio of Zn2+/2 < 0.5 which is then progressively converted into a 1
:
1 adduct. In addition, the Job's plot, which is the plot of absorbance at 384 nm, also shows a 1
:
1 stoichiometry between 2 and Zn2+ (Fig. 11a inset).
 |
| Fig. 11 (a) UV/vis absorption and (b) emission spectral changes on titration of Zn2+ into a methanol solution of 2. Inset shows the Job's plot corresponds to a 1 : 1 receptor–metal complex formation and the absorbance at 384 nm was plotted as a function of the molar ratio [Zn2+]/{[Zn2+] + [2]}. | |
Under similar condition, addition of Cd2+ to a methanol solution of 2 also results in absorption and emission spectral changes. Upon addition of only 0.2 equiv. of Cd2+, a significant blue-shift from 575 nm to 520 nm is observed. It is noteworthy that the emission intensity of 2 exhibits a pronounced enhancement upon addition of about 0.4 equiv. Cd2+, however, since a white precipitate is formed in the process, further investigation is not carried out.
The effects of meal ions on [RuII(phen)(PPh3)(CN)3]− has also been investigated in MeOH in order to compare with those of 2 and 3. As shown in Fig. S10,† the emission spectra of this complex reveal that its MLCT band exhibits little change towards monovalent cations such as Na+ and K+ and is slightly blue-shifted in the presence of the divalent Mg2+. However, the effect of Zn2+ is more pronounced. These variations are similar to those of 2 and 3 under similar conditions. However, titration experiment involving addition of Zn2+ to an methanolic solution of [RuII(phen)(PPh3)(CN)3]− shows different results (Fig. S11†). Upon the addition of Zn2+, the complex shows a new emission band at λem = 521 nm, while the initial emission band at λem = 582 nm gradually blue shifts and becomes a shoulder peak at λem = 552 nm (Fig. S11†). This result indicates that there are at least two species upon addition of Zn2+ in this system. The existence of a broadened emission band is the result of at least two emitting species that leads to an apparent blue shifting of the λem. Moreover, the intensity of the new emission band at λem = 521 nm exhibits a remarkable enhancement as compared with that at λem = 582 nm; while in 2, the intensity of these emission bands remains nearly equal.
Experimental
Materials and reagents
The ligands, 6-hydroxy-2,2′-bipyridine (bpyOH), 2-hydroxy-1,10-phenanthroline (phenOH)26 and the complex [RuII(PPh3)3Cl2]27 were synthesized according to literature procedures. Triphenylphosphine and RuCl3 were purchased from Strem Chemical Company and were used without further purification. All other reagents and solvents were of analytical grade and were used without further purification.
Physical measurements and instrumentation
1H NMR and 31P{1H} NMR spectra were recorded on a Bruker AV300 (300 MHz) FT-NMR spectrometer. Chemical shifts (δ) are reported relative to tetramethylsilane (Me4Si). Negative-ion ESI mass spectra were recorded on a PE-SCIEX API 150 EX single-quadruple mass spectrometer. Elemental analysis was performed on an Elementar Vario MICRO Cube elemental analyzer. IR spectra of the solid samples as KBr discs were obtained within the range 4000–400 cm−1 on an AVATAR 360 FTIR spectrometer.
Electronic absorption spectra were recorded on a Hewlett-Packard 8453 or a Hewlett-Packard 8452A diode-array spectrophotometer. Steady-state emission and excitation spectra were measured at RT and at 77 K on a Horiba Jobin Yvon Fluorolog-3-TCSPC spectrofluorometer. The solutions were rigorously degassed on a high-vacuum line in a two-compartment cell with not less than four successive freeze–pump–thaw cycles. The measurements at 77 K were carried out on dilute solutions of the samples in EtOH/MeOH (4
:
1 v/v) loaded in a quartz tube inside a quartz-walled Dewar flask that contained liquid nitrogen. The luminescence quantum yields were determined by using the optical-dilution method as described by Demas and Crosby28 with an aerated aqueous solution of [Ru(bpy)3Cl2](ϕem = 0.040 (ref. 29) with excitation at 436 nm) as the reference. Luminescence lifetimes were measured by using the time-correlated single-photon-counting (TCSPC) technique on a Fluorolog-3-TCSPC spectrofluorometer in a fast MCS mode with a Nano LED-375 LH excitation source, which had a peak excitation wavelength at 375 nm and a pulse width of less than 750 ps. The photon-counting data were analysed on Horiba Jobin Yvon Decay Analysis Software.
Cyclic voltammetric (CV) measurements were performed on a CH Instruments, Inc. model CHI 620 Electrochemical Analyser in MeCN solution with [nBu4N]PF6 (0.1 M) as a supporting electrolyte at room temperature. The reference electrode was a Ag/AgCl (0.1 M in MeCN) electrode and the working electrode was a glassy carbon electrode (CH Instruments, Inc.) with platinum wire as the counter electrode. The surface of the working electrode was polished with a 1 μm a-alumina slurry (Linde) and then with a 0.3 μm a-alumina slurry (Linde) on a microcloth (Buehler Co.). The ferrocenium/ferrocene+/0 couple (FeCp2) was used as an internal reference. Solutions for electrochemical studies were de-aerated with pre-purified argon gas prior to the measurements.
X-ray crystal structure determination
The crystal structures were determined on an Oxford Diffraction Gemini S Ultra X-ray single-crystal diffractometer by using graphite-monochromated CuKa radiation (λ = 1.5417 Å). The structures were solved by using direct methods with the SHELXS-97 program.30 The Ru atoms and many of the non-hydrogen atoms were located according to the direct methods. The positions of the other non-hydrogen atoms were located after refinement by full-matrix least-squares by using the SHELXL-97 program.31 In the final stage of the least-squares refinement, all non-hydrogen atoms were refined anisotropically. H atoms were generated by using the SHELXL-97 program.31 The positions of H atoms were calculated based on the riding model with thermal parameters that were 1.2 times that of the associated C atoms and participated in the calculation of the final R indices. CCDC-1487278 (1) and CCDC-1487279 (2) contain the supplementary crystallographic data for this paper.
Computer modeling
Molecular structures were optimized using DFT (hybrid Hartree-Fock DFT functional B3LYP level, which is a combination of the Becke 3-parameter exchange and Lee–Yang–Parr correlation functional, using the Gaussian 09 program package). In all cases, the LANL2DZ basis set was used for Ru atom and 6-31G(d) basis set for others. In all energy-minimized calculated structures, the optimized structure was confirmed to be a minimum from vibrational frequency calculations.32
Syntheses. All of the reactions were carried out under strictly anaerobic conditions in an inert argon atmosphere by using standard Schlenk techniques.
Synthesis of (PPh4)2[RuII(PPh3)2(CN)4] (1). [RuII(PPh3)3Cl2] (1.0 g, 1.04 mmol) and KCN (0.41 g, 6.26 mmol) in MeOH was heated at reflux overnight. The solvent was removed under the reduced pressure. The white residue was dissolved in water (500 mL) and then PPh4Cl (1.12 g, 3 mmol) was added to give product as white precipitate. The crystals suitable for X-ray determination were obtained by slow evaporation of a MeOH/H2O solution of the complex. Yield: 76.9%, 1.13 g. IR (KBr, cm−1): ν(C
N) 2059, 2100. 1H NMR (300 MHz, CDCl3): δ 7.91–7.96 (m, 14H, Ar–H), 7.82–7.71 (m, 28H, Ar–H), 7.21–7.27 (m, 28H, Ar–H). Elemental analysis for C88H70N4P4Ru: calcd C 75.04, H 5.01, N 3.98%; found C 75.11, H 5.28, N 3.79%. UV/vis (CH3CN): λmax/nm (ε/mol−1 dm3 cm−1): 268 (12
680), 275 (10
140), 310 sh (1800). ESI-MS: m/z 365.0 [M]2−, 1069.0 [M + PPh4]+.
Synthesis of (PPh4)[RuII(PPh3)(bpyOH)(CN)3] (2). A mixture of 1 (100 mg, 0.07 mmol) and bpyOH (14.7 mg, 0.08 mmol) in DMF (25 mL) was heated at reflux overnight under a nitrogen atmosphere. After removal of the solvent under reduced pressure, the residue was washed with Et2O and further purified by column chromatography on silica gel with CH2Cl2/MeOH (v/v 1
:
4) as the eluent to give orange band. Crystals suitable for X-ray determination were obtained by slow diffusion of Et2O vapour into the concentrated CH2Cl2 solution of 2. Yield: 66.5%, 45 mg. Elemental analysis for C55H43N5OP2Ru: calcd C 69.32, H 4.55, N 7.35%; found C 69.40, H 4.70, N 7.22%; IR (KBr, cm−1): ν(C
N) 2093, 2057, 2045. 1H NMR (300 MHz, CDCl3): δ 10.15 (s, 1H; Ar–OH), 9.03 (d, J = 5.4 Hz, 1H; bpy-H), 8.19 (d, J = 8.2 Hz, 1H; bpy-H), 8.04–7.94 (m, 3H, bpy-H), 7.92–7.68 (m, 17H, phenyl-H), 7.65 (d, J = 9.2 Hz, 1H, bpy-H), 7.35–7.05 (m, 18H, phenyl-H), 6.63 (d, J = 8.1 Hz, 1H, bpy-H); 31P{1H} NMR (162 MHz, CDCl3): δ 41.02 (s, PPh3), 22.32 (s, PPh4+). UV/vis (CH3CN): λmax/nm (ε/mol−1 dm3 cm−1): 260 sh (9790), 268 (9740), 276 (9540), 307 (12
680), 330 sh (5820), 433 (2200). ESI-MS: m/z 614 [M]−.
Synthesis of (PPh4)[RuII(PPh3)(phenOH)(CN)3] (3). The complex was synthesized using similar synthetic procedure of 2 except that phenOH was used instead of bpyOH. Yield: 57.6%, 40 mg. Elemental analysis for C57H43N5OP2Ru: calcd C 70.07, H 4.44, N 7.17%; found C 70.12, H 4.48, N 7.10%. IR (KBr, cm−1): ν(C
N) 2089, 2082, 2075. 1H NMR (300 MHz, CDCl3): δ 10.69 (s, 1H; Ar–OH), 9.26 (d, J = 5.1 Hz, 1H; phen H), 8.39–8.27 (m, 2H, phen H), 8.03–7.93 (m, 6H, phenyl H), 7.90–7.67 (m, 21H, phenyl H), 7.49 (dd, J = 8.2, 5.2 Hz, 1H; phen H), 7.21–6.98 (m, 13H, phenyl H); 31P{1H} NMR (162 MHz, CDCl3): δ 41.84 (s, PPh3), 22.32 (s, PPh4+). UV/vis (CH3CN): λmax/nm (ε/mol−1 dm3 cm−1): 268 (29
020), 274 sh (26
170), 295 sh (10
290), 383 (3000), 433 (3580). ESI-MS: m/z 638 [M]−.
Conclusion
Two mononuclear tricyanoruthenium(II) compounds fac-(PPh4)[RuII(bpyOH)(PPh3)(CN)3] (2) and fac-(PPh4)[RuII(phenOH)(PPh3)(CN)3] (3) have been obtained from the reactions of (PPh4)2[RuII(PPh3)2(CN)4] (1) with bpyOH and phenOH, respectively, in DMF. Both complexes display impressive quantum yields in various solvents, which are much higher than those of related ruthenate(II) diimine complexes. Due to the presence of two different and spatially separate CN− and −OH sites in these two complexes, solvatochromic behaviour, pH and ion effects of 2 and 3 are distinct from those of [RuII(phen)(PPh3)(CN)3]−. Our results demonstrate that these complexes useful for the development of new luminescent sensors/probes and sensitizers.
Acknowledgements
The authors gratefully acknowledge the financial support of Natural Science Foundation of China (21201023). This work was also supported by the Hong Kong University Grants Committee Area of Excellence Scheme (AoE/P-03-08).
Notes and references
-
(a) A. P. De Silva, H. Q. N. Gunaratne, T. Gunnlaugsson, A. J. M. Huxley, C. P. McCoy, J. T. Rademacher and T. E. Rice, Chem. Rev., 1997, 97, 1515–1566 CrossRef CAS PubMed;
(b) J. N. Demas and B. A. DeGraff, Coord. Chem. Rev., 2001, 211, 317–351 CrossRef CAS.
-
(a) E. Baggaley, S. W. Botchway, J. W. Haycock, H. Morris, I. V. Sazanovich, J. A. G. Williams and J. A. Weinstein, Chem. Sci., 2014, 5, 879–886 RSC;
(b) Q. Ju, Y. Liu, D. Tu, H. Zhu, R. Li and X. Chen, Chem.–Eur. J., 2011, 17, 8549–8554 CrossRef CAS PubMed;
(c) Y. Ma, S. J. Liu, H. R. Yang, Y. Q. Wu, H. B. Sun, J. X. Wang, Q. Zhao, F. Y. Li and W. J. Huang, J. Mater. Chem. B, 2013, 1, 319–329 RSC;
(d) H. E. Rajapakse, N. Gahlaut, S. Mohandessi, D. Yu, J. R. Turner and L. W. Miller, Proc. Natl. Acad. Sci. U. S. A., 2010, 107, 13582–13587 CrossRef CAS PubMed;
(e) J. N. Demas and B. A. DeGraff, Anal. Chem., 1991, 63, 829A–837A CrossRef CAS.
- J. N. Demas, B. A. DeGraff and P. B. Coleman, Anal. Chem., 1999, 71, 793A–800A CrossRef CAS PubMed.
- A. Mills and K. Eaton, Quim. Anal., 2000, 19, 75–87 CAS.
- C. Malins, H. G. Glever, T. E. Keyes, J. G. Vos, W. J. Dressick and B. D. MacCraith, Sens. Actuators, B, 2000, 67, 89–95 CrossRef CAS.
- C. Huber, I. Klimant, C. Krause and O. S. Wolfbeis, Anal. Chem., 2001, 73, 2097–2103 CrossRef CAS PubMed.
- J. N. Demas and B. A. DeGraff, Proc. SPIE-Int. Soc. Opt. Eng., 1992, 1796, 71–75 CrossRef.
-
(a) Y. Sun, Z. M. Hudson, Y. L. Rao and S. N. Wang, Inorg. Chem., 2011, 50, 3373–3378 CrossRef CAS PubMed;
(b) R. Zhang, Z. Q. Ye, B. Song, Z. C. Dai, X. An and J. L. Yuan, Inorg. Chem., 2013, 52, 10325–10331 CrossRef CAS PubMed;
(c) Z. Q. Ye, X. An, B. Song, W. Z. Zhang, Z. C. Dai and J. L. Yuan, Dalton Trans., 2014, 43, 13055–13060 RSC;
(d) R. Zhang, X. J. Yu, Z. Q. Ye, G. L. Wang, W. Z. Zhang and J. L. Yuan, Inorg. Chem., 2010, 49, 7898–7903 CrossRef CAS PubMed;
(e) H. Xu, H. C. Hu, C. S. Cao and B. Zhao, Inorg. Chem., 2015, 54, 4585–4587 CrossRef CAS PubMed.
-
(a) A. Juris, V. Balzani, F. Barigelletti, S. Campagna, P. Belser and A. Von Zelewsky, Coord. Chem. Rev., 1988, 84, 85–277 CrossRef CAS;
(b) V. Balzani, A. Juris, M. Venturi, S. Campagna and S. Serroni, Chem. Rev., 1996, 96, 759–833 CrossRef CAS PubMed and references therein.
(c) A. K. D. Mesmaeker, J. P. Lecomte and J. M. Kelly, Top. Curr. Chem., 1996, 177, 25–76 CrossRef CAS;
(d) V. Balzani and F. Scandola, in Comprehensive Supramolecular Photochemistry, ed. J. Atwood, D. MacNicol, E. Davies, F. Vogtle and J. M. Lehn, Pergamon Press, Oxford, U.K., 1996, vol. 10, p. 687 Search PubMed.
-
(a) C. A. Bignozzi, C. Chiorboli, M. T. Indelli, M. A. R. Scandola, G. Varani and F. J. Scandola, J. Am. Chem. Soc., 1986, 108, 7872–7873 CrossRef CAS PubMed;
(b) A. Ülveczky and A. Horváth, Inorg. Chim. Acta, 1995, 236, 173–176 CrossRef.
-
(a) J. L. Habib Jiwan, B. Wegewijs, M. T. Indelli, F. Scandola and S. E. Braslavsky, Recl. Trav. Chim. Pays-Bas, 1995, 114, 542 CrossRef CAS;
(b) M. D. Ward, Coord. Chem. Rev., 2006, 250, 3128 CrossRef CAS;
(c) C. J. Timpson, C. A. Bignozzi, B. P. Sullivan, E. M. Kober and T. J. Meyer, J. Phys. Chem., 1996, 100, 2915 CrossRef CAS.
-
(a) C. D. Borsarelli, S. E. Braslavsky, M. T. Indelli and F. Scandola, Chem. Phys. Lett., 2000, 317, 53–58 CrossRef CAS;
(b) M. A. Rampi, M. T. Indelli, F. Pina, F. Scandola and A. J. Parola, Inorg. Chem., 1996, 35, 3355–3361 CrossRef CAS PubMed;
(c) J. K. Evju and K. R. Mann, Chem. Mater., 1999, 11, 1425–1433 CrossRef CAS;
(d) C. E. Keller, C. Pollard, L. K. Yeung, W. D. Plessinger and C. J. Murphy, Inorg. Chim. Acta, 2000, 298, 209–215 CrossRef CAS;
(e) T. Lazarides, T. L. Easun, C. Veyne–Marti, W. Z. Alsindi, M. W. George, N. Deppermann, C. A. Hunter, H. Adams and M. D. Ward, J. Am. Chem. Soc., 2007, 129, 4014–4027 CrossRef CAS PubMed;
(f) F. Pina and A. J. Parola, Coord. Chem. Rev., 1999, 149, 185–186 Search PubMed;
(g) T. Megyes, G. Schubert, M. Kovács, T. Radnai, T. Grósz, I. Bakó, I. Pápai and A. Korváth, J. Phys. Chem. A, 2003, 107, 9903–9909 CrossRef;
(h) M. D. Ward, Coord. Chem. Rev., 2006, 250, 3128–3141 CrossRef CAS;
(i) G. Bergamini, C. Saudan, P. Ceroni, M. Maestri, V. Balzani, M. Groka, S. K. Lee, J. V. Heyst and F. Vögtle, J. Am. Chem. Soc., 2004, 126, 16466–16471 CrossRef CAS PubMed.
-
(a) M. D. Ward, Dalton Trans., 2010, 39, 8851–8867 RSC;
(b) T. Abe, T. Suzuki and K. Shinozaki, Inorg. Chem., 2010, 49, 1794–1800 CrossRef CAS PubMed;
(c) S. Encinas, A. F. Morales, F. Barigelletti, A. M. Barthram, C. M. White, S. M. Couchman, J. C. Jeffery, M. D. Ward, D. C. Grills and M. W. George, J. Chem. Soc., Dalton Trans., 2001, 3312–3319 RSC;
(d) M. E. Garca Posse, N. E. Katz, L. M. Baraldo, D. D. Polonuer, C. G. Colombano and J. A. Olabe, Inorg. Chem., 1995, 34, 1830–1835 CrossRef;
(e) S. G. Baca, H. Adams, C. S. Grange, A. P. Smith, I. Sazanovich and M. D. Ward, Inorg. Chem., 2007, 46, 9779–9789 CrossRef CAS PubMed;
(f) L. T. –L. Lo, S. W. Lai, S. M. Yiu and C. C. Ko, Chem. Commun., 2013, 49, 2311–2313 RSC.
- H. Feng, F. Zhang, S. W. Lai, S. M. Yiu and C. C. Ko, Chem.–Eur. J., 2013, 19, 15190–15198 CrossRef CAS PubMed.
-
(a) Y.-H. Chi, L. Yu, J. M. Shi, Y. Q. Zhang, T. Q. Hu, G. Q. Zhang, W. Shi and P. Cheng, Dalton Trans., 2011, 40, 1453–1462 RSC;
(b) X. M. Zhang, M. L. Tong, M. L. Gong, H. K. Lee, L. Luo, K. F. Li, Y. X. Tong and X. M. Chen, Chem.–Eur. J., 2002, 8, 3187–3194 CrossRef CAS PubMed;
(c) X. M. Zhang, M. L. Tong and X. M. Chen, Angew. Chem., Int. Ed., 2002, 41, 1029–1031 CrossRef CAS;
(d) K. Fujita, Y. Tanaka, M. Kobayashi and R. Yamaguchi, J. Am. Chem. Soc., 2014, 136, 4829–4832 CrossRef CAS PubMed;
(e) R. Kawahara, K. Fujita and R. Yamaguchi, Angew. Chem., Int. Ed., 2012, 51, 12790–12794 CrossRef CAS PubMed;
(f) T. Tomon, T. Koizumi and K. Tanaka, Eur. J. Inorg. Chem., 2005, 285–293 CrossRef CAS.
-
(a) J. Xiang, L. T.-L. Lo, C.-F. Leung, S.-M. Yiu, C.-C. Ko and T.-C. Lau, Organometallics, 2012, 31, 7101–7108 CrossRef CAS;
(b) J. Xiang, J.-S. Wu and Z. Anorg, Z. Anorg. Allg. Chem., 2013, 639, 606–610 CrossRef CAS.
- N. Nickita, M. J. Belouscoff, A. I. Bhatt, A. M. Bond, G. B. Deacon, G. Gasser and L. Spiccia, Inorg. Chem., 2007, 46, 8638–8651 CrossRef CAS PubMed.
-
(a) J. V. Houten and R. J. Watts, J. Am. Chem. Soc., 1975, 97, 3843 CrossRef;
(b) D. P. Rillema, G. Allen, T. J. Meyer and D. Gonrad, Inorg. Chem., 1983, 22, 1617 CrossRef CAS;
(c) E. M. Kober, B. P. Sullivan and T. J. Meyer, Inorg. Chem., 1984, 23, 2098 CrossRef CAS;
(d) J. V. Caspar and T. J. Meyer, J. Am. Chem. Soc., 1983, 105, 5583 CrossRef CAS;
(e) A. Juris, V. Balzani, F. Barigelletti, S. Campagna, P. Belser and A. Von Zelewsky, Coord. Chem. Rev., 1995, 33, 2438 Search PubMed.
-
(a) M. T. Indeli, M. Ghirotti, A. Prodi, C. Chiorboli, F. Scandola, N. D. McClenaghan, F. Puntoriero and S. Camanga, Inorg. Chem., 2003, 42, 5489–5497 CrossRef PubMed;
(b) N. R. M. Simpson, M. D. Ward, A. F. Morales, B. Ventura and F. Barigelletti, J. Chem. Soc., Dalton Trans., 2002, 2455–2461 RSC;
(c) N. R. M. Simpson, M. D. Ward, A. F. Morales, B. Ventura and F. Barigelletti, J. Chem. Soc., Dalton Trans., 2002, 2449–2454 RSC;
(d) D. J. Stufkens, Inorg. Chem., 1992, 13, 359 CAS;
(e) T. G. Kotch and A. J. Lees, Inorg. Chem., 1993, 32, 2570 CrossRef CAS.
- M. E. Garca Posse, N. E. Katz, L. M. Baraldo, D. D. Polonuer, C. G. Colombano and J. A. Olabe, Inorg. Chem., 1995, 34, 1830–1835 CrossRef.
-
(a) C. Hicks, G.-Z. Ye, C. Levi, M. Gonzales, I. Rutenburg, J.-W. Fan, R. Helmy, A. Kassis and H. D. Gafney, Coord. Chem. Rev., 2001, 211, 207 CrossRef CAS;
(b) Md. K. Nazeeruddin and K. Kalyanasundaram, Inorg. Chem., 1989, 28, 4251 CrossRef CAS;
(c) G. Y. Zheng, Y. Wang and D. P. Rillema, Inorg. Chem., 1996, 35, 7118 CrossRef CAS PubMed;
(d) F. Casalboni, Q. G. Mulazzani, C. D. Clark, M. Z. Hoffman, P. L. Orizondo, M. W. Perkovic and D. P. Rillema, Inorg. Chem., 1997, 36, 2252 CrossRef CAS PubMed;
(e) Md. K. Nazeeruddin, S. M. Zakeeruddin, R. Humphry-Baker, M. Jirousek, P. Liska, N. Vlachopoulos, V. Shklover, C.-H. Fischer and M. Gratzel, Inorg. Chem., 1999, 38, 6298 CrossRef CAS PubMed.
-
(a) A. M. W. C. Thompson, M. C. C. Smailes, J. C. JeVery and M. D. Ward, J. Chem. Soc., Dalton Trans., 1997, 737–744 RSC;
(b) J. G. Vos, Polyhedron, 1992, 11, 2285–2299 CrossRef CAS.
- J. R. Lakowicz, Principles of Fluorescence Spectroscopy, Springer–Verlag, New York, 3rd edn, 2006 Search PubMed.
- W. Cao, X. J. Zheng, D. C. Fang and L. P. Jin, Dalton Trans., 2014, 43, 7298–7303 RSC.
-
(a) A. Fodor–Kardos and A. Horváth, Photochem. Photobiol. Sci., 2005, 4, 185–190 RSC;
(b) T. Lazarides, T. L. Easun, C. Veyne–Marti, W. Z. Alsindi, M. W. George, N. Deppermann, C. A. Hunter, H. Adams and M. D. Ward, J. Am. Chem. Soc., 2007, 129, 4014–4027 CrossRef CAS PubMed.
- M. Zalas, B. Gierczyk, M. Cegłowski and G. Schroeder, Chem. Pap., 2012, 66, 733–740 CAS.
-
(a) T. A. Stephenson and G. J. Wilkinson, Inorg. Nucl. Chem. Lett., 1966, 28, 945–956 CrossRef CAS;
(b) P. S. Hallman, T. A. Stephenson and G. J. Wilkinson, Inorg. Synth., 1970, 12, 237–240 CAS;
(c) S. Nishimura, O. Yumoto, K. Tsuneda and H. Mori, Bull. Chem. Soc. Jpn., 1975, 48, 2603–2604 CrossRef CAS.
- J. N. Demas and G. A. Crosby, J. Phys. Chem., 1971, 75, 991–1024 CrossRef.
- J. Van Houten and R. Watts, J. Am. Chem. Soc., 1976, 98, 4853–4858 CrossRef CAS.
- G. M. Sheldrick, SHELX–97: Programs for Crystal Structure Analysis (Release 97–2), University of Göttingen, Göttingen, Germany, 1997 Search PubMed.
- A. Altomare, G. Cascarano, C. Giacovazzo, A. Guagliardi, M. Burla, G. Polidori and M. J. Camalli, Appl. Crystallogr., 1994, 27, 435–436 Search PubMed.
- L. Salassa, C. Garino, G. Salassa, R. Gobetto and C. Nervi, J. Am. Chem. Soc., 2008, 130, 9590 CrossRef CAS PubMed.
Footnotes |
† Electronic supplementary information (ESI) available. CCDC 1487278 and 1487279. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra16319j |
‡ These two authors contributed equally. |
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.