DOI:
10.1039/C5RA14570H
(Paper)
RSC Adv., 2015,
5, 73926-73934
Structurally simple phenanthroimidazole-based bipolar hosts for high-performance green and red electroluminescent devices†
Received
23rd July 2015
, Accepted 24th August 2015
First published on 25th August 2015
Abstract
With the aim of developing bipolar host materials, m-DPPI with two phenanthroimidazole groups meta-linked on a benzene ring and its donor–acceptor-type analogue, m-PPPI, with one of the phenanthroimidazole moieties replaced by an electron-withdrawing imidazophenanthroline group have been synthesized. The meta-linkage endows these compounds with relatively high triplet energies of ca. 2.6 eV for hosting both green and red phosphorescent emitters. The single-carrier devices indicate that both the neat and phosphor-doped films of these compounds have bipolar transporting properties, with relatively higher current densities for the films of m-DPPI. Efficient green (maximum efficiencies: 58.5 cd A−1, 67.1 lm W−1, external quantum efficiency (ηext) of 15.5%) and red (maximum efficiencies: 19.7 cd A−1, 26.9 lm W−1, ηext of 11.6%) PhOLEDs have been achieved using m-DPPI as the host. The performance of m-PPPI-based devices is also high, although a bit lower than that of m-DPPI-based devices. Moreover, all the devices exhibit low current-efficiency and ηext roll-off. The simple structures, easy syntheses and high device performance of these compounds are attractive regarding practical applications.
Introduction
Due to the ability of harvesting both electro-generated singlet and triplet excitons to achieve a 100% theoretical internal quantum efficiency, phosphorescent organic light-emitting diodes (PhOLEDs) exhibit great potential in both flat-panel displays and solid-state lighting.1,2 In PhOLEDs, a doped emitting layer (EML) with a fluorescent host as the matrix for dispersing phosphorescent emitters is usually adopted in order to avoid aggregation-caused luminescence quenching and triplet–triplet annihilation of the phosphors.3 As a result, the device performance is highly dependent on the properties of host materials. Recently, bipolar transporting hosts able to transport both electrons and holes have drawn considerable attention because they can provide relatively balanced hole and electron fluxes and a large hole–electron recombination zone in EML for improving device performance.4 One effective strategy for constructing bipolar hosts is incorporating electron-donating and electron-accepting groups into a molecule, as these two components could accept/transport holes and electrons, respectively.4a,5–8 However, the bipolar host is not limited to the donor–acceptor systems, as a few materials without obvious push–pull structures have also been proved to possess bipolar transporting ability.4a Besides the transporting requirements, another generally required character for host materials in PhOLEDs is a relatively large triplet energy for efficient triplet exciton transfer from host to guest and for preventing the reverse transfer.4a,9,10 In addition, a large triplet energy may make the host suitable for phosphorescent emitters with different emission colours.11 This is of great importance for full-colour/white OLEDs as it can reduce the varieties of applied materials and thus improve the process stability and reduce the production cost.11
Recently, phenanthroimidazole derivatives have attracted increasing attention as blue emitters12,13 and host materials.14,15 For this class of hosts, an intriguing feature is that they usually exhibit bipolar transporting character. For example, Yang and coworkers reported a series of phenanthroimidazole/carbazole hybrid bipolar hosts, and the green-emitting devices based on them showed a maximum current efficiency (ηc,max) of 77.6 cd A−1, a maximum power efficiency (ηp,max) of 80.3 lm W−1, and a maximum external quantum efficiency (ηext,max) of 21%;14a,b we also reported a series of bipolar hosts based on metal complexes and organic derivatives of phenanthroimidazole, which can realize high-performance red, green and blue OLEDs.15 These progresses suggest the great potential of phenanthroimidazole derivatives as candidates for commercial host materials. However, the research on this class of hosts is still limited. Therefore, the further development is important and attractive.
In this study, we have synthesized two phenanthroimidazole derivatives, m-DPPI and m-PPPI, with the aim of developing bipolar hosts. m-DPPI with two phenanthroimidazole moieties meta-linked on a benzene ring has no obvious push–pull character, while by replacing one phenanthroimidazole moiety with more electron-withdrawing imidazophenanthroline,16 m-PPPI possesses a donor–acceptor structure. The meta-linkage of two bulky substituents is beneficial to the realization of high triplet energies.7a,13j,17 The carrier transporting properties and OLED performance of these compounds have been evaluated. In addition, the photophysical, electrochemical and thermal properties of them have also been studied.
Results and discussion
Synthesis and crystal structure
As shown in Scheme 1, both m-DPPI and m-PPPI were synthesized by one-pot reaction through refluxing the mixture of phenanthrene-9,10-dione/1,10-phenanthroline-5,6-dione, aniline, ammonium acetate and corresponding aromatic aldehyde in acetate acid. The crude products separated by simple workup can be easily purified by vacuum sublimation to afford pure samples suitable for electronic devices. Both compounds were synthesized in good yields (56% for m-DPPI and 64% for m-PPPI). The simple preparation method is beneficial to the reduction of material cost, which is fairly important for materials' commercial application. The structures of both compounds were fully characterized by NMR spectroscopy, mass spectra and elemental analysis.
 |
| Scheme 1 Synthetic procedure of m-DPPI and m-PPPI. | |
To understand the molecular confirmations and reveal the intermolecular packing modes, the growth of single crystals was attempted, and the slab-like crystals of m-DPPI has been successfully obtained by slow evaporation of its saturated CH2Cl2 solution. The crystal belongs to orthorhombic centrosymmetric space group Pbcn with Z = 4, and the crystal structure is shown in Fig. 1. In the molecule, the phenanthroimidazole group has a good planarity and shows a dihedral angle of 34.9° with the central benzene ring. Adjacent molecules are stacked via two kinds of C–H⋯π interactions, and there are no C–H⋯N hydrogen-bonding and π⋯π interactions in the crystal.
 |
| Fig. 1 (a) Molecular structures of m-DPPI from single-crystal X-ray diffraction. Element (colour): carbon (gray), nitrogen (blue). Hydrogen atoms are omitted for clarity. Atomic displacement ellipsoids are drawn at 50% probability; (b) intermolecular C–H⋯π interactions: C–H18⋯π (H⋯π-ring centroid = 2.94 Å, H⋯π-ring plane = 2.81 Å, C⋯π-ring centroid = 3.77 Å, C–H⋯π-ring centroid = 149.5°) and C–H22⋯π (H⋯π-ring centroid = 2.85 Å, H⋯π-ring plane = 2.77 Å, C⋯π-ring centroid = 3.62 Å, C–H⋯π-ring centroid = 141.0°); (c) molecular packing structure of m-DPPI. | |
Thermal properties
The thermal properties of the compounds were characterized by thermogravimetric analysis (TGA) and differential scanning calorimetry (DSC) under a nitrogen atmosphere at a heating rate of 10 °C min−1 (Fig. S1 and S2†). m-DPPI and m-PPPI exhibit high decomposition temperatures (Td, corresponding to 5% weight loss) of 477 and 486 °C, respectively, which ensure sufficient stability for vacuum sublimation and device fabrication processes. The melting points (Tm) determined by DSC measurements are 358 °C for m-DPPI and 369 °C for m-PPPI. Compared with the 1,2-diphenyl-1H-phenanthro[9,10-d]imidazole (PPI) and its methyl-substituted derivatives, which possess the Td values of 336–343 °C and Tm values of 194–211 °C,12a m-DPPI and m-PPPI show much higher thermal stability partly due to their significantly increased molecular weight.
Photophysical properties
In CH2Cl2 (Fig. 2a), these compounds exhibit similar absorption-spectra profiles and close absorption edges (375 nm for m-DPPI and 379 nm for m-PPPI). m-DPPI displays ultraviolet emission, showing a structured emission spectrum with the maximum (λem) located at 388 nm. From m-DPPI to m-PPPI (λem = 402 nm), the emission spectrum is slightly red shifted by ca. 14 nm and becomes less structured and broader. In the thin-film state (Fig. 2b), the λem of m-DPPI is located at 384 nm, close to that of the CH2Cl2 solution. As for m-PPPI, the thin-film emission spectrum (λem = 434 nm) is obviously red-shifted and broader than the solution spectrum, suggesting strong intermolecular interactions in the solid state.18 The phosphorescence spectra of these compounds measured in 2-MeTHF at 77 K are very similar (Fig. 2b). The triplet energies (ET) estimated from the highest-energy vibronic sub-band of the phosphorescence spectra17,19 are 2.57 eV for both compounds, which is close to the ET value of classical phosphorescent host CBP (4,4′-bis(N-carbazolyl)-1,1′-biphenyl, 2.56 eV)20 and sufficiently high to host green and red phosphorescent dyes.4b We note that the ET value of these compounds is also quite close to that of PPI (ET = 2.61 eV), indicating that the meta-linkage is effective for retaining the triplet energy (Fig. S3†).
 |
| Fig. 2 (a) UV-vis absorption and emission spectra of m-DPPI and m-PPPI in CH2Cl2 (10−5 M); (b) emission spectra of m-DPPI and m-PPPI in thin-film state and their phosphorescence spectra in 2-MeTHF (10−5 M) at 77 K. | |
Electrochemical properties and theoretical calculations
Cyclic voltammetry (CV) measurements were carried out to estimate the HOMO and LUMO energy levels (Fig. 3). The oxidation onset potentials of m-DPPI and m-PPPI in CH2Cl2 were determined to be 0.80 and 0.82 V (vs. ferrocene/ferrocenium), respectively. Correspondingly, the HOMO energy levels were estimated to be −5.90 eV for m-DPPI and −5.92 eV for m-PPPI by assuming that the HOMO of ferrocene lies 5.1 eV (ref. 21) below vacuum level. Because there is no clear reduction process for these compounds within the electrochemical window of CH2Cl2 or even THF and DMF, the LUMO energy levels were deduced from HOMO energies and the optical band gaps determined from the edge of absorption spectra.12a,g,h,j,14b,15a,22 The deduced LUMO energies are −2.60 and −2.65 eV for m-DPPI and m-PPPI, respectively.
 |
| Fig. 3 Cyclic voltammograms of m-DPPI and m-PPPI in CH2Cl2. | |
To gain further insight into the electronic structures of these compounds, density functional theory (DFT) calculations were performed at B3LYP/6-31G(d) level (Fig. 4). The HOMO of m-DPPI is distributed on the whole molecule except the N-bonded phenyl groups, while the LUMO is distributed on the C6H4 linker and a part of each phenanthroimidazole moiety. For m-PPPI, due to the introduction of electron-withdrawing imidazophenanthroline moiety, HOMO and LUMO show clear spatial separation. The HOMO is mainly located on the phenanthroimidazole donor, while the LUMO is mainly dispersed on the imidazophenanthroline acceptor and the C6H4 linker. From m-DPPI to m-PPPI, replacing one phenanthroimidazole group with the imidazophenanthroline group slightly lower the HOMO and LUMO energies (Fig. 4).
 |
| Fig. 4 DFT (B3LYP/6-31G(d)) calculated molecular orbitals and energy levels of m-DPPI and m-PPPI. | |
Single-carrier devices
To evaluate the carrier injection/transporting properties, we fabricated single-carrier devices with the configurations of [ITO/MoO3 (10 nm)/m-DPPI or m-PPPI (60 nm)/MoO3 (10 nm)/Al (100 nm)] for hole-only devices and [ITO/TPBI (10 nm)/m-DPPI or m-PPPI (60 nm)/TPBI (10 nm)/LiF (1 nm)/Al (100 nm)] for electron-only devices (Fig. 5). MoO3 and TPBI (1,3,5-tris(N-phenylbenzimidazol-2-yl)benzene) are able to prevent electron injection from the cathode (Al) and hole injection from anode (ITO, indium-tin oxide), respectively, which guarantees the single-carrier injection characteristic of these devices.23 Since dopant molecules may influence carrier transporting properties, we also fabricated single-carrier devices based on the phosphor-doped films of these compounds (doping 8 wt% of typical green phosphor Ir(ppy)3 or red phosphor Ir(MDQ)2(acac)24), with the thickness of each layer unchanged for comparison (Fig. 5). The devices of neat m-DPPI show good conductivity for holes and electrons as indicated by the availability of large current densities (Fig. 5a and b), suggesting that m-DPPI has a bipolar transporting property.22 Although the single-carrier devices of m-PPPI can also conduct holes and electrons, their current densities at the same driving voltage is significantly lower (Fig. 5c and d). Considering the very close HOMO and LUMO energy levels of m-DPPI and m-PPPI determined by electrochemical data and optical band gaps, the hole injection from MoO3 and electron injection from TPBI to these compounds should have almost the same energy barrier. Therefore, the much lager current densities of m-DPPI-based devices imply a better carrier transporting property of this compound. The relatively poor transporting ability of m-PPPI could be related to the specific molecular arrangement in the thin-film state and the more localized HOMO and LUMO distributions (Fig. 4) that may reduce the possibility of intermolecular orbital overlap and therefore lower the charge-transfer integral.4a,25 Our results further confirm that the non-donor-acceptor systems could also be good bipolar transporting materials. For hole-only devices, the current density at a given driving voltage is decreased when Ir(ppy)3 or Ir(MDQ)2(acac) is doped into the film of m-DPPI or m-PPPI. In contrast, the current densities of electron-only devices are increased by doping the Ir-complexes (Fig. 5). Similar influence of phosphor-doping on current densities has been observed before, and the decreased hole current can be explained by the hole-trapping behaviour of the Ir-complexes and the increased electron current should be attributed to additional electron transporting channels formed by the dopants.26 The devices of doped m-DPPI show significantly higher current densities than corresponding devices of doped m-PPPI, implying more opportunities for electron–hole recombination in doped m-DPPI layer. This feature of doped m-DPPI is beneficial to the enhancement of OLED performance when it is used as the emitting layer in electroluminescent (EL) devices.
 |
| Fig. 5 Current density–voltage characteristics of hole-only and electron-only devices. | |
Phosphorescent OLEDs
To evaluate the practical utility of m-DPPI and m-PPPI as host materials, we initially examined the performance of Ir(ppy)3-based green phosphorescent device with the structure of [ITO/MoO3 (10 nm)/NPB (50 nm)/TCTA (5 nm)/m-DPPI (or m-PPPI): 8% Ir(ppy)3 (25 nm)/TPBI (30 nm)/LiF (1 nm)/Al (100 nm)] (devices G1 for m-DPPI and G2 for m-DPPI). NPB (N,N′-di-(naphthalen-1-yl)-N,N′-diphenylbenzidine) and TPBI were used as the hole- and electron-transporting layers, respectively; TCTA (4,4′,4′′-tri(N-carbazolyl)triphenylamine) was used as exciton-blocking layer; MoO3 and LiF served as hole- and electron-injecting layers, respectively. Fig. 6 depicts the energy level diagram of used materials and molecular structures of the phosphorescent dopants in the devices. The HOMO energies of m-DPPI and m-PPPI match well with that of TCTA and the LUMO energies match well with that of TPBI, which ensures the efficient carrier injection to the EML. The device performance is summarized in Table 1. The efficiency versus current density curves and current density–voltage–luminance (J–V–L) characteristics for the devices are shown in Fig. 7, S4 and S5.† Both devices show a low turn-on voltage of 2.7 V. Device G1 using m-DPPI as the host achieves maximum luminance (Lmax) of 48
016 cd m−2, ηc,max of 58.5 cd A−1, ηp,max of 67.1 lm W−1 and ηext,max of 15.5%. In comparison, device G2 hosted by m-PPPI exhibits slightly lower efficiencies with the ηc,max of 50.7 cd A−1, ηp,max of 58.9 lm W−1 and ηext,max of 13.5%. Notably, these devices show low ηc and ηext roll-off. At a brightness of 100 cd m−2, the efficiencies are 57.9 cd A−1 and 15.4% for device G1 and 47.5 cd A−1 and 12.8% for device G2, without significant decay compared with the maximum values. When the brightness reaches 1000 cd m−2, G1 and G2 still exhibit high ηc values of 54.6 and 43.2 cd A−1 and ηext values of 14.6 and 11.5%, respectively. The low efficiency roll-off is of great importance regarding the practical applications in displays and lighting.3g,5d,27
 |
| Fig. 6 Energy level diagram of the materials used in EL devices and the chemical structures of the phosphorescent emitters. The HOMO energy levels of m-DPPI and m-PPPI are calculated from the electrochemical data and the LUMO levels of both compounds are deduced from the HOMO levels and the optical band gaps. | |
Table 1 Electroluminescent properties of the devicesa
|
Dopant |
Von/V |
Lmax/cd m−2 |
ηcb/cd A−1 |
ηpb/lm W−1 |
ηextb/% |
CIE (x, y)c |
Abbreviation: Von: turn-on voltage; Lmax: maximum luminance; ηc: current efficiency; ηp: power efficiency; ηext: external quantum efficiency. Values given in the order of maximum, value at 100 cd m−2 and value at 1000 cd m−2. Measured at 100 cd m−2. |
G1 |
Ir(ppy)3 |
2.7 |
48 016 |
58.5, 57.9, 54.6 |
67.1, 60.4, 41.5 |
15.5, 15.4, 14.6 |
0.32, 0.61 |
G2 |
Ir(ppy)3 |
2.7 |
48 467 |
50.7, 47.5, 43.2 |
58.9, 44.0, 29.4 |
13.5, 12.8, 11.5 |
0.30, 0.63 |
R1 |
Ir(MDQ)2(acac) |
2.3 |
42 590 |
19.7, 17.1, 16.6 |
26.9, 17.8, 12.0 |
11.6, 11.2, 10.5 |
0.64, 0.36 |
R2 |
Ir(MDQ)2(acac) |
2.3 |
37 683 |
14.6, 14.3, 13.2 |
16.8, 11.4, 7.02 |
12.8, 12.6, 10.8 |
0.64, 0.36 |
 |
| Fig. 7 Current efficiency (a) and power efficiency (b) versus luminance curves, and EL spectra (c) of the devices. | |
Next, the red-emitting OLEDs using 8 wt% Ir(MDQ)2(acac) as the dopant with the configuration same as that of the green device were fabricated. The turn-on voltages of m-DPPI-hosted device, R1, and m-PPPI-hosted device, R2, were further reduced to 2.3 V (Table 1). The Lmax and maximum efficiencies are 42
590 cd m−2, 19.7 cd A−1, 26.9 lm W−1 and 11.6% for R1 (Fig. 7, S4 and S5†). From R1 to R2, the ηc,max and ηp,max are decreased to 14.6 cd A−1 and 16.8 lm W−1, and the ηext,max is only slightly influenced. It is worth noting that the power efficiency of device R1 is comparable with the best values reported for red-emitting PhOLEDs.3f,28 R1 and R2 also exhibit low ηc and ηext roll-off (Table 1). For example, R1 has the ηc values of 17.1 and 16.6 cd A−1 and ηext values of 11.2 and 10.5% at 100 and 1000 cd m−2, respectively.
All the devices display pure green or red emission from the dopant and there isn't residual emission from the host, indicating the complete energy transfer from host to dopant (Fig. 7c). Given the same triplet energy and close HOMO and LUMO energies of m-DPPI and m-PPPI, the relatively higher performance of m-DPPI-based devices should be partly ascribed to a better carrier transporting ability of the doped m-DPPI film as suggested by the single-carrier devices. The bipolar transporting property of the doped emitting layer results in a large recombination zone for electrons and holes. This can reduce the triplet–triplet annihilation and triplet-polaron quenching and thus reduce the efficiency roll-off at high luminance.26,29
Conclusion
In summary, two phenanthroimidazole-based host materials, m-DPPI and m-PPPI, have been designed and synthesized. These compounds exhibit bipolar transporting ability, regardless of whether they have donor–acceptor structures. Moreover, they possess high triplet energies of ca. 2.6 eV, which enable them to act as hosts for both the green and red phosphors. By using them as host materials, efficient green and red PhOLEDs have been achieved. The performance of m-DPPI-based devices is higher, partly due to the better transporting ability of doped m-DPPI layer. In addition, all the devices show low efficiency roll-off with slightly changed ηc and ηext values within the brightness range of 100–1000 cd m−2. The facile syntheses, the ability to host multicolour dopants and the good device performance of these compounds are all valuable merits with respect to commercial applications. Our studies further confirm that phenanthroimidazole is an excellent building block for bipolar hosts and both the donor–acceptor and non-donor–acceptor systems based on it could be good bipolar transporting materials.
Experimental
General information
MoO3, NPB, TCTA, TPBI, LiF, n-Bu4NPF6 and the starting materials for organic syntheses were obtained from commercial sources. All of the organic materials used for device fabrication were purified by vacuum sublimation prior to use. The starting materials for syntheses were used as received. All synthetic reactions were carried out under a nitrogen atmosphere by using Schlenk techniques. 1H NMR spectra were measured on Bruker Avance 500 MHz spectrometer with tetramethylsilane (TMS) as the internal standard. Mass spectra were recorded on a Shimadzu AXIMA-CFR MALDI-TOF mass spectrometer. Elemental analyses were performed on a flash EA 1112 analyser. UV-vis absorption spectra were recorded by a Shimadzu UV-2550 spectrophotometer. The emission spectra were recorded by a Shimadzu RF-5301 PC spectrometer. DSC measurements were performed on a NETZSCH DSC204 instrument at a heating rate of 10 °C min−1 under a nitrogen atmosphere. TGA measurements were performed on a TA Q500 thermogravimeter by measuring their weight loss while heating at a rate of 10 °C min−1 under nitrogen. Electrochemical measurements were performed with a BAS 100W Bioanalytical electrochemical work station with a scan rate of 100 mV s−1. A three-electrode cell configuration was employed for the measurement: a platinum disk as the working electrode, a platinum wire as the counter electrode, and an Ag/Ag+ electrode as the reference electrode. The oxidation potentials were measured in dichloromethane solution containing 0.1 M of n-Bu4NPF6 as the supporting electrolyte at a scan rate of 100 mV s−1.
Single-crystal structure
Diffraction data were collected on a Rigaku RAXIS-PRID diffractometer using the ω-scan mode with graphite-monochromator Mo Kα radiation. The structure was solved with direct methods using the SHELXTL programs and refined with full-matrix least-squares on F2.30 Non-hydrogen atoms were refined anisotropically. The positions of hydrogen atoms were calculated and refined isotropically. ESI.†
Theoretical calculations
The ground state geometries were optimized by the density functional theory (DFT)31 method with the Becke three-parameter hybrid exchange and the Lee–Yang–Parr correlation functional32 (B3LYP) and 6-31G(d) basis set using the Gaussian 03 software package.33 The molecular structure of m-DPPI as determined by X-ray crystallography was used as the input for optimizing ground-state geometries; by changing two carbon atoms with nitrogen atoms, the geometry of m-PPPI was optimized.
Fabrication and characterization of OLEDs and single-carrier devices
Before device fabrication, the ITO-coated glass substrates were pre-cleaned by sonication successively in a detergent solution, acetone, methanol, and deionized water and then treated with plasma for 2 min. The devices were prepared in vacuum at a pressure of 5 × 10−6 Torr. Organic layers, MoO3 and LiF were deposited onto the substrate at a rate of 0.3–0.5 Å s−1. The Al electrode was deposited at a rate of 10 Å s−1. Thicknesses of the material layer were controlled using a quartz crystal thickness monitor. The electrical characteristics of the OLEDs and single-carrier devices were measured with a Keithley 2400 source meter. The EL spectra and luminance of the OLEDs were obtained using a PR650 spectrometer. All measurements were carried out on the devices without encapsulations in ambient atmosphere under dark.
Synthesis
1,3-Bis(1-phenyl-1H-phenanthro[9,10-d]imidazol-2-yl)benzene(m-DPPI). A mixture of phenanthrene-9,10-dione (2.0 g, 9.6 mmol), isophthalaldehyde (0.64 g, 4.8 mmol), aniline (4.46 g, 48.0 mmol), ammonium acetate (2.96 g, 38.4 mmol) and glacial acetic acid (40 mL) was heated at 123 °C for 3 h. After cooled to room temperature, the reaction mixture was poured into distilled water under stirring. The resulting precipitate was filtered out and washed with water to give the crude product, which was further purified by vacuum sublimation to produce pure m-DPPI (1.78 g, 56%). 1H NMR (500 MHz, DMSO-d6): δ 8.95 (d, J = 8.0 Hz, 2H), 8.90 (d, J = 8.0 Hz, 2H), 8.72 (d, J = 8.0 Hz, 2H), 8.10 (s, 1H), 7.83 (t, J = 7.0 Hz, 2H), 7.75–7.69 (m, 12H),7.58 (t, J = 7.0 Hz, 2H), 7.46 (d, J = 7.5 Hz, 2H), 7.36 (t, J = 7.5 Hz, 2H), 7.28 (t, J = 8.0 Hz, 1H), 7.09 (d, J = 8.0 Hz, 2H). MS m/z: 662.1 [M]+ (calcd: 662.3). Anal. calcd (%) for C48H30N4: C, 86.98; H, 4.56; N, 8.45. Found: C, 86.76; H, 4.47; N, 8.53.
3-(1-Phenyl-1H-phenanthro[9,10-d]imidazol-2-yl)benzaldehyde (m-PPI-CHO). A mixture of phenanthrene-9,10-dione (0.50 g, 2.4 mmol), isophthalaldehyde (0.96 g, 7.2 mmol), aniline (2.23 g, 24 mmol), ammonium acetate (1.48 g, 19.2 mmol) and glacial acetic acid (30 mL) was heated at 123 °C for 3 h. After cooled to room temperature, the reaction mixture was poured into distilled water under stirring. The resulting precipitate was filtered out and washed with water to give the crude product, which was further purified by vacuum sublimation to produce m-PPI-CHO (726 mg, 76%). 1H NMR (500 MHz, DMSO-d6): δ 9.99 (s, 1H), 8.95 (d, J = 8.5 Hz, 1H), 8.90 (d, J = 8.5 Hz, 1H), 8.72 (d, J = 8.0 Hz, 1H), 8.16 (s, 1H), 7.93 (d, J = 8.0 Hz, 1H), 7.84–7.70 (m, 8H),7.56 (t, J = 7.5 Hz, 2H), 7.36 (t, J = 7.5 Hz, 1H), 7.13 (d, J = 8.0 Hz, 1H). MS m/z: 398.0 [M]+ (calcd: 398.1). Anal. calcd (%) for C28H18N2O: C, 84.40; H, 4.55; N,7.03. Found: C, 84.66; H, 4.47; N, 7.23.
1-Phenyl-2-(3-(1-phenyl-1H-phenanthro[9,10-d]imidazol-2-yl)phenyl)-1H-imidazo[4,5-f][1,10]phenanthroline (m-PPPI). The procedure is similar to that for synthesizing m-DPPI except that m-PPI-CHO and 1,10-phenanthroline-5,6-dione were used as the reactants instead of isophthalaldehyde and phenanthrene-9,10-dione, respectively. m-PPPI was obtained in a yield of 64%. 1H NMR (500 MHz, DMSO-d6): δ 9.28 (d, J = 4.5 Hz, 1H), 9.06 (d, J = 8.0 Hz, 1H), 8.98 (d, J = 4.5 Hz, 1H), 8.95 (d, J = 8.0 Hz,1H), 8.90 (d, J = 8.0 Hz, 1H), 8.71 (d, J = 8.0 Hz, 1H), 8.15 (s, 1H), 7.93 (dd, J = 8.5, 4.5 Hz, 1H), 7.84–7.69 (m, 12H), 7.58 (t, J = 8 Hz, 1H), 7.53–7.46 (m, 3H), 7.39–7.34 (m, 2H), 7.30 (t, J = 8 Hz, 1H), 7.09 (d, J = 7.5 Hz, 1H). MS m/z: 664.3 [M]+ (calcd: 664.2). Anal. calcd (%) for C46H28N6: C, 83.11; H, 4.25; N, 12.64. Found: C, 83.36; H, 4.47; N, 12.43.
Acknowledgements
This work was supported by the National Basic Research Program of China (2015CB655000), the National Natural Science Foundation of China (91333201 and 21221063) and Program for Chang Jiang Scholars and Innovative Research Team in University (No. IRT101713018).
Notes and references
-
(a) M. A. Baldo, D. F. O'Brien, Y. You, A. Shoustikov, S. Sibley, M. E. Thompson and S. R. Forrest, Nature, 1998, 395, 151 CrossRef CAS;
(b) M. A. Baldo, S. Lamansky, P. E. Burrows, M. E. Thompson and S. Forrest, Appl. Phys. Lett., 1999, 75, 4 CrossRef CAS PubMed;
(c) C. Adachi, M. A. Baldo, D. Brien, M. E. Thompson and S. R. Forrest, J. Appl. Phys., 2001, 90, 5048 CrossRef CAS PubMed;
(d) Y. Sun, N. C. Giebink, H. Kanno, B. Ma, M. E. Thompson and S. R. Forrest, Nature, 2006, 440, 908 CrossRef CAS PubMed;
(e) B. W. D'Andrade and S. R. Forrest, Adv. Mater., 2004, 16, 1585 CrossRef PubMed;
(f) S. Reineke, F. Lindner, G. Schwartz, N. Seidler, K. Walzer, B. Lussem and K. Leo, Nature, 2009, 459, 234 CrossRef CAS PubMed;
(g) K. Walzer, B. Maenning, M. Pfeiffer and K. Leo, Chem. Rev., 2007, 107, 1233 CrossRef CAS PubMed.
-
(a) K. T. Kamtekar, A. P. Monkman and M. R. Bryce, Adv. Mater., 2010, 22, 572 CrossRef CAS PubMed;
(b) Y. Zheng, A. S. Batsanov, M. A. Fox, H. A. Al-Attar, K. Abdullah, V. Jankus, M. R. Bryce and A. P. Monkman, Angew. Chem., Int. Ed., 2014, 53, 11616 CrossRef CAS PubMed;
(c) G. Li, D. Zhu, T. Peng, Y. Liu, Y. Wang and M. R. Bryce, Adv. Funct. Mater., 2014, 24, 7420 CrossRef CAS PubMed;
(d) G. M. Farinola and R. Ragni, Chem. Soc. Rev., 2011, 40, 3467 RSC;
(e) M. C. Gather, A. Köhnen and K. Meerholz, Adv. Mater., 2011, 23, 233 CrossRef CAS PubMed;
(f) G. Zhou, C.-L. Ho, W.-Y. Wong, Q. Wang, D. Ma, L. Wang, Z. Lin, T. B. Marder and A. Beeby, Adv. Funct. Mater., 2008, 18, 499 CrossRef CAS PubMed;
(g) T. Peng, H. Bi, Y. Liu, Y. Fan, H. Gao, Y. Wang and Z. Hou, J. Mater. Chem., 2009, 19, 8072 RSC;
(h) T. Peng, G. Li, K. Ye, C. Wang, S. Zhao, Y. Liu, Z. Hou and Y. Wang, J. Mater. Chem. C, 2013, 1, 2920 RSC;
(i) T. Peng, Y. Yang, H. Bi, Y. Liu, Z. Hou and Y. Wang, J. Mater. Chem., 2011, 21, 3551 RSC;
(j) X. Wang, S. L. Gong, D. Song, Z. H. Lu and S. Wang, Adv. Funct. Mater., 2014, 24, 7257 CrossRef CAS PubMed;
(k) X. Wang, Y.-L. Chang, J. S. Lu, T. Zhang, Z. H. Lu and S. Wang, Adv. Funct. Mater., 2014, 24, 1911 CrossRef CAS PubMed;
(l) C. Jiang, W. Yang, J. Peng, S. Xiao and Y. Cao, Adv. Mater., 2004, 16, 537 CrossRef CAS PubMed;
(m) T. Yu, Y. Cao, W. Su, C. Zhang, Y. Zhao, D. Fan, M. Huang, K. Yue and S. Z. D. Cheng, RSC Adv., 2014, 4, 554 RSC.
-
(a) S.-J. Su, H. Sasabe, T. I. Takeda and J. Kido, Chem. Mater., 2008, 20, 1691 CrossRef CAS;
(b) M. A. Baldo, C. Adachi and S. R. Forrest, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 62, 10967 CrossRef CAS;
(c) D. H. Kim, N. S. Cho, H.-Y. Oh, J. H. Yang, W. S. Jeon, J. S. Park, M. C. Suh and J. H. Kwon, Adv. Mater., 2011, 23, 2721 CrossRef CAS PubMed;
(d) H. Fukagawa, T. Shimizu, H. Hanashima, Y. Osada, M. Suzuki and H. Fujikake, Adv. Mater., 2012, 24, 5099 CrossRef CAS PubMed;
(e) J. Kwak, Y.-Y. Lyu, H. Lee, B. Choi, K. Char and C. Lee, J. Mater. Chem., 2012, 22, 6351 RSC;
(f) C.-H. Fan, P. Sun, T.-H. Su and C.-H. Cheng, Adv. Mater., 2011, 23, 2981 CrossRef CAS PubMed;
(g) Y.-C. Zhu, L. Zhou, H.-Y. Li, Q.-L. Xu, M.-Y. Teng, Y.-X. Zheng, J.-L. Zuo, H.-J. Zhang and X.-Z. You, Adv. Mater., 2011, 23, 4041 CrossRef CAS PubMed.
-
(a) L. Duan, J. Qiao, Y. Sun and Y. Qiu, Adv. Funct. Mater., 2011, 23, 1137 CrossRef CAS PubMed;
(b) Y. Tao, C. Yang and J. Qin, Chem. Soc. Rev., 2011, 40, 2943 RSC.
-
(a) S. O. Jeon, K. S. Yook, C. W. Joo and J. Y. Lee, Adv. Funct. Mater., 2009, 19, 3644 CrossRef CAS PubMed;
(b) M. Guan, Z. Chen, Z. Bian, Z. Liu, Z. Gong, W. Baik, H. Lee and C. Huang, Org. Electron., 2006, 7, 330 CrossRef CAS PubMed;
(c) L. Zeng, T. Y.-H. Lee, P. B. Merkel and S. H. Chen, J. Mater. Chem., 2009, 19, 8772 RSC;
(d) C.-H. Chang, M.-C. Kuo, W.-C. Lin, Y.-T. Chen, K.-T. Wong, S.-H. Chou, E. Mondal, R. C. Kwong, S. Xia, T. Nakagawa and C. Adachi, J. Mater. Chem., 2012, 22, 3832 RSC;
(e) D. Wagner, S. T. Hoffmann, U. Heinemeyer, I. Münster, A. Köhler and P. Strohriegl, Chem. Mater., 2013, 25, 3758 CrossRef CAS;
(f) H.-G. Jang, W. Song, J. Y. Lee and S.-H. Hwang, RSC Adv., 2015, 5, 17030 RSC;
(g) Y. Zhang, W. Haske, D. Cai, S. Barlow, B. Kippelen and S. R. Marder, RSC Adv., 2013, 3, 23514 RSC.
-
(a) Y. Tao, Q. Wang, L. Ao, C. Zhong, C. Yang, J. Qin and D. Ma, J. Phys. Chem. C, 2010, 114, 601 CrossRef CAS;
(b) D. Yu, F. Zhao, C. Han, H. Xu, J. Li, Z. Zhang, Z. Deng, D. Ma and P. Yan, Adv. Mater., 2012, 24, 509 CrossRef CAS PubMed;
(c) C. Fan, F. Zhao, P. Gan, S. Yang, T. Liu, C. Zhong, D. Ma, J. Qin and C. Yang, Chem.–Eur. J., 2012, 18, 5510 CrossRef CAS PubMed.
-
(a) Z. Ge, T. Hayakawa, S. Ando, M. Ueda, T. Akiike, H. Miyamoto, T. Kajita and M. Kakimoto, Org. Lett., 2008, 10, 421 CrossRef CAS PubMed;
(b) Z. Gao, M. Luo, X. Sun, H. Tam, M. Wong, B. Mi, P. Xia, K. Cheah and C. Chen, Adv. Mater., 2009, 21, 688 CrossRef CAS PubMed.
-
(a) M. Rothmann, S. Haneder, E. Como, C. Lennartz, C. Schildknecht and P. Strohriegl, Chem. Mater., 2010, 22, 2403 CrossRef CAS;
(b) A. Padmaperuma, L. Sapochak and P. Burrows, Chem. Mater., 2006, 18, 2389 CrossRef CAS;
(c) S. Jeon, K. Yook, C. Joo and J. Lee, Adv. Mater., 2010, 22, 1872 CrossRef CAS PubMed;
(d) C. Han, Z. Zhang, H. Xu, J. Li, G. Xie, R. Chen, Y. Zhao and W. Huang, Angew. Chem., Int. Ed., 2012, 51, 10104 CrossRef CAS PubMed;
(e) A. Wada, T. Yasuda, Q. Zhang, Y. Yang, I. Takasu, S. Enomoto and C. Adachi, J. Mater. Chem. C, 2013, 1, 2404 RSC.
- A. Chaskar, H.-F. Chen and K.-T. Wong, Adv. Mater., 2011, 23, 3876 CrossRef CAS PubMed.
-
(a) J. Ding, B. Zhang, J. Lü, Z. Xie, L. Wang, X. Jing and F. Wang, Adv. Mater., 2009, 21, 4983 CrossRef CAS PubMed;
(b) J. Ding, Q. Wang, L. Zhao, D. Ma, L. Wang, X. Jing and F. Wang, J. Mater. Chem., 2010, 20, 8126 RSC;
(c) X. Wang, S. Wang, Z. Ma, J. Ding, L. Wang, X. Jing and F. Wang, Adv. Funct. Mater., 2014, 24, 3413 CrossRef CAS PubMed;
(d) S. Shao, J. Ding, T. Ye, Z. Xie, L. Wang, X. Jing and F. Wang, Adv. Mater., 2011, 23, 3570 CrossRef CAS PubMed.
-
(a) K. Wang, S. Wang, J. Wei, S. Chen, D. Liu and Y. Liu, J. Mater. Chem. C, 2014, 2, 6817 RSC;
(b) S. Gong, Y. Chen, J. Luo, C. Yang, C. Zhong, J. Qin and D. Ma, Adv. Funct. Mater., 2011, 21, 1168 CrossRef CAS PubMed;
(c) Q. Wang, J. Ding, D. Ma, Y. Cheng, L. Wang, X. Jing and F. Wang, Adv. Funct. Mater., 2009, 19, 84 CrossRef CAS PubMed;
(d) H.-H. Chou and C.-H. Cheng, Adv. Mater., 2010, 22, 2468 CrossRef CAS PubMed;
(e) Y.-L. Chang, S. Yin, Z. Wang, M. G. Helander, J. Qiu, L. Chai, Z. Liu, G. D. Scholes and Z. Lu, Adv. Funct. Mater., 2013, 23, 705 CrossRef CAS PubMed;
(f) Y. Qian, Y. Ni, S. Yue, W. Li, S. Chen, Z. Zhang, L. Xie, M. Sun, Y. Zhao and W. Huang, RSC Adv., 2015, 5, 29828 RSC.
-
(a) Y. Yuan, D. Li, X. Zhang, X. Zhao, Y. Liu, J. Zhang and Y. Wang, New J. Chem., 2011, 35, 1534 RSC;
(b) Y. Zhang, S.-L. Lai, Q.-X. Tong, M.-Y. Chan, T.-W. Ng, Z.-C. Wen, G.-Q. Zhang, S.-T. Lee, H.-L. Kwonge and C.-S. Lee, J. Mater. Chem., 2011, 21, 8206 RSC;
(c) Y. Zhang, S.-L. Lai, Q.-X. Tong, M.-F. Lo, T.-W. Ng, M.-Y. Chan, Z.-C. Wen, J. He, K.-S. Jeff, X.-L. Tang, W.-M. Liu, C.-C. Ko, P.-F. Wang and C.-S. Lee, Chem. Mater., 2012, 24, 61 CrossRef;
(d) Y. Yuan, J.-X. Chen, F. Lu, Q.-X. Tong, Q.-D. Yang, H.-W. Mo, T.-W. Ng, F.-L. Wong, Z.-Q. Guo, J. Ye, Z. Chen, X.-H. Zhang and C.-S. Lee, Chem. Mater., 2013, 25, 4957 CrossRef CAS;
(e) C.-J. Kuo, T.-Y. Li, C.-C. Lien, C.-H. Liu, F.-I. Wu and M.-J. Huang, J. Mater. Chem., 2009, 19, 1865 RSC;
(f) Y.-N. Yan, W.-L. Pan and H.-C. Song, Dyes Pigm., 2010, 86, 249 CrossRef CAS PubMed;
(g) S. Zhuang, R. Shangguan, H. Huang, G. Tu, L. Wang and X. Zhu, Dyes Pigm., 2014, 101, 93 CrossRef CAS PubMed;
(h) S. Zhuang, R. Shangguan, J. Jin, G. Tu, L. Wang, J. Chen, D. Ma and X. Zhu, Org. Electron., 2012, 13, 3050 CrossRef CAS PubMed;
(i) D. Kumar, K. R. Thomas, C. C. Lin and J. H. Jou, Chem.–Asian J., 2013, 8, 2111 CrossRef CAS PubMed;
(j) B. Wang, G. Mu, J. Tan, Z. Lei, J. Jin and L. Wang, J. Mater. Chem. C, 2015, 3, 7709 RSC.
-
(a) Z. Wang, P. Lu, S. Chen, Z. Gao, F. Shen, W. Zhang, Y. Xu, H. S. Kwokb and Y. Ma, J. Mater. Chem., 2011, 21, 5451 RSC;
(b) W. Li, D. Liu, F. Shen, D. Ma, Z. Wang, T. Feng, Y. Xu, B. Yang and Y. Ma, Adv. Funct. Mater., 2012, 22, 2797 CrossRef CAS PubMed;
(c) Z. Wang, P. Lu, S. Chen, Z. Gao, F. Shen, W. Zhang, Y. Xu, H. S. Kwok and Y. Ma, J. Mater. Chem., 2011, 21, 5451 RSC;
(d) Z. Wang, Z. Gao, S. Xue, Y. Liu, W. Zhang, C. Gu, F. Shen, P. Lu and Y. Ma, Polym. Bull., 2012, 69, 273 CrossRef CAS;
(e) Z. M. Wang, X. H. Song, Z. Gao, D. W. Yu, X. J. Zhang, P. Lu, F. Z. Shen and Y. G. Ma, RSC Adv., 2012, 2, 9635 RSC;
(f) Z. Wang, Y. Feng, H. Li, Z. Gao, X. Zhang, P. Lu, P. Chen, Y. Ma and S. Liu, Phys. Chem. Chem. Phys., 2014, 16, 10837 RSC;
(g) Z. Gao, G. Cheng, F. Shen, S. Zhang, Y. Zhang, P. Lu and Y. Ma, Laser Photonics Rev., 2014, 8, L6 CrossRef CAS PubMed;
(h) Z. Gao, Y. Liu, Z. Wang, F. Shen, H. Liu, G. Sun, L. Yao, Y. Lv, P. Lu and Y. Ma, Chem.–Eur. J., 2013, 19, 2602 CrossRef CAS PubMed;
(i) Z. Wang, Y. Feng, S. Zhang, Y. Gao, Z. Gao, Y. Chen, X. Zhang, P. Lu, B. Yang, P. Chen, Y. Ma and S. Liu, Phys. Chem. Chem. Phys., 2014, 16, 20772 RSC;
(j) H. Liu, Q. Bai, L. Yao, H. Zhang, H. Xu, S. Zhang, W. Li, Y. Gao, J. Li, P. Lu, H. Wang, B. Yang and Y. Ma, Chem. Sci., 2015, 6, 3797 RSC.
-
(a) H. Huang, Y. Wang, S. Zhuang, X. Yang, L. Wang and C. Yang, J. Phys. Chem. C, 2012, 116, 19458 CrossRef CAS;
(b) H. Huang, Y. Wang, B. Wang, S. Zhuang, B. Pan, X. Yang, L. Wang and C. Yang, J. Mater. Chem. C, 2013, 1, 5899 RSC;
(c) X. Zhang, J. Lin, X. Ouyanga, Y. Liu, X. Liu and Z. Ge, J. Photochem. Photobiol., A, 2013, 268, 37 CrossRef CAS PubMed.
-
(a) K. Wang, F. Zhao, C. Wang, S. Chen, D. Chen, H. Zhang, Y. Liu, D. Ma and Y. Wang, Adv. Funct. Mater., 2013, 23, 2672 CrossRef CAS PubMed;
(b) K. Wang, S. Wang, J. Wei, Y. Miao, Y. Liu and Y. Wang, Org. Electron., 2014, 15, 3211 CrossRef CAS PubMed;
(c) C. Li, S. Wang, W. Chen, J. Wei, G. Yang, K. Ye, Y. Liu and Y. Wang, Chem. Commun., 2015, 51, 10632 RSC;
(d) D. Liu, M. Du, D. Chen, K. Ye, Z. Zhang, Y. Liu and Y. Wang, J. Mater. Chem. C, 2015, 3, 4394 RSC;
(e) K. Wang, S. Wang, J. Wei, S. Chen, D. Liu, Y. Liu and Y. Wang, J. Mater. Chem. C, 2014, 2, 6817 RSC.
- R. M. F. Batista, S. P. G. Costa, M. Belsley, C. Lodeiro and M. M. M. Raposo, Tetrahedron, 2008, 64, 9230 CrossRef CAS PubMed.
- Y. Tao, Q. Wang, C. Yang, Q. Wang, Z. Zhang, T. Zou, J. Qin and D. Ma, Angew. Chem., Int. Ed., 2008, 47, 8104 CrossRef CAS PubMed.
-
(a) Z. Zhang, Y. Zhang, D. Yao, H. Bi, I. Javed, Y. Fan, H. Zhang and Y. Wang, Cryst. Growth Des., 2009, 9, 5069 CrossRef CAS;
(b) Z. Zhang, R. M. Edkins, J. Nitsch, K. Fucke, A. Eichhorn, A. Steffen, Y. Wang and T. B. Marder, Chem.–Eur. J., 2015, 21, 177 CrossRef CAS PubMed;
(c) X. Wang, Q. Liu, H. Yan, Z. Liu, M. Yao, Q. Zhang, S. Gong and W. He, Chem. Commun., 2015, 51, 7497 RSC.
-
(a) Y. Tao, Q. Wang, Y. Shang, C. Yang, L. Ao, J. Qin, D. Ma and Z. Shuai, Chem. Commun., 2009, 77 RSC;
(b) Y. Tao, Q. Wang, L. Ao, C. Zhong, J. Qin, C. Yang and D. Ma, J. Mater. Chem., 2010, 20, 1759 RSC;
(c) Y. Tao, Q. Wang, C. Yang, C. Zhong, J. Qin and D. Ma, Adv. Funct. Mater., 2010, 20, 2923 CrossRef CAS PubMed.
- V. Adamovich, J. Brooks, A. Tamayo, A. M. Alexander, P. I. Djurovich, B. W. D'Andrade, C. Adachi, S. R. Forrest and M. E. Thompson, New J. Chem., 2002, 26, 1171 RSC.
- C. M. Cardona, W. Li, A. E. Kaifer, D. Stockdale and G. C. Bazan, Adv. Mater., 2011, 23, 2367 CrossRef CAS PubMed.
-
(a) L. Deng, J. Li, G.-X. Wang and L.-Z. Wu, J. Mater. Chem. C, 2013, 1, 8140 RSC;
(b) X. Ouyang, D. Chen, S. Zeng, X. Zhang, S. Su and Z. Ge, J. Mater. Chem., 2012, 22, 23005 RSC;
(c) Q. Li, L.-S. Cui, C. Zhong, Z.-Q. Jiang and L.-S. Liao, Org. Lett., 2014, 16, 1622 CrossRef CAS PubMed.
-
(a) C.-J. Zheng, J. Ye, M.-F. Lo, M.-K. Fung, X.-M. Ou, X.-H. Zhang and C.-S. Lee, Chem. Mater., 2012, 24, 643 CrossRef CAS;
(b) W. Jiang, L. Duan, J. Qiao, G. Dong, L. Wang and Y. Qiu, Org. Lett., 2011, 13, 3146 CrossRef CAS PubMed;
(c) F.-M. Hsu, C.-H. Chien, Y.-J. Hsieh, C.-H. Wu, C.-F. Shu, S.-W. Liu and C.-T. Chen, J. Mater. Chem., 2009, 19, 8002 RSC;
(d) S.-J. Su, E. Gonmori, H. Sasabe and J. Kido, Adv. Mater., 2008, 20, 4189 CAS.
- J. P. Duan, P. P. Sun and C. H. Cheng, Adv. Mater., 2003, 15, 224 CrossRef CAS PubMed.
- B. C. Lin, C. P. Cheng, Z.-Q. You and C.-P. Hsu, J. Am. Chem. Soc., 2005, 127, 66 CrossRef CAS PubMed.
- X. Qiao, Y. Tao, Q. Wang, D. Ma, C. Yang, L. Wang, J. Qin and F. Wang, J. Appl. Phys., 2010, 108, 034508 CrossRef PubMed.
-
(a) C. Murawski, K. Leo and M. C. Gather, Adv. Mater., 2013, 25, 6801 CrossRef CAS PubMed;
(b) L. Zhu, Z. Wu, J. Chen and D. Ma, J. Mater. Chem. C, 2015, 3, 3304 RSC.
-
(a) Y. Feng, P. Li, X. Zhuang, K. Ye, T. Peng, Y. Liu and Y. Wang, Chem. Commun., 2015, 51, 12544 RSC;
(b) C.-H. Chen, L.-C. Hsu, P. Rajamalli, Y.-W. Chang, F.-I. Wu, C.-Y. Liao, M.-J. Chiu, P.-Y. Chou, M.-J. Huang, L.-K. Chu and C.-H. Cheng, J. Mater. Chem. C, 2014, 2, 6183 RSC;
(c) C.-H. Shih, P. Rajamalli, C.-A. Wu, M.-J. Chiu, L.-K. Chu and C.-H. Cheng, J. Mater. Chem. C, 2015, 3, 1491 RSC.
- T. Komino, H. Nomura, T. Koyanagi and C. Adachi, Chem. Mater., 2013, 25, 3038 CrossRef CAS.
-
(a) SHELXTL, version 5.1, Siemens Industrial Automation, Inc., 1997 Search PubMed;
(b) G. M. Sheldrick, SHELXS-97, Program for Crystal Structure Solution, University of Göttingen, Göttingen, 1997 Search PubMed.
- E. Runge and E. K. U. Gross, Phys. Rev. Lett., 1984, 52, 997 CrossRef CAS.
- A. D. J. Becke, Chem. Phys., 1993, 98, 5648 CAS.
- M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, J. A. Montgomery Jr, T. Vreven, K. N. Kudin, J. C. Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G. Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y. Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S. Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K. Raghavachari, J. B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J. Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez and J. A. Pople, Gaussian 03, Revision C.02, Gaussian, Inc., Pittsburgh, PA, 2003 Search PubMed.
Footnote |
† Electronic supplementary information (ESI) available: TGA and DSC curves and additional photophysical, crystallographic and electroluminescent data. CCDC 1008378. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/c5ra14570h |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.