Synthesis, photophysical properties and application in organic light emitting devices of rhenium(I) carbonyls incorporating functionalized 2,2′:6′,2′′-terpyridines

Tomasz Klemensa, Anna Świtlicka-Olszewskaa, Barbara Machura*a, Marzena Grucelab, Henryk Janeczekb, Ewa Schab-Balcerzak*bc, Agata Szlapad, Slawomir Kulad, Stanisław Krompiecd, Karolina Smolareke, Dorota Kowalskae, Sebastian Mackowskie, Karol Erfurtf and Piotr Lodowskig
aDepartment of Crystallography, Institute of Chemistry, University of Silesia, 9th Szkolna St., 40-006 Katowice, Poland. E-mail: bmachura@poczta.onet.pl
bCentre of Polymer and Carbon Materials, Polish Academy of Sciences, 34M. Curie-Sklodowska Str., 41-819 Zabrze, Poland
cDeparment of Polymer Chemistry, University of Silesia, 9th Szkolna St., 40-006 Katowice, Poland. E-mail: ewa.schab-balcerzak@us.edu.pl
dDepartment of Inorganic, Organometallic Chemistry and Catalysis, Institute of Chemistry, University of Silesia, 9th Szkolna St., 40-006 Katowice, Poland
eInstitute of Physics, Faculty of Physics, Astronomy and Informatics, Nicolaus Copernicus University, 5 Grudziadzka Str., 87-100 Torun, Poland
fDepartment of Chemical Organic Technology and Petrochemistry, Silesian University of Technology, Krzywoustego 4, 44-100 Gliwice, Poland
gDepartment of Theoretical Chemistry, Institute of Chemistry, University of Silesia, 9th Szkolna St., 40-006 Katowice, Poland

Received 7th April 2016 , Accepted 31st May 2016

First published on 3rd June 2016


Abstract

Several new rhenium(I) complexes [ReCl(CO)3(4′-R-terpy-κ2N)] incorporating 2,2′:6′,2′′-terpyridine-based ligands were successfully synthetized and characterized by IR, NMR (1H and 13C), UV-vis spectroscopy and single crystal X-ray analysis. The luminescent properties of [ReCl(CO)3(4′-R-terpy-κ2N)] were studied in solution and solid state, at 298 and 77 K. To better understand the photophysical properties of [ReCl(CO)3(4′-R-terpy-κ2N)], density functional theory (DFT) and time-dependent DFT (TD-DFT) calculations were performed. Preliminary studies towards application of these complexes in organic light emitting diodes (OLEDs) were carried out, including testing the possibility of electroluminescence intensity increase by including metallic nanowires in the structure design.


Introduction

Since the first observation of electronically excited luminescence in [ReCl(CO)3(4,7-(Ph)2phen)] and [ReCl(CO)3(5-R-phen)] (R – H, CH3, Cl, Br, NO2),1 rhenium(I) tricarbonyl complexes based on 1,10-phenanthroline, 2,2′-bipyridine or related bidentate diimine derivatives [Re(CO)3La(NN)]n+ (La – axial ligand; n = 0 or 1) have attracted significant attention due to their photochemical and photophysical properties as well as potential applicability.2 The studies revealed that the axial ligand L and structural variations of N,N ligand strongly affect the nature and energetic order of low-lying excited states and, thus, the photophysics, photochemistry and electrochemistry of [Re(CO)3La(NN)]n+ complexes. Depending on the relative energy levels of the metal and ligand orbitals, the occurrence of various excited states (metal-to-ligand charge transfer (MLCT), ligand-to-ligand charge transfer (LLCT), σ-bond-to-ligand charge transfer (σ → π*) and intraligand (IL)) is possible, as well as tuning of interaction between them.3 Importantly, most of [Re(CO)3La(NN)]n+ exhibit good phosphorescence properties thanks to the strong spin–orbit coupling induced by the Re ion that enhances the singlet–triplet mixing, affording a fairly long-lived excited state and appreciable quantum efficiency of emission. Intersystem crossing processes (ISC) in this class of compounds were interpreted on the basis of density functional theory (DFT) calculations including spin–orbit coupling (SOC) effects.4 Richness and variability of the photophysical and photochemical behaviours make these compounds attractive for solar energy conversion,5 luminescence sensing,6 for photosensitization of organic substrates,7 organic light-emitting diodes (OLEDs)8 and in other fields such as catalysis9 and medicinal chemistry.10

Herein, we present the synthesis, characterization and photophysical properties of six new Re(I) complexes incorporating 2,2′:6′,2′′-terpyridine ligand framework (Scheme 1). In particular, we aim to elucidate the possibility to tune the emission performance of Re(I)-terpyridine carbonyls by structural modifications of terpy ligand. 2,2′:6′,2′′-Terpyridine derivatives are noticeable ligands in coordination chemistry, but investigations of Re(I)-terpyridine carbonyls are scarce and the structure–property relationship remains unexplored.11 The paucity of data on Re(I)-terpyridine carbonyls can be attributed to the fact that [Re(CO)3(terpy)Cl], synthesized in 1988 by Juris and co-workers,12 was found to be nonluminescent in solution at room temperature. Recent studies, however, revealed the possibility of improving the emission performance of [ReCl(CO)3(4′-R-terpy-κ2N)] by structural modifications of terpy ligand and proved that Re(I) complexes incorporating 2,2′:6′,2′′-terpyridine ligand framework are suitable candidates for OLED materials.11c,13


image file: c6ra08981j-s1.tif
Scheme 1 2,2′:6′,2′′-Terpyridine derivatives employed in this study.

The luminescent properties of [ReCl(CO)3(4′-R-terpy-κ2N)] were studied in solution and solid state, at 298 and 77 K. To get detailed insight into their electronic structure and spectroscopic properties, the density functional theory (DFT) and time-dependent DFT (TD-DFT) calculations were performed.

Experimental section

Materials

Re(CO)5Cl was commercially available (Sigma Aldrich) and was used without further purification. Poly(9-vinylcarbazole) PVK (Mn = 25[thin space (1/6-em)]000–50[thin space (1/6-em)]000) was purchased from Sigma Aldrich and used without additional purification. Poly(3,4-(ethylenedioxy)thiophene)[thin space (1/6-em)]:[thin space (1/6-em)]poly-(styrenesulfonate) (PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS) (0.1–1.0 S cm−1) and substrates with pixilated ITO anodes were supplied by Ossila. 2,2′:6′,2′′-Terpyridine derivatives were obtained as described in our previous work.14 NMR spectra of those derivatives were also reported previously.14 All solvents for synthesis were of reagent grade and were used as received. For spectroscopy studies HPLC grade solvents were used.

Silver nanowires (AgNWs) were synthesized using the polyol process, in which ethylene glycol (EG) served as the reducing and solvent reagent.15 The product was purified by centrifugation process and the mixture was diluted with isopropyl alcohol and centrifuged. The supernatant containing silver particles and unreacted substrates was removed. Finally the product was redispersed in 2 mL of pure water. Scanning electron microscopy yields diameters of the nanowires in the range from 50 to 150 nm, while their lengths range from 4 to even 30 microns.16 The samples for investigating the influence of silver nanowires on the electroluminescence properties of the devices were fabricated by placing the nanowires into PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS layer. The morphology of the samples were not substantially affected by the presence of the metallic nanoparticles.

General synthesis route for [ReCl(CO)3(4′-R-terpy-κ2N)] complexes (1–6)

Complexes 1–5. Re(CO)5Cl (0.10 g, 0.27 mmol) and suitable 4′-R-terpy ligand (0.27 mmol) were dissolved in acetonitrile (80 mL) and refluxed under argon for 6 hours. The resulting reaction solution was reduced in volume to 10 mL and allowed to cool to room temperature. The resulting yellow (1, 2, 3 and 4) or orange (5) solid was collected by filtration washed with diethyl ether, and dried. X-ray quality crystals were obtained by slow recrystallization from acetonitrile.
Complex 6. Re(CO)5Cl (0.10 g, 0.27 mmol) and suitable 4′-R-terpy ligand (0.27 mmol) were dissolved in argon-saturated acetonitrile (80 mL). Resulting solution was heated in autoclave for 24 hours to 150 °C, remained in that temperature for 30 hours and was cooled to room temperature for another 30 hours. X-ray quality brown crystals were obtained by filtration and washed with diethyl ether.
[ReCl(CO)3(4′-R1-terpy-κ2N)] (1). Yield: 75%. IR (KBr, cm−1): 2021(vs), 1899(vs) and 1887(vs) ν(C[triple bond, length as m-dash]O); 1615(m) ν(C[double bond, length as m-dash]N) and ν(C[double bond, length as m-dash]C). 1H NMR (400 MHz, DMSO-d6): δ/ppm = 9.15 (d, J = 10.6 Hz, 2H), 9.08 (d, J = 5.0 Hz, 1H), 8.81 (d, J = 4.2 Hz, 1H), 8.41 (t, J = 8.0 Hz, 1H), 8.33 (d, J = 8.2 Hz, 2H), 8.26 (s, 1H), 8.07 (t, J = 7.7 Hz, 1H), 7.94 (d, J = 7.3 Hz, 3H), 7.80 (dd, J = 15.4, 7.5 Hz, 3H), 7.67–7.61 (m, 1H), 7.53 (t, J = 7.4 Hz, 2H), 7.44 (t, J = 7.2 Hz, 1H). 13C NMR (125 MHz, DMSO-d6): δ/ppm = 198.29, 194.95, 191.49 (3CO), 161.88, 158.30, 157.63, 156.74, 153.19, 150.53, 149.72, 142.92, 140.42, 139.34, 137.43, 134.04, 129.58, 128.96, 128.70, 127.93, 127.35, 125.98, 125.64, 125.45, 124.60, 120.89. HRMS (ESI): calcd for C30H19N3O3Re [M − Cl]+ 656.0985 found 656.0981. DSC Tm = 330 °C.
[ReCl(CO)3(4′-R2-terpy-κ2N)] (2). Yield: 70%. IR (KBr, cm−1): 2019(vs), 1911(vs) and 1885(vs) ν(C[triple bond, length as m-dash]O); 1609(s) and 1577(m) ν(C[double bond, length as m-dash]N) and ν(C[double bond, length as m-dash]C). 1H NMR (400 MHz, DMSO-d6): δ/ppm = 9.09 (d, J = 5.2 Hz, 1H), 8.99 (s, 1H), 8.96 (d, J = 8.2 Hz, 1H), 8.79 (d, J = 4.4 Hz, 1H), 8.38–8.29 (m, 2H), 8.05 (t, J = 7.6 Hz, 1H), 7.96 (d, J = 6.8 Hz, 3H), 7.77 (d, J = 8.0 Hz, 2H), 7.69–7.58 (m, 3H), 7.20 (d, J = 8.1 Hz, 1H), 4.08 (s, 3H). 13C NMR (100 MHz, DMSO-d6): δ/ppm = 198.21, 194.99, 191.50 (3CO), 161.41, 158.20, 157.17, 156.80, 153.21, 152.31, 149.74, 140.49, 137.38, 131.25, 129.59, 128.58, 128.50, 127.92, 127.37, 126.47, 125.90, 125.72, 125.50, 125.43, 125.02, 124.71, 122.71, 104.81, 56.50. HRMS (ESI): calcd for C29H19N3O4Re [M − Cl]+ 660.0934 found 660.0935. DSC Tm = 279, 294 °C.
[ReCl(CO)3(4′-R3-terpy-κ2N)] (3). Yield: 65%. IR (KBr, cm−1): 2023(vs), 1944(vs) and 1911(vs) ν(C[triple bond, length as m-dash]O); 1615(m) and 1589(m) ν(C[double bond, length as m-dash]N) and ν(C[double bond, length as m-dash]C). 1H NMR (400 MHz, DMSO-d6): δ/ppm = 9.09 (dd, J = 12.0, 6.3 Hz, 3H), 8.80 (d, J = 4.2 Hz, 1H), 8.40 (t, J = 7.7 Hz, 1H), 8.26 (d, J = 8.4 Hz, 2H), 8.22 (s, 1H), 8.06 (t, J = 7.6 Hz, 1H), 7.91 (d, J = 7.7 Hz, 1H), 7.81–7.76 (m, 1H), 7.70 (d, J = 8.4 Hz, 2H), 7.66–7.60 (m, 1H). 13C NMR (125 MHz, DMSO-d6): δ/ppm = 198.25, 194.90, 191.43 (3CO), 161.93, 158.19, 157.69, 156.64, 153.21, 149.76, 149.71, 140.43, 137.43, 136.44, 134.00, 130.22, 129.81, 128.00, 126.00, 125.63, 125.47, 124.79, 121.04. HRMS (ESI): calcd for C24H14N3O3ReCl [M − Cl]+ 614.0273 found 614.0266. DSC Tm = 335 °C.
[ReCl(CO)3(4′-R4-terpy-κ2N)] (4). Yield: 80%. IR (KBr, cm−1): 2022(vs), 1942(vs) and 1911(vs) ν(C[triple bond, length as m-dash]O); 1615(m) and 1589(m) ν(C[double bond, length as m-dash]N) and ν(C[double bond, length as m-dash]C). 1H NMR (400 MHz, DMSO-d6): δ/ppm = 9.09 (dd, J = 12.3, 6.7 Hz, 3H), 8.80 (d, J = 4.4 Hz, 1H), 8.40 (t, J = 7.8 Hz, 1H), 8.22 (d, J = 6.3 Hz, 1H), 8.19 (d, J = 8.5 Hz, 2H), 8.06 (t, J = 7.4 Hz, 1H), 7.91 (d, J = 7.8 Hz, 1H), 7.83 (d, J = 8.5 Hz, 2H), 7.81–7.76 (m, 1H), 7.67–7.60 (m, 1H). 13C NMR (125 MHz, DMSO-d6): δ/ppm = 198.24, 194.89, 191.41 (3CO), 161.95, 158.19, 157.71, 156.63, 153.21, 149.88, 149.71, 140.44, 137.43, 134.38, 132.75, 130.42, 128.01, 126.00, 125.64, 125.48, 125.35, 124.75, 121.00. HRMS (ESI): calcd for C24H14N3O3ReBr [M − Cl]+ 657.9759 found 657.9749. DSC Tm = 336 °C.
[ReCl(CO)3(4′-R5-terpy-κ2N)] (5). Yield: 75%. IR (KBr, cm−1): 2017(vs), 1938(vs) and 1884(vs) ν(C[triple bond, length as m-dash]O); 1612(m) and 1512(m) ν(C[double bond, length as m-dash]N) and ν(C[double bond, length as m-dash]C). 1H NMR (400 MHz, DMSO-d6): δ/ppm = 9.08 (d, J = 5.0 Hz, 1H), 9.05–8.95 (m, 2H), 8.80 (d, J = 4.2 Hz, 1H), 8.39 (t, J = 7.8 Hz, 1H), 8.14–8.02 (m, 3H), 7.89 (d, J = 7.8 Hz, 1H), 7.79 (t, J = 6.4 Hz, 1H), 7.67–7.53 (m, 2H), 7.40 (t, J = 8.4 Hz, 1H). 13C NMR (100 MHz, DMSO-d6): δ/ppm = 198.14, 194.84, 191.30 (3CO), 165.35, 162.87, 161.67, 159.13, 158.06, 157.40, 156.50, 153.27, 149.80, 146.15, 140.55, 137.46, 128.09, 125.63, 123.47, 120.86, 113.24, 105.63. HRMS (ESI): calcd for C24H13N3O3ReF2 [M − Cl]+ 616.0483 found 616.0485. DSC (I scan) Tm = 223 °C, Tc = 228 °C Tm = 282 °C, (II scan) Tg = 136 °C.
[ReCl(CO)3(4′-R6-terpy-κ2N)] (6). Yield: 75%. IR (KBr, cm−1): 2019(vs), 1924(vs) and 1892(vs) ν(C[triple bond, length as m-dash]O); 1600(s) and 1536(m) ν(C[double bond, length as m-dash]N) and ν(C[double bond, length as m-dash]C). 1H NMR (400 MHz, DMSO-d6): δ/ppm = 9.09 (d, J = 7.8 Hz, 1H), 9.04 (d, J = 5.3 Hz, 1H), 8.95 (s, 1H), 8.79 (s, 1H), 8.37 (t, J = 8.0 Hz, 1H), 8.12 (d, J = 8.9 Hz, 2H), 8.03 (d, J = 10.8 Hz, 2H), 7.86 (d, J = 7.7 Hz, 1H), 7.79–7.73 (m, 1H), 7.64–7.58 (m, 1H), 6.87 (d, J = 8.8 Hz, 2H), 3.06 (s, 6H). 13C NMR (100 MHz, DMSO-d6): δ/ppm = 198.32, 195.04, 191.67 (3CO), 161.45, 158.60, 157.11, 157.05, 153.05, 152.62, 150.84, 149.61, 140.31, 137.34, 129.29, 127.65, 125.62, 125.54, 125.25, 122.29, 120.97, 118.79, 112.53, 39.89. HRMS (ESI): calcd for C26H20N4O3Re [M − Cl]+ 623.1094 found 623.1092. DSC Tm = 347 °C.

Crystal structure determination and refinement

The X-ray intensity data of 1, 3, 4, 6 were collected on a Gemini A Ultra diffractometer equipped with Atlas CCD detector and graphite monochromated MoKα radiation (λ = 0.71073 Å) at room temperature (Table 1). The unit cell determination and data integration were carried out using the CrysAlis package of Oxford Diffraction. Lorentz, polarization and empirical absorption correction using spherical harmonics implemented in SCALE3 ABSPACK scaling algorithm were applied.17 The structures were solved by the Patterson method using SHELXS97 and refined by full-matrix least-squares on F2 using SHELXL97.18 All the non-hydrogen atoms were refined anisotropically, and hydrogen atoms were placed in calculated positions refined using idealized geometries (riding model) and assigned fixed isotropic displacement parameters, d(C–H) = 0.93 Å, Uiso(H) = 1.2Ueq.(C) (for aromatic); and d(C–H) = 0.96 Å, Uiso(H) = 1.5Ueq.(C) (for methyl). The methyl groups were allowed to rotate about their local threefold axis. CCDC reference numbers: 1471443 (1), 1471444 (3), 1471445 (4) and 1471446 (4).
Table 1 Crystal data and structure refinement of 1, 3, 4 and 6 complexes
  1 3 4 6
Empirical formula C30H19ClN3O3Re C24H14Cl2N3O3Re C24H14BrClN3O3Re C26H20ClN4O3Re
Formula weight 691.13 649.48 693.94 658.11
Temperature [K] 298.0(2) 298.0(2) 298.0(2) 298.0(2)
Wavelength [Å] 0.71073 0.71073 0.71073 0.71073
Crystal system Monoclinic Monoclinic Monoclinic Triclinic
Space group P21/c P21/c P21/c P[1 with combining macron]
Unit cell dimensions [Å, °] a = 14.7833(3) a = 9.4121(6) a = 9.4203(6) a = 7.2934(3)
b = 11.8128(3) b = 20.5407(9) b = 20.6584(8) b = 11.1958(6)
c = 30.1691(8) c = 11.8912(7) c = 11.8992(7) c = 15.1296(6)
      α = 91.434(4)
β = 100.552(2) β = 104.084(6) β = 104.218(6) β = 92.141(3)
      γ = 107.978(4)
Volume [Å3] 5179.4(2) 2229.8(2) 2244.7(2) 1173.40(9)
Z 8 4 4 2
Density (calculated) [Mg m−3] 1.773 1.935 2.053 1.863
Absorption coefficient [mm−1] 4.833 5.722 7.345 5.329
F(000) 2688 1248 1320 640
Crystal size [mm] 0.15 × 0.09 × 0.06 0.26 × 0.07 × 0.03 0.14 × 0.09 × 0.04 0.16 × 0.15 × 0.04
θ range for data collection [°] 3.30 to 25.05 3.31 to 25.05 3.44 to 25.05 3.66 to 25.05
Index ranges −17 ≤ h ≤ 17 −11 ≤ h ≤ 11 −8 ≤ h ≤ 11 −8 ≤ h ≤ 8
−12 ≤ k ≤ 14 −19 ≤ k ≤ 24 −24 ≤ k ≤ 18 −13 ≤ k ≤ 12
−35 ≤ l ≤ 32 −12 ≤ l ≤ 14 −14 ≤ l ≤ 14 −18 ≤ l ≤ 15
Reflections collected 30[thin space (1/6-em)]073 9767 9568 9547
Independent reflections 9158 (Rint = 0.033) 3933 (Rint = 0.040) 3965 (Rint = 0.051) 4145 (Rint = 0.0484)
Completeness to 2θ = 50° [%] 99.8 99.8 99.7 99.7
Max. and min. transmission 1.000 and 0.630 1.000 and 0.455 1.000 and 0.085 1.000 and 0.582
Data/restraints/parameters 9158/0/685 3933/0/298 3965/0/298 4145/0/318
Goodness-of-fit on F2 1.040 1.106 1.085 1.143
Final R indices [I > 2σ(I)] R1 = 0.0266 R1 = 0.0387 R1 = 0.0451 R1 = 0.0309
wR2 = 0.0512 wR2 = 0.0677 wR2 = 0.0971 wR2 = 0.0646
R indices (all data) R1 = 0.0385 R1 = 0.0540 R1 = 0.0581 R1 = 0.0409
wR2 = 0.0544 wR2 = 0.0716 wR2 = 0.1036 wR2 = 0.0863
Largest diff. peak and hole [eÅ−3] 0.601 and −0.596 0.769 and −0.671 1.732 and −2.151 1.055 and −0.948
CCDC number 1471443 1471444 1471445 1471446


Physical measurements

The IR spectra were recorded with a Nicolet iS5 FTIR spectrophotometer in the spectral range 4000–400 cm−1 with the samples in the form of KBr pellets. The electronic spectra were measured using Perkin Elmer Lambda 40 UV/vis spectrometer (in CH3CN and CHCl3 solution) and Jasco V570 UV-V-NIR spectrometer (in solid state as film deposited on glass substrate and as blends with poly(N-vinylcarbazole) (PVK) on glass substrate). The 1H NMR and 13C NMR spectra were recorded (295 K) on Bruker Avance 400 NMR spectrometer at a resonance frequency of 400 MHz for 1H NMR spectra and 100 MHz for 13C NMR spectra or on Bruker Avance 500 NMR spectrometer at a resonance frequency of 500 MHz for 1H NMR spectra and 125 MHz for 13C NMR spectra using DMSO-d6 or CDCl3 as solvent and TMS as an internal solvent.

Electrospray ionization mass spectrometry (ESI-MS) was performed on a Varian 500-MS IT Mass Spectrometer ion trap apparatus. The instrument was operating in the negative ion mode with a capillary voltage of 50 V, needle voltage of 5 kV and a spray shield voltage of 600 V. Drying gas (N2) temperature was 150 °C. High resolution mass spectrometry analyses were performed on a Waters Xevo G2 Q-TOF mass spectrometer (Waters Corporation) equipped with an ESI source operating in positive-ion modes. Full-scan MS data were collected from 100 to 1000 Da in positive ion mode with scan time of 0.1 s. To ensure accurate mass measurements, data were collected in centroid mode and mass was corrected during acquisition using leucine enkephalin solution as an external reference (Lock-Spray™), which generated reference ion at m/z 556.2771 Da ([M + H]+) in positive ESI mode. The accurate mass and composition for the molecular ion adducts were calculated using the MassLynx software (Waters) incorporated with the instrument.

Steady-state luminescence spectra of solid state and solution samples were measured with FLS-980 fluorescence spectrophotometer equipped with a 450 W Xe lamp and high-gain photomultiplier PMT + 500 nm (Hamamatsu, R928P) detector. The PL lifetime measurement was performed with a time correlated single photon counting (TCSPC) or multi-channel scaling (MCS) method. Excitation wavelength (375 nm or 470 nm) for TCSPC was obtained using the TCSPC diode with various pulse periods as light source. For MCS excitation wavelength was obtained using 60 W microsecond Xe flash lamp.

Photoluminescence spectra in solid state as film deposited on glass substrate and as blends with poly(N-vinylcarbazole) (PVK) on glass substrate were registered using Hitachi F-2500 spectrometer. In order to collect electroluminescence (EL) spectra the voltage was applied using a precise voltage supply (Gw Instek PSP-405) and the sample was fixed to an XYZ stage. Light from the OLED device was collected through a 30 mm lens, focused on the entrance slit (50 μm) of a monochromator (Shamrock SR-303i) and detected using a CCD detector (Andor iDus 12305). Typical acquisition times were equal to 10 seconds. The pre-alignment of the setup was done using a 405 nm laser. Differential Scanning Calorimetry (DSC) was performed with a TA-DSC 2010 apparatus, under nitrogen atmosphere using sealed aluminum pans with heating rate 20 °C min−1. Electrochemical measurements were carried out using ATLAS 0531 Electrochemical Unit & Impedance Analyzer potentiostat. Cyclic voltammetry experiments were conducted in acetonitrile (Aldrich, HPLC grade) and 0.2 M tetrabutylammonium hexafluorophosphate (Bu4NPF6 Aldrich, 99%) was used as the supporting electrolyte. Pt electrode, platinum coil and a quasi-Ag/AgCl electrode served as working, auxiliary and reference electrode, respectively. Potentials were referenced with respect to ferrocene (Fc), which was used as the internal standard. The HOMO and LUMO levels were calculated by assuming the absolute energy level of Fc/Fc+ as −5.1 eV to vacuum. Active layers thickness was measured by atomic force microscopy (AFM) Topometrix Explorer TMX 2000.

Film and blend preparation

Films and blends (15% concentration of compound in PVK) on glass substrate were prepared by spin coating (1000 rpm, 60 s) from chloroform solution (10 mg mL−1).

Devices preparation

Devices were prepared on OSILLA substrates with pixilated ITO anodes, cleaned sequentially with detergent, deionized water, 10% NaOH solution, water and isopropanol in an ultrasonic bath. OLEDs with three different configurations: ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/compound/Al, ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/PVK[thin space (1/6-em)]:[thin space (1/6-em)]Re(I)/Al and ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS[thin space (1/6-em)]:[thin space (1/6-em)]AgNWs/PVK[thin space (1/6-em)]:[thin space (1/6-em)]Re(I)/Al were fabricated. Substrates were covered with PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS or PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS[thin space (1/6-em)]:[thin space (1/6-em)]AgNWs thin film (40 nm) by spin coating at 5000 rpm for 60 s and annealed for 15 min at 120 °C. Active layer was spin-coated on top of the PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS layer from chloroform solution (10 mg mL−1) at 1000 rpm for 60 s. Finally an aluminum cathode (100 nm) was vacuum-deposited. Current density–voltage (JV) measurements were carried out using a Keithley 6517A source-measure unit.

Computational details

The calculations were performed using the GAUSSIAN-09 program package.19 The geometries of the singlet ground state (S0) and the lowest triplet state (T1) of 1–6 were fully optimized without any symmetry restrictions at the DFT level with the B3LYP hybrid exchange–correlation functional. The calculations were performed using the def2-TZVPD basis set for rhenium, 6-31+G** basis set for fluorine, chlorine, bromine, oxygen and nitrogen, 6-31G* for carbon and 6-31G for hydrogen atoms.20 The starting point for geometry optimization was taken from X-ray structure, and all the subsequent calculations were performed based on the optimized geometries. Vibrational frequencies were calculated on the basis of the optimized geometry to verify that each of the geometries is a minimum on the potential energy surface. Furthermore, on the basis of the optimized ground and excited state geometries, the absorption and emission properties in acetonitrile (MeCN) media were calculated by TD-DFT at the B3LYP hybrid functional level and with the polarized continuum model (PCM).21 The predicted bond lengths of 1, 3, 4, and 6 are compared with the experimental data in Table S1. The predicted bond lengths and angles for the ground state are within the range of error expected for DFT calculations of rhenium(I) complexes, and the general trends observed in the experimental data are well reproduced in the calculations, providing confidence on the reliability of the chosen method to reproduce the geometry of studied complexes (Table S1).

Results and discussion

Synthesis and general characterization of Re(I) complexes

The preparation of [ReCl(CO)3(4′-R1-terpy-κ2N)] (1), [ReCl(CO)3(4′-R2-terpy-κ2N)] (2), [ReCl(CO)3(4′-R3-terpy-κ2N)] (3), [ReCl(CO)3(4′-R4-terpy-κ2N)]·2H2O (4), [ReCl(CO)3(4′-R5-terpy-κ2N)] (5) and [ReCl(CO)3(4′-R6-terpy-κ2N)] (6) complexes was carried out using standard procedure by refluxing equimolar amounts of [Re(CO)5Cl] and 2,2′:6′,2′′-terpyridine under argon atmosphere.

The presence of three intense ν(C[triple bond, length as m-dash]O) bands in the region 2023–1873 cm−1 is consistent with a facial arrangement of the carbonyl groups and indicates bidentate chelate coordination of 4′-R-terpy ligand in 1–6. Medium intensity absorptions associated with the stretching modes ν(CN), ν(C[double bond, length as m-dash]C) modes of the 4′-R-terpy ligand occur in the range 1617–1512 cm−1.22

The unsymmetrical bidentate coordination mode of 4′-R-terpy in the examined complexes is also evidenced by NMR data as the resonances attributed to the terminal pyridine protons are clearly differentiated in both 1H and 13C NMR spectra.

Molecular structures of [ReCl(CO)3(4′-R-terpy-κ2N)]

Perspective views of the asymmetric units of [ReCl(CO)3(4′-R1-terpy-κ2N)] (1), [ReCl(CO)3(4′-R3-terpy-κ2N)] (3), [ReCl(CO)3(4′-R4-terpy-κ2N)] (4) and [ReCl(CO)3(4′-R6-terpy-κ2N)] (6) are presented in Fig. 1. The crystal structure of 1 comprises two independent molecules per asymmetric unit. As shown in Table S1, only small variations in bond lengths and bond angles can be noticed between these molecules. The crystal structures of the examined Re(I) complexes (1, 3, 4 and 6) are stabilized by intra- and intermolecular C–H⋯Cl, C–H⋯N type and C–H⋯N contacts (Table S2) as well as π–π type interactions (Fig. S1).
image file: c6ra08981j-f1.tif
Fig. 1 A perspective views showing the asymmetric units of 1, 3, 4 and 6 together with the atom numbering. Displacement ellipsoids are drawn at 50% probability.

In [ReCl(CO)3(4′-R-terpy-κ2N)] complexes, the Re(I) is six-coordinated in a distorted octahedral geometry defined by three facially arranged carbonyl ligands, two nitrogen atoms from two pyridine rings of the 4′-R-terpy ligand and chlorine atom. The major angular distortion from the idealized octahedral geometry is attributed to geometrical constraints issued from the occurrence of five-member chelate ring of the bidentate 4′-R-terpy ligand, resulting in N(2)–Re(1)–N(1) angle of 77.61(11)° for 1, 74.97(19)° for 3, 75.0(2)° for 4 and 74.03(19)° for 6 (Table S1). The bidentate 4′-R-terpy ligand as a whole is far from planarity. The dihedral angle between the mean planes of two coordinated pyridine rings is 10.75(1)° in molecule A and 12.80(4)° in molecule B of 1, 5.42(7)° in 3, 6.30(6)° in 4 and 15.91(5)° in 6, whereas the uncoordinated pyridyl ring and R substituent in the 4′-terpy position are inclined to the central pyridine at 53.42(1) and 22.63(6)° in molecule A and 67.09(6)° and 15.40(5)° in molecule B of 1, 61.84(8)° and 19.89(1)° in 3, 60.88(2)° and 18.94(1)° in 4, 50.02(8)° and 19.46(8) in 6, respectively.

The Re–C bond distances of 1, 3, 4, and 6 compare well to the values reported for complexes [Re(CO)3La(NN)]n+ with facially arranged carbonyl ligands.11a–f The Re(1)–N(2) bond length is significantly longer than Re(1)–N(1), which seems to be a feature typical for structures involving bidentate terpy ligands.11c–f In contrast, in the complexes incorporating tridentate-coordinated terpy ligand [ReCl(CO)2(4′-R-terpy-κ3N)], the metal-nitrogen distance to the central ring is shorter than the corresponding distances to the outer pyridyl rings.11a,b The elongation of the Re(1)–N(2) distance in [ReCl(CO)3(4′-R-terpy-κ2N)] is attributed to the steric bulk of the uncoordinated pyridyl ring. The steric hindrance induced by uncoordinated pyridyl ring is also responsible for the enlargement of the C(1)–Re(1)–N(2) angle (101.97(14)–103.2(2)°), which is the largest one between any two cis-located ligands (Table S1).

Solubility of the Re(I) complexes

Solubility of the synthesized compounds in chloroform solution in concentration 10 mg in 1 mL required for thin film preparation was tested. The influence of ligand structure on solubility of the prepared complexes was pronounced. Compound with the ligand containing 4-dimethylaminophenyl (6) substituent was practically insoluble, whereas the complexes with ligand bearing biphenyl-4-yl (1) and 4-methoxynaphthalen-1-yl (2) moieties were completely soluble. The other compounds were partially soluble, but sufficient enough for dispersing them in PVK matrix with 15 wt% concentration.

Thermal properties of Re(I) complexes

Differential scanning calorimetry (DSC) was used for investigation of the thermal behaviour of the complexes. When samples were heated, an endothermic peak due to melting was observed, except for the compound containing phenyl ring with two fluorine atoms (5). In the case of compound 5, after melting temperature (Tm) of 223 °C the crystallization temperature (Tc) of 229 °C and Tm of 282 °C were determined from DSC first heating scan. When the isotropic liquid was cooled down, in their 2nd heating DSC scan, the thermogram displays a glass transition (Tg) temperature. Thus, this compound exhibited behavior characteristic for molecular glasses.23 Fig. S2 (in ESI) shows a DSC thermogram of compound 5. For all other samples during further heating above melting maximum temperature, an exothermic peak due to thermal decomposition is seen. However, the complexes showed high Tm in the range of 294–359 °C (cf. Experimental section) and they are thermally stable enough for application in most optoelectronic devices.

Absorption spectroscopy and DFT calculations

The absorption properties in the UV-vis range of the prepared compounds 1–6 were studied in two solvents of different polarity CHCl3 (dielectric constants ε = 4.8) and MeCN (ε = 37.5). Additionally, the UV-vis spectra of selected molecules were registered in solid state, that is, in thin film (compound 1 and 2) and in blend with poly(9-vinylcarbazole) (PVK) (compound 1–5). PVK matrix was applied because it is frequently used in guest/host light emitting diodes.24 The obtained UV-vis data are collected in Table 2.
Table 2 Electronic spectral data for the complexes [ReCl(CO)3(4′-R-terpy-κ2N)]
Compound Medium λ/nm (103 × ε/dm3 mol−1 cm−1)
1 CHCl3 389.8 (6.1), 317.1 (30.4), 285.2 (24.9)
MeCN 383.9 (8.7), 315.0 (37.9), 278.6 (28.4), 191.49 (126.9)
Film 393, 320 sh
Blend 344, 330
2 CHCl3 383.9 (9), 306.1 (23.4), 240.9 (39.3)
MeCN 374.0 (42.3), 306.0 (90.7), 230.7 (184), 206.35 (235.1)
Film 398, 315 sh
Blend 389, 345, 330
3 CHCl3 398.6 (4.0), 303.0 (25.1), 267.6 (23.4)
MeCN 385.4 (3.3), 284.9 (19.6), 264.0 (18.2), 196.6 (178.2)
Blend 345, 330
4 CHCl3 401.2 (4.8), 304.2 (30.2), 267.9 (27.1)
MeCN 385.4 (5.4), 294.5 (32.5), 264.7 (30.8), 194.5 (71.8)
Blend 344, 330
5 CHCl3 399.2 (6.5), 299.9 (35.3), 261.4 (43.3)
MeCN 383.2 (9.5), 300.5 (46.5), 260.1 (60.9), 197.4 (125.8)
Blend 345, 330
6 CHCl3 429.8 (14.4), 380.2 (8.2) sh, 312.6 (14.7), 247.2 (17.1)
MeCN 419.2 (16.5), 354.1 (9.3), 308.7 (14.2), 246.2 (17.7), 192.2 (178.2)


The electronic absorption spectra of 1–6 in acetonitrile (10−5 M), plotted as a function of molar extinction coefficient versus wavelength, are shown in Fig. 2. The dominant absorption bands of 1–6 with molar extinction coefficients on the order of 1.42–23.51 × 104 M−1 cm−1 in the range 200–330 nm can be assigned to the spin allowed intraligand (IL) π → π* transitions of the diimine moiety. These bands are accompanied by the lower-energy absorptions, which are tentatively attributed to the metal-to-ligand transitions (1MLCT). Lower molar extinction coefficients (0.33–4.23 × 104 M−1 cm−1) and wide breadth of these bands are consistent with this assignment. Furthermore, as expected for 1MLCT bands, the longest wavelength absorption band of [ReCl(CO)3(4′-R-terpy-κ2N)] in chloroform is red-shifted by 5.9 nm for 1, 9.9 nm for 2, 13.2 nm for 3, 15.8 nm for 4, 16 nm for 5 and 10.6 nm for 6 with reference to more polar acetonitrile solution. From Table 2 and Fig. 2, it can also be noted that the absorption spectra of [ReCl(CO)3(4′-R-terpy-κ2N)] are clearly affected by the strong electron-donating substituents R introduced into 4′-position of the terpy ring. For complex 2, the substituent 4-methoxynaphthalen-1-yl draws electron density from the 2,2′:6′,2′′-terpyridine, facilitating the photoinduced metal-to-ligand charge transfer. It manifests itself in significant increase of the intensity of the longest wavelength absorption band of 2 compared to other studied complexes. In turn, introduction of strong electron-donating dimethylamine group in 6 results in both intensity increase and clear red-shift of the absorption in acetonitrile compared to 3, 4 and 5, while the energy of the low-energy band of [ReCl(CO)3(4′-R-terpy-κ2N)] containing chloro, bromo and fluoro substituents R remains nearly unchanged. For compounds 2 and 6, the lower-energy absorption is probably attributed to intraligand charge transfer transitions (ILCT) occurring from electron-donating substituent R to terpy moiety.25


image file: c6ra08981j-f2.tif
Fig. 2 The UV-vis absorption spectra of [ReCl(CO)3(4′-R-terpy-κ2N)] in MeCN (10−5 M).

In the UV-vis spectra of the investigated compounds in solid state, that is, as thin film on glass substrate two absorption band are seen in the similar range as were observed in chloroform solution. In the case of blend with poly(9-vinylcarbazole) (PVK) (15 wt% of compound in PVK) absorption typical for matrix is dominated. The absorption band in the lowest energy region is not clearly pronounced, except of 2.

In order to get more insight into the nature of the electronic transitions involved in the absorption processes, computational studies at DFT/B3LYP/DEF2-TZVPD/6-31+G* level were undertaken. As shown in Table S1, the predicted bond lengths and angles of 1, 3, 4 and 6 are in good agreement with the values based upon the X-ray crystal structure data, as well as the general trends observed in the experimental data are well reproduced in the calculations, providing confidence on the reliability of the chosen method to reproduce the geometries of the studied complexes.

Schematic representation of molecular orbital (MO) energy levels is presented in Fig. 3, whereas the contours of the frontier molecular orbitals of 1–6 are depicted in Fig. 4. The percentage contributions of Re, CO, terpy ring, R substituent and Cl fragments into the selected molecular orbitals of 1–6 are summarized in Tables S3–S8 (ESI Materials).


image file: c6ra08981j-f3.tif
Fig. 3 Molecular orbital energy level graph of 1–6 at the DFT/B3LYP/DEF2-TZVPD/6-31+G* level.

image file: c6ra08981j-f4.tif
Fig. 4 Frontier molecular orbitals of 1–6 computed at the DFT/B3LYP/DEF2-TZVPD/6-31+G* level.

As can be seen in Fig. 3, the HOMO level of 2 and 6 is clearly destabilized in comparison with other studied complexes, whereas the substituents R1, R3, R4 and R5 introduced into terpy ring seem to have almost the same impact on the energy of the highest molecular orbital of [ReCl(CO)3(4′-R-terpy-κ2N)]. The LUMO energy of 1–6 is only marginally influenced by the substituents attached to 2,2′:6′,2′′-terpyridine. As a result, the HOMO–LUMO gap is significantly lower in 2 (3.21 eV) and 6 (2.90 eV) compared to the other studied complexes. Notably, the HOMO–LUMO gap of 1, 3, 4 and 5 was found to lie in a narrow range of 3.57 eV to 3.59 eV. In case of complex 1, the HOMO–LUMO difference energy remains comparable to the energy gap in the complexes 3, 4 and 5, although HOMO orbitals is considerably localized on R substituent and it has to a large extend πR character.

As can be seen from Tables S3–S8, the LUMOs of 1–6 have similar distribution and they are predominantly composed of π* orbitals localized on the terpyridine moiety with percentage participation of 74.63% for 1, 73.95% for 2, 74.74% for 3, 74.66% for 4, 74.01% for 5 and 76.12% for 6. These orbitals also are contributed by 5d(Re) and π* orbitals of the R substituent, but to a much smaller extent.

For complexes 3, 4 and 5, the HOMO and HOMO−1 orbitals are constituted by 5dyz and 5dxz rhenium orbitals (∼45%) in bonding relation to the carbonyl π* orbitals (∼27%) and antibonding arrangement to chlorine p orbitals (∼22%). The HOMO−2 of these compounds is composed of 5dxy rhenium orbital (∼55%) and carbonyl π* orbitals (∼32%).

Attached to the terpy ring, biphenyl-4-yl (R1), 4-methoxynaphthalen-1-yl (R2) and 4-dimethylaminophenyl (R6) substituents raise the energy of 4′-R-terpy-localized orbitals. Consequently, the HOMOs of 1, 2 and 6 are essentially π orbitals on the 4′-R-terpy ligand and they are predominately localized on the R substituent – 57.11% for 1, 84.97% for 2 and 81.84% for 6. The orbitals 5dπ(Re), π*(CO) and 2pπ(Cl) mainly contribute to HOMO−1 and HOMO−2, while and HOMO−3 has d(Re)/π*(CO) character. For complex 1, 5dπ(Re) and π*(CO) orbitals also participate in the HOMOs of 1 – 16.17% (5dπ(Re)) and 9.46% (π*(CO)).

For complexes 3, 4, 5, π orbital of 4′-R-terpy lies below MOs with the large participation of d orbitals of rhenium. Differences in the energy order of the π orbital and d based orbitals is clearly visible amongst the value of the energy gap between the HOMO and LUMO orbitals for respective complexes (Fig. 3).

The influence of the basis sets on the energetics of MOs was tested with using larger basis sets TZVP, def2-TZVPD, TZ2P in calculations for complexes 2, 3 and 6. The most essential results obtained from this part of calculations are collected in the Tables S9 and S10 and shown in the Fig. S3 in ESI materials. In the Fig. S3, the energy diagram of frontier MOs of 2, 3 and 6 species, calculated using four different bases, is presented. Comparison of MOs energy indicates that the tested basis sets have very little effect on energy order and the nature of highest occupied and lowest unoccupied molecular orbitals. In the calculations with TZVP and def2-TZVPD basis sets, energy gap for the complexes 2 and 6 are slightly higher compared to the calculations with the def2-TZVPD/6-31+G* basis. For these two larger basis sets, energy gap increases by approximately 0.09 eV and 0.04 eV for 2 and 6, respectively. For complex 3, this gap reduces by about 0.16 eV compared to the results of calculation in the def2-TZVPD/6-31+G* basis set. Composition of the frontier molecular orbitals for complexes 2, 3 and 6 obtained from the calculations in TZVP and def2-TZVPD basis sets is also very similar to the results of calculations in the def2-TZVPD/6-31+G* basis (Table S9 in ESI materials). Also in the calculation with the TZ2P basis set and the ZORA relativistic approximation, no significant changes in the properties of HOMOs and LUMOs orbitals were observed compared to remaining bases (Table S10).

In aromatic species chemistry, split degenerated π orbitals of aromatic ring under the influence of substituent with the electronegative atom is known. Such a split increases the energy of one HOMO π orbitals as a result of the formation the antibonding combination between π orbital of aromatic system and the p orbital of electronegative atom in a substituent. Degree of such a split can be approximately correlated with the electro-donating properties of the substituent, for example the split is smaller in case of chlorine, bromine and larger in case of such substituents like –OH, –NH2 or –N(CH3)2. In relation to the presented computational results, one could suggest that a similar effect is visible in the case of π orbitals of the 4′-R-terpy ligand with the substituents R2 and R6. In these two complexes (2 and 6), HOMO orbital is characterized by a predominating contribution coming from the πR orbital. On the other hand, it cannot be excluded that the energy of this HOMO orbital is an overestimate. Perhaps its energy should be lower and be within the range of d based MOs energy.

The experimental and calculated electronic absorption spectra of 1–6 are compared in Fig. 5. The excited states calculated for 1–6 (Tables S11–S16) demonstrated that excited-state electronic structures are best described in terms of multiconfigurations, wherein a linear combination of several occupied-to-virtual MO excitations comprises a given optical absorption band. The lowest energy absorption band of 1–6 is generally attributed to a few allowed electronic transitions corresponding one-electron excitations H → L and H−2 → L for 1, H → L, H → L+1 for 2 and 6, H−3 → L, H−1 → L for 3, H−3 → L, H−1 → L and H−1 → L+1 for 4, H−1 → L for 5.


image file: c6ra08981j-f5.tif
Fig. 5 Experimental (black) and calculated (red) electronic absorption spectra of 1–6 in MeCN solution.

Interestingly, the low-wavelength range of theoretical absorption spectrum of 2 and 6 is dominated by intense transitions, mainly of ILCT (πR/terpyimage file: c6ra08981j-t1.tif) character. Especially for complex 6, the S1 excited state corresponds to the very intensive electronic transition between π orbitals of the ligand 4′-R6-terpy. For complexes 1, 3, 4 and 5 lowest excited state has basically d → image file: c6ra08981j-t2.tif character and is of MLCT type. In regard of differences in the description of the electronic structure of investigated complexes, the influence of the functionals and the basis sets on the calculation results were taken into account. For complex 6, additional test calculations were performed with the use of different functionals, i.e. hybrid-GGA, pure GGA, and long range corrected (Table S17 in ESI). For all tested functionals, the lowest electronic transitions S1 are of ILCT character, and calculated oscillator strengths of these transitions are significantly larger (0.18–0.94 a.u.) compared to oscillator strengths of S2 transitions being basically of MLCT type (0.007–0.448 a.u.). Among applied functionals, the smallest difference of the energy between the lowest experimental band and the calculated S1 transitions, is obtained in calculations with B3LYP functional (0.42 eV) providing confidence on the reliability of the chosen method to reproduce the spectral features of the studied complexes.

Inspection of the orbital percentage composition, oscillator strengths and major contributions of transitions assigned to the low-energy absorption band of 1, 3, 4 and 5 (Fig. 4 and Tables S3–S8) demonstrates that these absorption bands have a mixed character 1MLCT/1LLCT/1ILCT/1IL, but the balance between 1MLCT/1LLCT and 1ILCT/1IL is clearly shifted toward the former.

Emission spectroscopy and DFT calculations

The luminescence properties of 1–6 were studied in CHCl3 and MeCN solutions at room temperature, as well as in solid state and in 4[thin space (1/6-em)]:[thin space (1/6-em)]1 ethanol–methanol glass matrix at 77 K. Moreover, photoluminescence of selected compounds was tested in thin films (1 and 3) and blends with PVK (15% w/w of compound in matrix) (1–5) on glass substrate. The photophysical data of 1–6 are summarized in Table 3, and Fig. 6 shows the normalized emission spectra of the Re(I) complexes in MeCN and in solid state. All emission and excitation spectra of compounds 1–6 as well as decay curves are summarized in Fig. S4 in ESI.
Table 3 Summary of photoluminescent properties of the complexes [ReCl(CO)3(4′-R-terpy-κ2N)]
Compound (R) Medium λex [nm] λem [nm] Stokes shift [cm−1] ϕ [%] τ, ns (weight) χ2
a (4[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v).
image file: c6ra08981j-u1.tif CHCl3 392 664 10[thin space (1/6-em)]450 0.52 3.67 (93.89%), 32.77 (6.11%) 1.137
MeCN 366 442, 671 4698, 12[thin space (1/6-em)]194 0.47 4.46 (52.20%), 1.57 (47.80%) for λem = 442 nm; 2.94 (60.41%), 1.75 (39.59%) for λem = 671 nm 1.208; 1.124
Solid 462 592 4753 12.46 494.42 1.049
EtOH–MeOHa (77 K) 263, 306, 363 516, 550 8168   88[thin space (1/6-em)]506.6 (66.94%), 12[thin space (1/6-em)]857.7 (33.06%) 1.039
Film 330 395, 587        
390 440, 584
Blend 330 380, 580        
390 440
image file: c6ra08981j-u2.tif CHCl3 384 658 10[thin space (1/6-em)]844 0.35 4.38 1.088
MeCN 375 650 11[thin space (1/6-em)]282 0.53 4.77 (83.09%), 0.66 (16.91%) 1.047
Solid 471 574 3810 1.18 9539.0 (87.12%), 2402.9 (12.88%) 1.072
EtOH–MeOHa (77 K) 264, 316, 389 536, 568 7050   465[thin space (1/6-em)]271.6 (77.26%), 144[thin space (1/6-em)]424.4 (22.74%) 1.109
Film 330 375, 409, 580        
390 438, 467, 585
Blend 330 377, 570        
390 442, 467
image file: c6ra08981j-u3.tif CHCl3 397 669 10[thin space (1/6-em)]241 0.13 3.03 (91.19%), 17.99 (8.81%) 1.063
MeCN 376 664 11[thin space (1/6-em)]536 0.09 2.12 (86.66%), 5.61 (13.34%) 1.120
Solid 475 591 4132 21.00 493.10 1.060
EtOH–MeOHa (77 K) 264, 305, 335, 365 542 8947   5520 (66.83%), 1903 (33.17%) 1.047
Blend 330 375, 577        
390 414, 538
image file: c6ra08981j-u4.tif CHCl3 396 665 10[thin space (1/6-em)]215 0.44 2.94 (63.71%), 71.60 (36.29%) 1.134
MeCN 376 664 11[thin space (1/6-em)]536 0.15 2.21 (87.33%), 12.91 (12.67%) 1.158
Solid 479 592 3985 16.93 425.34 1.077
EtOH–MeOHa (77 K) 267, 304, 333, 365 551 9248   5110 (75.16%), 1805 (24.84%) 1.035
Blend 330 375, 579        
390 411, 436, 568
image file: c6ra08981j-u5.tif CHCl3 397 676 10[thin space (1/6-em)]396 0.40 2.72 1.092
MeCN 380 516, 666 6936, 11[thin space (1/6-em)]301 0.27 3.22 for λem = 516 nm; 2.36 for λem = 666 nm 1.109; 0.983
Solid 489 640 4825 0.96 13.43 (82.87%), 57.35 (17.13%) 1.066
EtOH–MeOHa (77 K) 268, 298, 331, 370 549 8812   3906.3 1.006
Blend 330 377, 576        
390 438, 465
image file: c6ra08981j-u6.tif CHCl3 421 636 8030 1.13 22.54 (84.36%), 7.11 (15.64%) 1.100
MeCN 452 687 7568 1.91 0.18 (96.30%), 2.48 (3.70%) 1.037
Solid 494 636 4520 0.67 281.68 (61.85%), 3081.25 (38.15%) 1.192
EtOH–MeOHa (77 K) 312, 325, 368, 444 580 5281   263[thin space (1/6-em)]293.5 (77.92%), 103[thin space (1/6-em)]380.6 (22.08%) 1.056



image file: c6ra08981j-f6.tif
Fig. 6 The normalized emission spectra of 1–6 in MeCN (a) and in solid state (b).

In solution, the complexes [ReCl(CO)3(4′-R-terpy-κ2N)] are weakly emissive. The average lifetimes and quantum yields ϕ fall in the regions 0.26–27.86 ns and 0.09–1.91%, respectively. Excitation of [ReCl(CO)3(4′-R-terpy-κ2N)] at the HOMO → LUMO absorption band gave rise to well-separated dual emission, with maxima at 442 and 671 nm for 1 in MeCN, 516 and 666 nm for 5 in MeCN, or broad structure-less emission band in the range 636–687 nm for other studied complexes. With reference to the photophysical studies of related tricarbonylrhenium(I) polypyridine systems, the emission in the range 636–687 nm seems to be consistent with the metal-ligand-to-ligand charge transfer phosphorescence (3MLLCT), while the higher energy emission peaks (442 and 516 nm) may be tentatively assigned to the 3LC (π → π*) exited state.9e,11b,12,26 The free ligands display fluorescence at 359 nm for 4′-R1-terpy and 437 nm for 4′-R5-terpy, ruling out the possibility that any of the emissions originate from ligand impurities.

As can be seen from Table 3, the emission of 6 was found to be significantly red-shifted with increasing solvent polarity. It appeared at ∼50 nm longer in wavelength as the polarity of the solvent increased from chloroform to acetonitrile. Such a behaviour is consistent with 4′-R-terpy-centred excited state with significant contributing CT character.27 In contrast, the emission maxima of the other studied complexes are influenced only marginally by changes in solvent polarity. A slight negative solvatochromic effect, reflected in blue-shifted emission with increasing solvent polarity, can be noticed for 2 and 3.

In low-temperature 77 K EtOH–MeOH (4[thin space (1/6-em)]:[thin space (1/6-em)]1 v/v) glassy medium, the emission bands of the complexes occurred at higher energy (516–580 nm) owing to the rigidochromic effect, which is responsible for raising the energy of the emissive 3MLLCT due to the lack of solvent reorganization following excitation.1,11c,28 Upon cooling to low temperature (77 K), the complexes 3, 4, and 5 exhibit broad and unstructured emission, which is a characteristic feature of MLCT emission. The microsecond excited state lifetimes support the phosphorescence assignment. The emissive bands of 1, 2 and 6 show the appearance of partial vibronic coupling features, indicating some degree of mixing between 3MLLCT and 3LC states.

All the examined complexes are also emissive in solid state. In similarity with frozen matrix, the emissions of 1–6 appear at lower energy relative to that reported for solution, consistent with the rigidochromic effect. Most notably, the introduction of biphenyl-4-yl (R1), 4-chlorophenyl (R3) and 4-bromophenyl (R4) substituents into the terpy ring leads to a significant increase in quantum yield ϕ: 12.46% for 1, 21.00% for 3 and 16.93% for 4. The stronger emission in solid than that in solution may be attributed the aggregation of the complexes due to intermolecular π–π interactions between the adjacent molecules in the crystal lattice. As a result, the triplet energy level of the charge transfer state from Re(I) to the interacting ligands is reduced and the ligands participate in the excited state in the solid state. In addition, the rotation of the free pyridine and substituent rings were also limited.

Triplet excited states of [ReCl(CO)3(4′-R-terpy-κ2N)] were also investigated computationally at DFT/UB3LYP/def2-TZVPD/6-31+G* level. The energies of phosphorescence emissions for [ReCl(CO)3(4′-R-terpy-κ2N)] were calculated from the energy difference between the ground singlet and the triplet state ΔET1−S0 as well as they were computed using TD-DFT. The calculated values of phosphorescence emission energy together with a description of character of the triplet states and experimental data are gathered in Table 4. Isosurfaces of the difference between α and β spin densities for [ReCl(CO)3(4′-R-terpy-κ2N)] at its T1 state are presented in Fig. 7 as well as the HSOMO and LSOMO involved in TD-DFT triplet excitations are available in ESI (Fig. S5). Taken into consideration that calculated emission energies are in reasonably good agreement with the experimental emission band maxima of 1–6, it is possible to conclude that luminescence has character of the phosphorescence and occurs from the low-lying triplet state (T1). Theoretical calculations predicted mixing between 3MLCT, 3LLCT, 3ILCT and 3IL excitations in the lowest excited states of 1–6. The complexity of the triplet state character results from the complicated electron structure of excited states of such complexes, in which electron excitations involve both d orbitals of metal and the π orbitals of ligand 4′-R-terpy. For complexes 1, 2 and 6, however, the lowest triplet state seems to be mainly of 3ILCT type, while 3MLCT character dominates in T1 for remaining complexes.

Table 4 Calculated phosphorescence emission energies of 1–6, compared to the experimental values recorded in acetonitrile solutiona
Compound DFT TD-DFT λexp [nm eV−1]
ΔET1−S0 (eV)/(nm) Character Major contribution (CI coefficient) E [eV] λcal [nm] Character
a ΔET1−S0-energy difference between the ground singlet and triplet states.
1 2.08/596.4 3ILCT/3IL/3MLCT/3LLCT H → L (0.6414) 1.89 656.30 3ILCT/3IL/3MLCT 671/1.85
2 1.93/642.0 3ILCT/3IL/3LLCT/3MLCT H → L (0.6094) 1.76 706.21 3ILCT/3IL 650/1.91
3 1.81/685.7 3MLCT/3LLCT/3ILCT/3IL H → L (−0.6790) 1.82 679.89 3MLCT/3LLCT/3IL 664/1.87
4 1.77/700.2 3MLCT/3LLCT/3ILCT/3IL H → L (−0.6790) 1.82 682.11 3MLCT/3LLCT/3IL 664/1.87
5 1.74/711.3 3MLCT/3LLCT/3ILCT/3IL H → L (0.6883) 1.83 678.56 3MLCT/3LLCT/3IL 666/1.86
6 1.87/662.6 3ILCT/3IL/3LLCT/3MLCT H → L (0.6769) 1.81 684.99 3ILCT/3IL 687/1.80



image file: c6ra08981j-f7.tif
Fig. 7 Isodensity surface electron spin density for the complexes 1–6 at their T1 state geometry. Blue and green colours show regions of positive and negative spin density values, respectively.

The emission ability of the compounds in the form of film and blend was examined under two excitation wavelengths, in the maximum of the absorption band, that is, 330 and 390 nm. The exemplary emission spectra of 5 in the film and blend are presented in Fig. 8. In the PL spectra of the compounds in thin film on glass two emission bands are seen under both applied λex (330 and 390 nm). In the case of compounds molecularly dispersed in a PVK matrix the lack of emission at lower energetic region under λex = 390 nm was observed. Additionally, taking onto account the fact that the emission spectrum of the host significantly overlaps with the guest's UV-vis spectrum (cf. Fig. 8a) it can be supposed that Förster Resonance Energy Transfer (FRET) takes place. Similar behavior was found in the case of the Re(I) complex reported in our former article.13b Thus, PVK matrix served both as a hole transport material and as an energy donor.


image file: c6ra08981j-f8.tif
Fig. 8 (a) UV-vis absorption spectra of compound 1, compound 1 in blend with PVK and PL spectrum of PVK; (b) PL spectra of 1 in film and blend with PVK.

Electrochemistry

Electrochemical properties of studied Re(I) complexes (except for 6 because of its insufficient solubility) were investigated in acetonitrile solution by cyclic voltammetry (CV) at scan rate of 10 mV s−1. Frontier molecular orbitals represented by the HOMOs and LUMOs which play crucial role in charge transport properties were calculated using the ferrocene (Fc) ionization potential value of −5.1 eV as the standard.29 Electrochemical data is summarized in Table 5, and Fig. 9 shows representative voltammograms of Re(I) complexes. In voltammograms of investigated compounds one or two oxidation processes were observed. First irreversible oxidation process, at around 0.75 V seems to be attributed to Re(II/I) oxidation.13b,30 As these values are relatively close to the potentials observed for one electron oxidation of the 4′-R-terpy ligands,14 however, precise determination of the oxidation potentials are unfortunately precluded. Coordination of 4′-R-terpy to the rhenium(I) can shift the 4′-R-terpy oxidation potential to lower potentials.31 Second oxidation process was observed only for complexes based on the terpyridine ligand with halogen atoms (3, 4 and 5). Quasi-reversible process (peak-to-peak separation around 0.15 V) was found for complexes 3 and 5. According to the literature, second oxidation process can be connected to losing of CO ligand and Re(III/II) oxidation, which is usually irreversible, however, higher electron density on Re center, probably caused by presence of halogen atoms, might cause stabilization of CO ligand and make second oxidation process more reversible.11b For other compounds second oxidation peak was not observed, probably due to electrochemical window of acetonitrile. The first reversible reduction process at potential around −1.75 V occurs at terpyridine ligand, similar like in our previously described Re(I) complexes,13b however, potential Ered1 of compounds reported in this work was slightly lower (ca. 0.3 V) than for complex with terpyridine ligand substituted with heterocyclic aromatic rings. Second irreversible reduction process was detected only for compounds 3 and 5. Considering the calculated HOMO and LUMO energy levels, practically the lack of the ligand structure effect was observed (Table 5). Thus, the studied compounds exhibited similar energy band gap (Eg) slightly above 2 eV. Comparing the results obtained for the compounds described herein and rhenium complexes described in our former work,13b it can be seen that the HOMO is almost the same for all compounds (around −5.7 eV). However, a higher LUMO energy level was calculated (ca. 0.3 to 0.4 eV). Consequently, Eg is also higher than for the previously described complexes (Eg about 1.8 eV).
Table 5 Electrochemical data for investigated Re(I) complexesa
Code Eox1 [V] Eox2 [V] EpcEpa [V] Eox,onset [V] Ered1 [V] EpcEpa [V] Ered2 [V] Ered,onset [V] HOMO [eV] LUMO [eV] Eg [eV]
a HOMO = −5.1 − Eox,onset, LUMO = −5.1 − Ered,onset, Eg = Eox,onsetEred,onset = HOMO − LUMO.
1 0.76 0.60 −1.76 0.07 −1.63 −5.70 −3.47 2.23
2 0.72 0.59 −1.83 0.09 −1.70 −5.69 −3.40 2.29
3 0.75 1.12 0.14 0.59 −1.75 0.08 −2.23 −1.63 −5.69 −3.47 2.22
4 0.77 1.31 0.66 −1.70 0.06 −1.50 −5.76 −3.60 2.16
5 0.77 1.21 0.16 0.60 −1.78 0.09 −2.35 −1.65 −5.70 −3.45 2.25



image file: c6ra08981j-f9.tif
Fig. 9 Cyclic voltammograms of compound 2 and 3 (scan rate 10 mV s−1, electrolyte 0.2 M Bu4NPF6 in acetonitrile).

Electroluminescence

The ability of selected compounds (1–5) for emission of light under applied voltage was investigated. All tested compounds were applied as guest dispersed in PVK matrix (15% w/w).

The appropriate alignment of the energy levels between the host matrix and the guest molecules is fulfilled for obtained compounds. Comparing the HOMO and LUMO energy of the compounds (HOMO ca. −5.7 eV, LUMO ca. 3.5 eV) and PVK (HOMO −5.8 eV and LUMO 2.2 eV) it is evident that the guest possess HOMO and LUMO energy higher and lower, respectively, than the energy levels of a matrix. OLEDs with configuration ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/PVK[thin space (1/6-em)]:[thin space (1/6-em)]compound/Al were fabricated. Additionally, devices in which compounds 2 and 3 formed an emitting single layer were obtained (ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/compound 2 or 3/Al). The morphology of the active layer, as another key factor for the OLEDs functioning, was investigated by atomic force microscopy (AFM) and the exemplary AFM images for PVK doped with compound 3 are shown in Fig. S6. The AFM images of blends show uniform and flat surfaces, indicating good film-forming properties. The surface root-mean-square roughness (RMS) of these layers were in the range of 22.8–0.7 nm. The current density–voltage (JV) characteristics of the prepared devices were registered and the exemplary JV plot is depicted in Fig. 10a, whereas the obtained data for all devices are summarized in Table 6. The turn-on voltage (Von) of all devices was about 3.5 V, except for the one based on compound 2 applied as single emitting layer. However, the OLEDs differ in value of current density (J). The significant increase in the reached J was found after utilization of PVK doped with the compounds in comparison with a neat film of 1 and 2. The lowest value of J exhibited device containing complex 3. For the devices containing compounds dispersed in matrix the electroluminescence (EL) spectra were registered and the exemplary spectra are presented in Fig. 10b together with intensity dependencies on the applied external voltage (Fig. 10c). EL spectrum becomes fully dominated by the dye emission. The EL spectra measured for all compounds match the photoluminescence spectra rather well, and we observe some variation regarding both the energetic position of the EL bands as well as their shapes. The maximum of the EL bands shifts towards longer wavelengths. In all of the compounds the emergence of EL signal was observed between 15 V and 26 V. In the case of ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/PVK[thin space (1/6-em)]:[thin space (1/6-em)]2/Al (Fig. 10c) and ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/PVK[thin space (1/6-em)]:[thin space (1/6-em)]3/Al the experiment was carried out in such a way that the external voltage was first ramped up and then decreased. It should be noted that in this article OLEDs with the simplest structure are presented. It is well known that the device performance can be improved by introducing additional charge transporting layers such as carrier injection and charge transporting layers. However, in our investigations another possibility of EL emission enhancement was applied by incorporating silver nanowires (AgNWs) into PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS layer. Thus, the devices with the following structure ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS[thin space (1/6-em)]:[thin space (1/6-em)]AgNWs/PVK[thin space (1/6-em)]:[thin space (1/6-em)]5/Al were prepared. The comparison between JV characteristics and EL spectra measured for the compound 5 with and without the nanowires incorporated to the PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS layer is displayed in Fig. 11. There are two remarkable observations that can be ascribed to the effects caused by silver nanowires. First of all, the emergence of the EL spectrum occurs significantly faster in the case of plasmonic structure: the spectrum appears at external voltage of 13–15 V as compared to around 19 V for the device devoid of silver nanowires. It indicates that the efficiency of carrier generation and injection is considerably higher when silver nanowires are incorporated into the structure. Furthermore, the observed intensities differ considerably between the two structures, and the difference is approximately an order of magnitude of comparable external voltages. This result indicates that it is possible not only to increase absorption with metallic nanoparticles,32 but also influence the EL response of the actual device.


image file: c6ra08981j-f10.tif
Fig. 10 (a) Current density–voltage characteristics of devices based on compound 2, (b) electroluminescence spectra of OLED with ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/2[thin space (1/6-em)]:[thin space (1/6-em)]PVK 15%/Al structure for increasing and decreasing external voltages, and (c) electroluminescence intensity as a function of voltage for OLED with structure ITO/PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS/2[thin space (1/6-em)]:[thin space (1/6-em)]PVK 15%/Al. The black and red curves correspond to increasing and decreasing external voltage, respectively.
Table 6 JV characteristics parameters of devices based on compounds 2–6 together with electroluminescence wavelength maximum (λEL) and thickness and surface roughness (RMS) of active layer
Code Active layer J [mA cm−2] Von [V] λEL [nm] Active layer thickness [nm] RMS [nm]
1 Film 0.05 3.5
Blend 29 3.5 630 135 11.6
2 Film 0.77 2.4
Blend 44 3.5 615 130 22.8
3 Blend 5.3 3.2 635 150 0.70
4 Blend 20 3 625 140 7.70
5 Blend 20 3.5 630 158 1.83



image file: c6ra08981j-f11.tif
Fig. 11 Comparison between (a) JV characteristics and (b) EL spectra measured as a function of external voltage for the compound 5 for devices without and with silver nanowires incorporated into PEDOT[thin space (1/6-em)]:[thin space (1/6-em)]PSS layer.

Conclusions

A comprehensive investigation on photophysical properties of new Re(I) complexes [ReCl(CO)3(4′-R-terpy-κ2N)] incorporating 2,2′:6′,2′′-terpyridine-based ligands was conducted and a decisive role of the nature of the substituent R in determining photophysical properties of [ReCl(CO)3(4′-R-terpy-κ2N)] was demonstrated. The electron-rich substituents which have lone pairs able to conjugate with the π system of the terpy core, give rise to new π → π* transitions with considerable intra-ligand, substituent-to-terpy charge-transfer (ILCT) character. The participation of ILCT transitions was supported theoretically at the DFT level. The studied complexes showed similar and promising low energy band gap, electrochemically estimated, slightly above 2 eV. They exhibited photoluminescence both in solution and solid state. Thus, Re(I) complexes, except for one based on ligand substituted with tertiary amine (compound 6), were applied for OLED fabrication with active layer consisting of compounds (15% w/w) dispersed in PVK matrix. The emission of light under applied voltage with maximum electroluminescence (EL) band in the range between 615 and 635 nm was observed. Additionally, the possibility of significant emission enhancement was demonstrated by incorporation of metallic nanowires into an OLED device.

Acknowledgements

The research was co-financed by the National Research and Development Center (NCBiR) under Grant ORGANOMET No. PBS2/A5/40/2014. The calculations have been carried out in Wroclaw Centre for Networking and Supercomputing (http://www.wcss.wroc.pl).

Notes and references

  1. M. Wrighton and D. L. Morse, J. Am. Chem. Soc., 1974, 96, 998–1003 CrossRef CAS.
  2. (a) D. J. Stufkens and A. Vlcek Jr, Coord. Chem. Rev., 1998, 177, 127–179 CrossRef CAS; (b) S. Sato, A. Sekine, Y. Ohashi, O. Ishitani, A. M. Blanco–Rodriguez, A. Vlcek Jr, T. Unno and K. Koike, Inorg. Chem., 2007, 46, 3531–3540 CrossRef CAS PubMed; (c) M. P. Coogan, V. Fernandez-Moreira, J. B. Hess, S. J. A. Pope and C. Williams, New J. Chem., 2009, 33, 1094–1099 RSC; (d) M. Chergui, Acc. Chem. Res., 2015, 48, 801–808 CrossRef CAS PubMed; (e) J. Agarwal, R. P. Johnson and G. Li, J. Phys. Chem. A, 2011, 115, 2877–2881 CrossRef CAS PubMed; (f) M. J. Li, X. Liu, Y. Q. Shi, R. J. Xie, Q. H. Wei and G. N. Chen, Dalton Trans., 2012, 41, 10612–10618 RSC; (g) J. M. Smieja and C. P. Kubiak, Inorg. Chem., 2010, 49, 9283–9289 CrossRef CAS PubMed.
  3. (a) F. Zhao, J.-x. Wang, W.-q. Liu and Y.-b. Wang, Comput. Theor. Chem., 2012, 985, 90–96 CrossRef CAS; (b) F. L. Thorp-Greenwood, J. A. Platts and M. P. Coogan, Polyhedron, 2014, 67, 505–512 CrossRef CAS; (c) S. R. Stoyanov, J. M. Villegas, A. J. Cruz, L. L. Lockyear, J. H. Reibenspies and D. P. Rillema, J. Chem. Theory Comput., 2005, 1, 95–106 CrossRef PubMed.
  4. (a) C. Gourlaouen, J. Eng, M. Otsuka, E. Gindensperger and C. Daniel, J. Chem. Theory Comput., 2015, 11, 99–110 CrossRef CAS PubMed; (b) R. Heydova, E. Gindensperger, R. Romano, J. Sykora, A. Vlcek, S. Zalis and C. Daniel, J. Phys. Chem. A, 2012, 116, 11319–11329 CrossRef CAS PubMed; (c) R. Bakova, M. Chergui, C. Daniel, A. Vlcek and S. Zalis, Coord. Chem. Rev., 2011, 255, 975–989 CrossRef CAS; (d) J. Eng, C. Gourlaouen, E. Gindensperger and C. Daniel, Acc. Chem. Res., 2015, 48, 809–817 CrossRef CAS PubMed.
  5. (a) Y. Matsubara, S. E. Hightower, J. Chen, D. C. Grills, D. E. Polyansky, J. T. Muckerman, K. Tanaka and E. Fujita, Chem. Commun., 2014, 50, 728–730 RSC; (b) S. Roy and C. P. Kubiak, Dalton Trans., 2010, 39, 10937–10943 RSC; (c) M. D. Doherty, D. C. Grills, J. T. Muckerman, D. E. Polyansky and E. Fujita, Coord. Chem. Rev., 2010, 254, 2472–2482 CrossRef CAS; (d) A. Sarto Polo, M. K. Itokazu and N. Y. Murakami Iha, Coord. Chem. Rev., 2004, 248, 1343–1361 CrossRef.
  6. (a) C. S. K. Mak, H. Ling Wong, Q. Y. Leung, W. Y. Tam, W. K. Chan and A. B. Djurisic, J. Organomet. Chem., 2009, 694, 2770–2776 CrossRef CAS; (b) A. O. T. Patrocinio and N. Y. Murakami Iha, Inorg. Chem., 2008, 47, 10851–10857 CrossRef CAS PubMed; (c) A. Wing-Tat, V. Man-Wai, H.-W. Liu and K. K.-W. Lo, Chem.–Eur. J., 2014, 20, 9633–9642 CrossRef PubMed; (d) X.-F. Wang, J. Lumin., 2013, 134, 508–514 CrossRef CAS.
  7. (a) C.-C. Ko and V. W.-W. Yam, J. Mater. Chem., 2010, 20, 2063–2070 RSC; (b) P. H.-M. Lee, C.-C. Ko, N. Zhu and V. W.-W. Yam, J. Am. Chem. Soc., 2007, 129, 6058–6059 CrossRef CAS PubMed; (c) O. S. Wenger, L. M. Henling, M. W. Day, J. R. Winkler and H. B. Gray, Inorg. Chem., 2004, 43, 2043–2048 CrossRef CAS PubMed; (d) S.-S. Sun and A. J. Lees, Organometallics, 2002, 21, 39–49 CrossRef CAS.
  8. (a) M. R. Goncalves and K. P. M. Frin, Polyhedron, 2015, 97, 112–117 CrossRef CAS; (b) W.-K. Chu, C.-C. Ko, K.-C. Chan, S.-M. Yiu, F.-L. Wong, C.-S. Lee and V. A. L. Roy, Chem. Mater., 2014, 26, 2544–2550 CrossRef CAS; (c) H. Yersin, A. F. Rausch, R. Czerwieniec, T. Hofbeck and T. Fischer, Coord. Chem. Rev., 2011, 255, 2622–2652 CrossRef CAS; (d) X. Li, D. Zhang, G. Lu, G. Xiao, H. Chi, Y. Dong, Z. Zhang and Z. Hu, J. Photochem. Photobiol., A, 2012, 241, 1–7 CrossRef CAS.
  9. (a) A. Zarkadoulas, E. Koutsouri, C. Kefalidi and C. A. Mitsopoulou, Coord. Chem. Rev., 2015, 304–305, 55–72 CrossRef CAS; (b) H. Takeda and O. Ishitani, Coord. Chem. Rev., 2010, 254, 346–354 CrossRef CAS; (c) W. Belliston-Bittner, A. R. Dunn, Y. H. L. Nguyen, D. J. Stuehr, J. R. Winkler and H. B. Gray, J. Am. Chem. Soc., 2005, 127, 15907–15915 CrossRef CAS PubMed; (d) C. D. Windle and R. N. Perutz, Coord. Chem. Rev., 2012, 256, 2562–2570 CrossRef CAS; (e) A. A. Abdel-Shafi, J. L. Bourdelande and S. S. Ali, Dalton Trans., 2007, 2510–2516 RSC.
  10. (a) G. Santoro, T. Zlateva, A. Ruggi, L. Quaroni and F. Zobi, Dalton Trans., 2015, 44, 6999–7008 RSC; (b) J. A. Czaplewska, F. Theil, E. Altuntas, T. Niksch, M. Freesmeyer, B. Happ, D. Pretzel, H. Schafer, M. Obata, S. Yano, U. S. Schubert and M. Gottschaldt, Eur. J. Inorg. Chem., 2014, 6290–6297 CrossRef CAS; (c) M. Kaplanis, G. Stamatakis, V. D. Papakonstantinou, M. Paravatou-Petsotas, C. A. Demopoulos and C. A. Mitsopoulou, J. Inorg. Biochem., 2014, 135, 1–9 CrossRef CAS PubMed; (d) K. K.-W. Lo, M.-W. Louie and K. Y. Hang, Coord. Chem. Rev., 2010, 254, 2603–2622 CrossRef CAS; (e) G. Balakrishnan, T. Rajendran, K. S. Murugan, M. S. Kumar, V. K. Sivasubramanian, M. Ganesan, A. Mahesh, T. Thirunalasundari and S. Rajagopal, Inorg. Chim. Acta, 2015, 434, 51–59 CrossRef CAS.
  11. (a) B. Laramee-Milette, C. Lachance-Brais and G. S. Hanan, Dalton Trans., 2015, 44, 41–45 RSC; (b) B. A. Frenzel, J. E. Schumaker, D. R. Black and S. E. Hightower, Dalton Trans., 2013, 42, 12440–12451 RSC; (c) D. Wang, Q.-L. Xu, S. Zhang, H.-Y. Li, C.-C. Wang, T.-Y. Li, Y.-M. Jing, W. Huang, Y.-X. Zheng and G. Accorsi, Dalton Trans., 2013, 42, 2716–2723 RSC; (d) A. J. Amoroso, A. Banu, M. P. Coogan, P. G. Edwards, G. Hossain and K. M. Abdul Malik, Dalton Trans., 2010, 39, 6993–7003 RSC; (e) M. P. Coogan, V. Fernandez-Moreira, B. M. Kariuki, S. J. A. Pope and F. L. Thorp-Greenwood, Angew. Chem., Int. Ed., 2009, 48, 4965–4968 CrossRef CAS PubMed; (f) V. Fernandez-Moreira, F. L. Thorp-Greenwood, R. J. Arthur, B. M. Kariuki, R. L. Jenkins and M. P. Coogan, Dalton Trans., 2010, 39, 7493–7503 RSC.
  12. A. Juris, S. Campagna, I. Bidd, J.-M. Lehn and R. Ziessel, Inorg. Chem., 1988, 27, 4007–4011 CrossRef CAS.
  13. (a) G. Velmurugan and P. Venuvanalingam, Dalton Trans., 2015, 44, 8529–8542 RSC; (b) T. Klemens, A. Świtlicka-Olszewska, B. Machura, M. Grucela, E. Schab-Balcerzak, K. Smolarek, S. Mackowski, A. Szlapa, S. Kula, S. Krompiec, P. Lodowski and A. Chrobok, Dalton Trans., 2016, 45, 1746–1762 RSC.
  14. A. Maroń, A. Szlapa, T. Klemens, S. Kula, B. Machura, S. Krompiec, J. G. Małecki, A. Świtlicka-Olszewska, K. Erfurt and A. Chrobok, Org. Biomol. Chem., 2016, 14, 3793–3808 Search PubMed.
  15. Y. Sun, Y. Yin, B. Mayers, T. Herricks and Y. Xia, Chem. Mater., 2002, 14, 4736–4745 CrossRef CAS.
  16. D. Kowalska, B. Krajnik, M. Olejnik, M. Twardowska, N. Czechowski, E. Hofmann and S. Mackowski, Sci. World J., 2013, 2013, 670412 Search PubMed.
  17. Oxford Diffraction Ltd., CrysAlis RED, Version 1.171.35.11, 2011 Search PubMed.
  18. G. M. Sheldrick, Acta Crystallogr., Sect. A: Found. Crystallogr., 2008, 64, 112–122 CrossRef CAS PubMed.
  19. 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, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision D.01, Gaussian, Inc., Wallingford CT, 2013 Search PubMed.
  20. (a) D. Andrae, U. Haeussermann, M. Dolg, H. Stoll and H. Preuss, Theor. Chim. Acta, 1990, 77, 123–141 CrossRef CAS; (b) B. Metz, H. Stoll and M. Dolg, J. Chem. Phys., 2000, 113, 2563–2569 CrossRef CAS; (c) K. A. Peterson, D. Figgen, E. Goll, H. Stoll and M. Dolg, J. Chem. Phys., 2003, 119, 11113–11123 CrossRef CAS; (d) T. Leininger, A. Nicklass, W. Kuechle, H. Stoll, M. Dolg and A. Bergner, Chem. Phys. Lett., 1996, 255, 274–280 CrossRef CAS; (e) M. Kaupp, P. V. Schleyer, H. Stoll and H. Preuss, J. Chem. Phys., 1991, 94, 1360–1366 CrossRef CAS.
  21. (a) B. Mennucci and J. Tomasi, J. Chem. Phys., 1997, 106, 5151–5158 CrossRef CAS; (b) M. T. Cances, B. Mennucci and J. Tomasi, J. Chem. Phys., 1997, 107, 3032–3041 CrossRef; (c) M. Cossi, V. Barone, B. Mennucci and J. Tomasi, Chem. Phys. Lett., 1998, 286, 253–260 CrossRef CAS.
  22. K. Nakamoto, Infrared and Raman Spectra of Inorganic and Coordination Compounds, Wiley-Interscience, New York, 6th edn, 2009 Search PubMed.
  23. A. Plante, S. Palato, O. Lebel and A. Soldera, J. Mater. Chem. C, 2013, 1, 1037–1042 RSC.
  24. B. Luszczynska, E. Dobruchowska, I. Glowacki, J. Ulanski, F. Jaiser, X. Yang, D. Neher and A. Danel, J. App. Phys., 2006, 99, 024505 CrossRef.
  25. C. B. Larsen, H. Salm, Ch A. Clark, A. B. S. Elliott, M. G. Fraser, R. Horvath, N. T. Lucas, X.-Z. Sun, M. W. George and K. C. Gordon, Inorg. Chem., 2014, 53, 1339–1354 CrossRef CAS PubMed.
  26. (a) A. Vlcek, Top. Organomet. Chem., 2010, 29, 73–114 CrossRef CAS; (b) P. J. Wright, M. G. Affleck, S. Muzzioli, B. W. Skelton, P. Raiteri, D. S. Silvester, S. Stagni and M. Massi, Organometallics, 2013, 32, 3728–3737 CrossRef CAS.
  27. (a) T. Mutai, J.-D. Cheon, G. Tsuchiya and K. Araki, J. Chem. Soc., Perkin Trans. 2, 2002, 862–865 RSC; (b) W. Goodall, K. Wild, K. J. Arm and J. A. G. Williams, J. Chem. Soc., Perkin Trans. 2, 2002, 1669–1681 RSC; (c) J.-D. Cheon, T. Mutai and K. Araki, Org. Biomol. Chem., 2007, 5, 2762–2766 RSC.
  28. A. J. Lees, Chem. Rev., 1987, 87, 711–743 CrossRef CAS.
  29. R. Rybakiewicz, I. Tszydel, J. Zapala, L. Skorka, D. Wamil, D. Djurado, J. Pécaut, J. Ulanski, M. Zagorska and A. Pron, RSC Adv., 2014, 4, 14089–14100 RSC.
  30. (a) K. Koike, J. Tanabe, S. Toyama, H. Tsubaki, K. Sakamoto, J. R. Westwell, F. P. A. Johnson, H. Hori, H. Saitoh and O. Ishitani, Inorg. Chem., 2000, 39, 2777–2783 CrossRef CAS PubMed; (b) H. Hori, K. Koike, M. Ishizuka, K. Takeuchi, T. Ibusuki and O. Ishitani, J. Organomet. Chem., 1997, 530, 169–176 CrossRef CAS.
  31. L. D. Ramos, R. N. Sampaio, F. F. de Assis, K. T. de Oliveira, P. Homem-de-Mello, A. O. T. Patrocinio and K. P. M. Frin, Dalton Trans., 2016 10.1039/c6dt01112h.
  32. K. Smolarek, B. Ebenhoch, N. Czechowski, A. Prymaczek, M. Twardowska, I. D. W. Samuel and S. Mackowski, Appl. Phys. Lett., 2013, 103, 203302 CrossRef.

Footnote

Electronic supplementary information (ESI) available: Short intra- and intermolecular contacts, MO composition of selected compounds, calculated absorption and emission data, excitation and emission spectra with PL lifetime curves, DSC thermograms of 2 and AFM images of the blend PVK with 3. CCDC 1471443–1471446. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c6ra08981j

This journal is © The Royal Society of Chemistry 2016