Sunhee
Lee
and
Won-Sik
Han
*
Department of Chemistry, Seoul Women's University, 621 Hwarang-ro, Seoul 01797, Republic of Korea. E-mail: wshan@swu.ac.kr
First published on 7th May 2020
The main difficulties hindering development of a deep-blue phosphorescent cyclometalated Ir(III) complex are insufficient colour purity, i.e., failure to achieve ideal Commission Internationale de L'Eclairage (CIE) coordinates of (0.14, 0.09), and insufficient emission efficiency and stability. The latter problem is due to the highly energetic and hot excited states of these complexes, which yield faster decomposition. Therefore, control of the excited-state properties of cyclometalated Ir(III) complexes through systematic chemical modification of the ligands is being extensively investigated, with the aim of developing efficient and stable blue phosphorescent materials. The most common strategies towards achievement of a blue phosphorescent cyclometalated Ir(III) complex involve (1) substitution of electron-withdrawing F atoms at the cyclometalating ligands that stabilise the HOMO orbitals and (2) use of a heteroleptic system with electron-rich ancillary ligands bearing a 5-membered ring heterocycle to increase the LUMO energy level. However, the C–F bonds on the cyclometalating ligands have been found to be inherently unstable during device operation; thus, other types of electron-withdrawing groups (e.g., the cyano, trifluoromethyl, and sulfonyl groups) have been applied. Along with phosphorescence colour tuning to blue, the influence of the ligand structure on the photoluminescence quantum yield (PLQY) is also being intensively investigated. Two major PLQY lowering mechanisms for blue emissive Ir(III) complexes have been identified: (1) the vibronic-coupled non-radiative decay process and (2) crossing from the emissive state to an upper non-emissive 3MC excited state. To enhance the PLQY, mechanism (1) can be suppressed by employing rigid ligand frameworks to restrict intramolecular motion, whereas mechanism (2) can be prevented by destabilising the 3MC state using strong σ donor ligands such as N-heterocyclic carbenes. This review summarises the fundamental photophysics of cyclometalated Ir(III) complexes and surveys design strategies for efficient blue phosphorescent Ir(III) complexes, to provide a guide for future research in this field.
In this regard, cyclometalated Ir(III) complexes play an essential role in OLED applications such as flat panel displays and solid-state lighting.1,13–15 Previous studies have established the structure–property relationships within cyclometalated Ir(III) complexes that enable high-efficiency phosphorescence emission with full spectral coverage spanning the near ultraviolet (UV) to near infrared. However, phosphorescence tuning over the entire visible spectrum remains a challenge. The design and synthesis of deep-blue-emitting cyclometalated Ir(III) complexes is particularly difficult. Although high external quantum efficiencies (EQEs) exceeding 30% have been achieved for green and red OLED devices,16–20 high EQEs for deep-blue OLED devices have rarely been reported.21–24
The main difficulties in developing deep-blue-emissive Ir(III) complexes are the following: (1) they lack sufficient colour purity for the ideal Commission Internationale de L'Eclairage (CIE) coordinates of (0.14, 0.09), (2) they have insufficient emission efficiency, and (3) attainment of an appropriate host and carrier transport materials with sufficient triplet energy levels is challenging. (Note: The materials for the host and carrier transport materials are not discussed in this review.) These difficulties are due to the highly energetic excited states of blue phosphors, which can yield faster decomposition than those of green and red phosphors.25–28 In addition, the upper-lying excited states of blue-emissive Ir(III) complexes can induce several degradation processes, for example, triplet–triplet annihilation,29,30 triplet–polaron annihilation,31 and polaron–polaron annihilation.32
Note that review articles of blue phosphorescent cyclometalated Ir(III) complexes for applications in OLEDs have been well presented previously.33–50 In this review, we will discusses the current research status regarding development of cyclometalated Ir(III) complexes towards a blue-emissive dopant for OLEDs. We summarise the fundamental photophysics of cyclometalated Ir(III) complexes, considering radiative and non-radiative (NR) decay processes (section 2). We then discuss the structure–property relationship dependence on the ancillary ligand (section 3). We hope this review will be helpful to OLED researchers working to design efficient deep-blue phosphorescent Ir(III) complexes.
Fig. 1 Modified Jablonski diagram with time scales for photophysical processes in molecular system. Bold and dotted lines indicate radiative and non-radiative (NR) decay processes, respectively. |
(1) Photoabsorption: As a molecule absorbs photons, an electron is excited from the singlet ground state (S0), to the lowest (S1) or higher singlet excited states (Sn). The transition rate is very fast, at >1015 s−1. In general, a transition from S0 to the lowest triplet state (T1) has low probability because the electron spin is parallel to the ground-state spin; thus, this transition is called ‘forbidden’. However, this forbidden transition can be observed under specific conditions, for example, for internal and external heavy-atom effects with very low extinction coefficients (ε), with for which εmax = 10−1–10−2.
(2) Vibrational relaxation: Immediately after excitation to S1 or Sn, the molecule population distribution in the Franck–Condon state of the higher excited-vibrational states (v′ ≥ 1) relaxes to a less energetic vibrational state (v′ = 0) through vibrational energy transfer. This process also occurs immediately after IC and intersystem crossing. The vibrational relaxation rate is in the range of a few picoseconds, at >1012 s−1.
(3) IC: This process involves a change in the electronic states, for example, Sn to Sn−1, which have energetically degenerate vibrational states. Therefore, there is no energy change and this process is radiationless. The IC range is <108 s−1 and depends on the energy gap between the Sn and Sn−1 states according to the ‘Energy Gap Law’.
(4) Fluorescence: This process is a radiative transition from S1 to S0, as allowed by the selection rules, at a rate of ≤109 s−1. As the IC rate from Sn to Sn−1 is very high (in the picoseconds range), most fluorescence occurs from the lowest S1 state; this is called ‘Kasha's Rule’.
(5) ISC: This process is similar to IC. However, although the molecular spin state remains the same for IC, the ISC process requires a change in spin state. The latter process involves transitions from S1 to the higher excited-vibrational states of T1 and Tn, followed by relaxation to less energetic vibrational states (similar to IC). The ISC rate is in the range of 1012 s−1 and depends on the degree of spin–orbit coupling and the ‘El-Sayed Rule’. The transition from T1 to S0 is also an ISC process, but with a lower transition rate (≤106 s−1) than the S1 → T1 transition according to the “Energy Gap Law” (the energy gap between the T1 and S0 states is usually larger than that between the S1 and T1 states).
(6) Phosphorescence: This process is another radiative transition and involves transition from the T1 to S0 states. As the direct formation of T1 from S0 is a forbidden transition, the T1 state is usually generated via the following sequence: excitation from S0 to Sn → rapid IC from Sn to S1 → ISC from S1 to T1. The phosphorescence rate for organic compounds is approximately 103 s−1 but is much higher for organometallic complexes that possess heavy metals, being in the range of >106 s−1.
(7) Photochemical processes: The energy or electron in the excited state can participate in chemical reactions. In general, these processes occur from the S1 and T1 states. In the case of the unimolecular reaction process, the reaction rate is in the range of ≤1012 s−1. For the intermolecular (bimolecular) reaction process, the reaction rate depends on the solvent temperature (T) and viscosity (≤1010 M−1 s−1).
In particular, triplet-related processes are a major focus in the context of cyclometalated Ir(III) complexes for OLED application, as the maximum internal quantum efficiency for conversion of electric energy to photons can be achieved through phosphorescence.
As an example, Fig. 2 shows the absorption spectrum of the well-known cyclometalated homoleptic Ir(III) complex, fac-tris[2-phenylpyridine]iridium(III), fac-Ir(ppy)3, (1). The absorption spectrum involves strong 1LC spin-allowed π–π* transitions at <300 nm and spin-allowed d–π* transitions (1MLCT) in the 320–430 nm range. The >435 nm absorptions are attributable to a spin-forbidden 3MLCT transition with very low ε; this is a consequence of the heavy-atom effect yielding a strong SOC.52 However, it should be noted that the lowest energy state of 1 is a hybrid state with 3MLCT and 3LC characteristics (vide infra).
Fig. 2 (a) Absorption spectrum in 10 μM solution of dichloromethane (inset: chemical structure) and (b) orbital contributions of HOMO and LUMO of 1. |
These assignments for the transitions of 1 are supported by theoretical calculations.53–55 According to those calculations, the molecular orbitals mainly participating in generation of the lowest triplet state are from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO). As shown in Fig. 2b, the HOMO is delocalised over the Ir t2g orbital and the π orbitals of the phenyl ring in the 2-ppy ligand, whereas the LUMO exclusively involves the π* orbitals of the pyridine ring in the 2-ppy. Therefore, one of the causes of the S0 → T1 transition is charge transfer from the 5d Ir metal orbital to the ppy ligands. Further studies involving time-dependent density functional theory (TD-DFT) calculations have indicated that the lowest triplet state is in a hybrid form of the MLCT and LC transition states.56,57 In the triplet manifold, the dominant LC state may be lowest. This is because the singlet–triplet splitting from the electron-exchange interaction is much smaller for MLCT than LC π–π*, as the orbitals in the MLCT excited state have greater spatial extension. The optimised structure of the lowest triplet energy state indicates broken C3 symmetry and excitation localised on a single ligand, supporting hybridisation in the triplet excited states. More detailed analysis based on multi-configurational self-consistent field orbitals and second-order configurational interactions have been performed to study the different radiative and NR processes of fac-1 and mer-1.58 Heteropletic59,60 and bis-tridentate61 Ir(III) complexes have also yielded similar calculation results.
Yersin et al. investigated the detailed properties of the emitting triplet state of 1, including the corresponding radiative and NR transitions.52,66 The lowest triplet state can be divided into three substates: I, II, and III. For most organometallic complexes containing transition metal ions, the transition between the lowest substate I and S0 is forbidden, but the other transitions from II and III to S0 are permitted.68–71 The same results were found for 1.66 Indeed, three emissive triplet substates, I–III, of 1 in tetrahydrofuran (THF) solution were identified from the temperature-dependent emission spectra (1.2 K ≤ T ≤ 300 K) and classified as substates of a 3MLCT state. Through further combined analysis with application of high magnetic fields, the energy differences between the substates were successfully estimated, i.e., ΔEII–I = 13.5 cm−1 and ΔEIII–I = 83.5 cm−1, as depicted in Fig. 3. This relatively large total zero-field splitting (ZFS) indicates that the lowest triplet state of 1 is a 3MLCT state. The same assignments were concluded from theoretical calculations.53,56,72
Fig. 3 Energy levels for three lowest triplet substates, I, II, and III, and decay times of 1. Adapted with permission from ref. 66. Copyright 2003, Elsevier. |
Heteroleptic Ir(III) complexes possess more complex photophysical processes than those of a homolectic system, because the presence of the ancillary ligand (L′) induces additional transitions such as L′C and ML′CT. This is in addition to the ligand-to-ligand charge transfer (LL′CT) transition that occurs between a main ligand and ancillary ligand, as shown in Fig. 4.
The most well-known blue-emissive heteroleptic Ir(III) is likely a bis[2-(4,6-difluorophenyl)pyridinato-C2,N](picolinato)iridium(III), commonly known as FIrpic (2).73Fig. 5 shows typical emission spectra of 2 at RT and 77 K. The RT emission spectrum exhibits blue phosphorescence peaks at 471 and 495 nm with fine vibronic structure, which are virtually independent of the solvent polarity. These observations indicate that the emitting lowest triplet state has a predominantly 3LC character with a minor 3MLCT contribution. Thus, if the cyclometalated Ir(III) complex emission has a non-vibronic structure and the emission changes depending on the solvent polarity, the lowest triplet state has a predominantly 3MLCT character. At low temperature (77 K), the emission spectrum of 2 shows narrower and highly structured emission bands at 461 and 495 nm, indicating that the 3LC character is increased at this temperature. A small thermally induced Stokes shift (ΔES) with of 10 nm also indicates that the nature of the lowest triplet state in 2 is based on the 3LC state. In general, small and large ΔES indicate 3LC and 3MLCT excited states, respectively.74–76 The PLQY of 2 in dilute solution was initially reported to be approximately 0.6. Note, however, that this value was later corrected to higher values of >0.90.77–79 Accordingly, the radiative- and NR constants of 2 were estimated to be kr = 5 × 105 s−1 and knr = 0.5 × 105 s−1, respectively.
Temperature-independent decay processes occur via direct crossing between the two potential energy surfaces of the triplet and ground states. These are vibronic-coupled NR decay process, as studied in detail by Samuel et al.80,81 Those researchers suggested that stronger vibrational coupling yields an increased NR decay rate and decreased PLQY. They estimated the vibrational coupling values using the Huang–Rhys factor, SM, which roughly quantifies the structural distortion (ΔQ) of the excited state relative to the ground state.82,83 If SM = 0, the geometries of the excited and ground states should be identical, and only a sharp peak corresponding to the 0–0 transition should be observed. With increased ΔQ and, hence, appearance of the 0–1 transition, the SM value can be estimated using the intensity ratios of the 0–0 and 0–1 vibrational peaks: SM = (I0–1/I0–0). Thus, higher SM indicates increased vibrational coupling between the potential surfaces of the triplet excited state and the ground state; this yields lower PLQY as the transition probability is increased with increasing overlap between the initial and final states.
Samuel et al. also found that the experimental phosphorescence lifetime (τP) value of 4 (Fig. 6) in 2-methyl-tetrahydrofuran (2-MeTHF) increases from 0.9 μs at 300 K to 2.2 μs at 200 K and to 3.2 μs at 77 K.81 An Arrhenius plot showing the temperature-dependence of τP gives activation energy of 0.27 eV (2178 cm−1). In the case of 5, τP changes from 0.15 μs at 290 K to 3 μs at 200 K, but is constant (3 μs) at ≤200 K. The temperature-dependence of τP gives an activation energy of 0.67 eV (5404 cm−1). On polymethyl methacrylate (PMMA) solid films, the emission decays of 3–5 (Fig. 6) reveal bi-exponential behaviour. The longer component for 5 increases from 1.35 μs at 310 K to 3.7 μs at 77 K, and the temperature dependence of τP gives an activation energy of 0.13 eV (1049 cm−1); this is much smaller than that for the homogeneous solution in 2-MeTHF, as shown in Fig. 6. As the activation energies differ considerably between homogeneous solutions and solid films, they suggested that a vibronic-coupled NR decay mechanism could play an important role in lowering the PLQY. Such NR decay processes can be suppressed by employing rigid ligand frameworks.
Fig. 6 (a) Chemical structures of 3–5 and (b) Arrhenius (activation energy) plot of 5 dissolved in 2-MeTHF or solid PMMA host. Adapted with permission from ref. 81. Copyright 2008, Elsevier. |
We next consider, temperature-dependent NR decay processes, which occur for crossing from the emissive state to an upper non-emissive 3MC excited state. As the non-emissive 3MC states of cyclometalated Ir(III) complexes are high-energy, green and red phosphorescent Ir(III) complexes cannot access this state at RT. However, the upper-lying 3MLCT states of blue phosphorescent Ir(III) complexes can easily access this non-emissive 3MC state. In particular, Thompson et al. reported the generation of the 3MC state in blue phosphorescent Ir(III) complexes, 6–12.84 The temperature-dependences of τP in 2-MeTHF reveal two different lower- and higher-temperature regions. These can be analysed according to the Boltzmann model eqn (1):
(1) |
where, k0 is the decay rate from the lowest-energy triplet substate (TI), k1 and k2 are pre-exponential factors (decay rate constants), E1 and E2 are the activation energies for NR decay, kB is the Boltzmann constant, and T is temperature in kelvin. In the lower-temperature region, k1 = 105–106 s−1 and E1 = 40–120 cm−1 were obtained; the latter corresponds to the ZFS energies between substates I and III. The NR decay at low temperatures proceeds from the non-emissive substate III populated from the emissive substate I. In the higher-temperature region, E2 values of 4700 ± 100 (6 and 7), 3400 ± 100 (8 and 9), and 1600 ± 100 cm−1 (10–12) were obtained, along with k2 values of 4 × 1014 (6), 3.4 × 1013 (7), 5 × 1012 (8), 6.1 × 1012 (4), 1.2 × 1012 (10), 2.6 × 1011 (11), and 3.7 × 109 s−1 (12). The NR decay in the high-temperature region was attributed to crossing from the emissive lowest triplet state to the non-emissive 3MC state. The argument for this mechanism is supported by (E0–0 + E2) = 25680–27370 cm−1 for 6–11, respectively (corresponding to 74–76 kcal mol−1), which are close to the dissociation energies of Ir–ligand bonds. The large k2 values, which are associated with high-frequency vibrations accompanied by bond breaking, also support this interpretation. Based on the ‘Energy-Gap Law’, the knr value for vibration-coupled deactivation has been estimated as 4 × 103 s−1, sufficiently low for a minor impact of vibration-coupled decay on the net NR decay. Accordingly, the authors concluded that NR decay rates can be decreased and PLQY can be improved by increasing the energy separation between the emissive and non-emissive states. In addition, on the basis of a DFT calculation for 10, they propose a change from the 6-coordinated lowest triplet state to the 3MC state with a 5-coordinated trigonal bipyramid structure accompanied by rupturing of one Ir–N bond, as shown in Fig. 7. Later, structural change in the excited state is supported by theoretical calculations, which showed a lengthened Ir–N bond up to 2.70 Å in the 3MC state due to the strong σ-antibonding interactions between the metal and the N atom in the ppy ligand.85
Fig. 7 (a) Chemical structures of 6–12. (b) Three types of decay process from T1 state to the ground state: temperature-dependent radiative process, kr(T), or one of two NR decay processes. (c) Changes in chemical structures and spin density surfaces calculated for triplet states of six- and five-coordinated forms of 10. Adapted with permission from ref. 84. Copyright 2009, American Chemical Society. |
Haga et al. reported a similar phenomenon in Ir(III) complexes with tridentate pyrazolyl ligands, 13 and 14.86 The intense absorptions at 320–400 nm and weak bands at >410 nm were attributed to the 1MLCT transition and 3LC mixed with 3MLCT transitions, respectively. In those complexes, the emission spectra at 77 K are highly structured, with the 0–0 vibration band being highest; this is characteristic of emission from a 3LC-dominated state with a minor 3MLCT contribution. The emission behaviour is highly temperature-dependent, as revealed by the shift from considerably short lifetimes (0.18–140 ns) at 298 K to much longer lifetimes (3.9–13.1 μs) at 77 K. The temperature dependence of τp for 13 at 90–300 K reveal biphasic behaviour, which was analysed based on the following bi-exponential eqn (2):
(2) |
Fig. 8 Potential–energy curves involving thermal deactivation of phosphorescent state (P) calculated as function of bond length Ir–N(1) (Ir–N(4)) for 13 and 14. Adapted with permission from ref. 86. Copyright 2008, American Chemical Society. |
Photophysical properties | Oxidation (V) | Device performances | Ref. | |||||||
---|---|---|---|---|---|---|---|---|---|---|
λ em (nm) | τ em (μs) | PLQY | k r (s−1) | k nr (s−1) | EL (λmax, nm) | CIE (x, y) | EQEmax (%) | |||
a Measured at 77 K. | ||||||||||
1 | 510 | 1.9 | 0.40 | 2.1 × 105 | 3.2 × 105 | 0.31 | — | — | — | 9 |
— | — | — | — | — | — | 518 | (0.30, 0.63) | 14.3 | 160 | |
514 | — | — | — | — | — | 510 | (0.28, 0.63) | 5.7 | 88 | |
2 | 471, 495 | 1.74 | 0.94 | 5.5 × 105 | 0.3 × 105 | — | — | (0.15, 0.36) | 5.1 | 174 |
3 | 449, 479 | 1.08 | 0.66 | 6.1 × 105 | 3.1 × 105 | 0.28 | — | — | — | 80 and 96 |
4 | 428, 456 | 1.25 | 0.27 | 2.2 × 105 | 5.8 × 105 | 0.50 | — | — | — | 80 and 96 |
5 | 425, 450 | 0.15 | 0.03 | 2.0 × 105 | 6.5 × 106 | 0.72 | — | — | — | 80 and 96 |
6 | 500 | 1.7 | 0.95 | 5.6 × 105 | 0.3 × 105 | — | — | — | — | 84 |
7 | 475 | 2.6 | 0.93 | 3.6 × 105 | 0.3 × 105 | — | — | — | — | 84 |
8 | 500 | 1.2 | 0.55 | 4.6 × 105 | 3.8 × 105 | — | — | — | — | 84 |
9 | 457 | 1.3 | 0.60 | 4.6 × 105 | 3.1 × 105 | — | — | — | — | 84 |
10 | 412a | 0.002 | <0.01 | — | — | — | — | — | — | 84 |
11 | 388a | 0.007 | <0.01 | — | — | — | — | — | — | 84 |
12 | 382 | 1.1 | 0.37 | 1.1 × 105 | 3.4 × 105 | — | — | — | — | 84 |
13 | 456a | 0.0002 | <0.001 | — | — | 0.75 | — | — | — | 86 |
14 | 555 | 1.78 | 0.78 | 4.4 × 105 | 1.2 × 105 | 0.42 | — | — | — | 86 |
In 2006, De Cola et al. explored the number of F atoms in the same molecular architecture.89 When one more F atom is substituted at the 3-position of the phenyl ring, (16, Fig. 9), the emission shows a hypsochromic shift to 459 nm. Interestingly, however, the emission of the tetra-fluorinated complex (17, Fig. 9), shifts to lower energy at 468 nm. De Cola et al. suggested that the lower emission energy of 17 is due to the more positive reduction potentials caused by the presence of more F atoms. Note also that these two complexes have different emission quantum yields (0.30 and 0.53 for 16 and 17, respectively) and τem at RT (1600 and 2300 ns for 16 and 17, respectively). The highest EQE value of 5.5% was observed for an electroluminescent device containing 17. Those authors also observed different stabilities for devices fabricated with the fac- and mer-isomers. That is, the latter exhibited fast spectral changes in emission from the blue to green region of the spectrum.
Further theoretical studies, performed by Tian et al. in 2011 revealed that the differences between 16 and 17 are due to the different transition dipole moments.90 The same effects were also observed for cationic Ir(III) complexes. De Angelis et al. reported a combined experimental and theoretical study comparing 18 (a dfppy-based complex) and 19 (a ppy-based complex, see Fig. 9).91 Those authors controlled the phosphorescent emission wavelength and improved the quantum yields by modulating the electronic structures of the cyclometalated Ir(III) complexes through F functionalisation. The F-functionalised 19 showed blue-shifted emission at 463 nm with a higher PLQY (0.85) than 18 (λmax = 491 nm, PLQY = 0.80) in acetonitrile solution. This was related to a dramatic decrease in the NR deactivation rate constant, in agreement with the ‘Energy-Gap Law’. The τem values were also found to increase from 18 (2.4 μs) to 19 (4.1 μs), yielding an overall reduction in the kr; this suggests increasing π–π* character in the emitting excited states. DFT and TD-DFT calculations with solvent effects were conducted to characterise the lowest triplet excited states and revealed that the extensive mixing of the 3MLCT and π–π* contributions agrees with the τem increment for 19 compared to 18.
Other modifications were made by replacing phenyl pyridine ligands with phenyl heterocyclic ring systems, such as imidazole, pyrazole, and triazole. Note that imidazole-based homoleptic Ir(III) complexes have been used as blue dopants because the imidazole group can heighten the LUMO energy level, thereby enlarging the energy gap and increasing the T1 energy level. Kitamura et al. reported homoleptic and heteroleptic Ir(III) tirs(phenylimidazolinate) complexes.92 Upon replacement of the pyridyl ring with the imidazolyl ring, the LUMO mainly populates the phenyl ring; this is quite different from the LUMOs of ppy-based Ir(III) complexes. Accordingly, substitution of an F atom into the phenyl ring (20, Fig. 9) yields the most blue-shifted emission of 453 nm at RT (PLQY = 0.60), with a value of 446 nm at 77 K. In this case, π-electron-donating substituents can induce a blue-shift of the emission spectra. They fabricated an OLED device using 22, which exhibited efficient luminescence compared with a 1-based device. The two devices showed similar emission colours, but the emission luminance of the former was smaller (Lmax = 889 cd m−2 at 15 V compared to Lmax = 3490 cd m−2 at 13 V), because of inefficient carrier injection into the emitting layer.
Kang et al. also studied imidazole-based Ir(III) complexes, but with a bulky terphenyl unit on the N atom of the imidazoyl ring, (23–25, Fig. 9).93 Those researchers concluded that a terphenyl ligand without alkyl chains (23) is advantageous in terms of lifetime, whereas the terphenyl with alkyl chains (24, 25) is efficient in terms of PLQY. When employed in OLED devices, the EQEs of 24 and 25 were 21.1% and 21.3%, respectively, which were higher than that of 23 (19.2%). The authors explained that the high PLQYs of 24 and 25 assisted the triplet emission of the blue devices by effectively harvesting triplet excitons.
In 2018, Lee et al. reported diisopropylphenyl-functionalised phenylimidazole-based Ir(III) complexes (rather than a terphenyl group) (26 and 27, Fig. 9), which showed more blue-shifted emission at 454 nm.94 Even though the diisopropylphenyl group has lesser bulkiness than the terphenyl ring, this group still efficiently limits intermolecular aggregation and prevents the different self-quenching processes. Substitution of CN at position-5 (27), was found to significantly affect the device lifetime; a longer device lifetime exceeding 550 h at 200 cd m−2 was obtained, with EQEmax = 17.6% and CIE (0.15, 0.28).
Pyrazole-attached Ir(III) complexes were also investigated. For example, Thompson et al. reported that the MLCT transitions of the pyrazolyl-based complexes, 30–32 (Fig. 9) are hypsochromically shifted relative to pyridyl-based complexes (1, 28, and 29), due to the higher triplet energy of phenylpyrazole (3.28 eV) compared to ppy (2.88 eV).9 F substitution effectively shifts the emission further to the blue region, up to 21 nm. Notably, phosphorescence of 31 was observed at 390 nm, which has very rarely been reported to date. However, these homoleptic pyrazolyl-based complexes are not emissive at the RT. Therefore, pryrazolyl-based ligands tend to be used in heteroleptic rather than homoleptic systems (vide infra).
Samuel et al. previously reported a series of phenyl triazoletype Ir(III) complexes.80 As triazole has a higher LUMO energy than pyridine,95 replacement of the pyridyl moiety with a triazolyl ring was expected to shift the emission further to the blue region than that of the corresponding ppy-based Ir(III) complex. As expected, 6 has λmax at 449 nm, which is significantly blue shifted compared to that of 1 (λmax = 510 nm). Ir(III) complexes with F atom(s) 6–8 show hypsochromically shifted emissions relative to non-F-based complexes, by up to 425 nm. However, the PLQYs decrease with increased F atoms. This is because of the increased ligand triplet energy in the emissive energy state, which decreases the material radiative decay rate through vibronic-coupled NR decay, as explained in the previous section.
Powell et al. theoretically studied the effects of fluorination on the role of metal–ligand bond fission in the NR decay of excited states in cylcometalated Ir(III) complexes.26,96 The calculated activation barrier to the 3MC state shows a clear correlation with the experimentally obtained NR decay rate for a series of Ir(III) complexes. As 3MC state formation requires breaking of an Ir–N bond, Powell et al. compared Ir–N bond distances in the 3MC states of cylcometalated Ir(III) complexes. For 33 and 34 (Fig. 9), the Ir–N bond length changes in the 3MC state relative to the ground-state structure are much smaller than those of 35 and 36 (Fig. 9). Accordingly, the activation barrier to the 3MC state is lower for 35 and 36, which yields low PLQY.
Samuel et al. extended their work to dendronised triazole-based Ir(III) complexes, to attain a solution-processable material.97–99 F-Attached dendrimers 38–40 (Fig. 10) show phosphorescence emissions at approximately 441 nm; however, these materials are unsuitable for device applications because of their low triplet energy and vibrational quenching, which yield luminescence quenching or low thermal properties.97,98 To solve this problem, dendrons were later modified to have a twisted geometry and successfully utilised in a blue PH-OLED device.99 Dendrimer 41 shown in Fig. 10 is highly emissive, with PLQY = 0.94 and 0.6, and τem of 3.6 and 2.8 μs, in solution and neat film, respectively. Its high emissivity is attributed to the dendron rigidifying effect, which reduces the geometry change in the excited state. Using 41, Samuel et al. fabricated a preliminary OLED device and showed EQEmax of 3.9% with CIE coordinates of (0.16, 0.17). Dendritic approaches are still being investigated, especially for solution-processed phosphorescent OLEDs.100–106 Available photophysical- and electrochemical data of 15–41 with their device performances are summarised in Table 2.
Photophysical properties | Oxidation (V) | Device performances | Ref. | |||||||
---|---|---|---|---|---|---|---|---|---|---|
λ em (nm) | τ em (μs) | PLQY | k r (s−1) | k nr (s−1) | EL (λmax, nm) | CIE (x, y) | EQEmax (%) | |||
a Measured at 77 K. | ||||||||||
15 | 468 | 1.6 | 0.43 | 2.7 × 105 | 3.6 × 105 | 0.78 | — | — | — | 9 |
495 | — | 0.06 | — | — | — | — | (0.27, 0.38) | 0.6 | 88 | |
16 | 459, 486 | 0.1 | 0.03 | 1.9 × 105 | 4.4 × 105 | — | — | — | — | 89 |
17 | 468, 497 | 0.1 | 0.03 | 2.3 × 105 | 2.0 × 105 | — | 478, 511 | — | 3.2 | 89 |
18 | 491, 520 | 2.4 | 0.80 | 3.3 × 105 | 0.8 × 105 | 0.72 | — | — | — | 91 |
19 | 463, 493 | 4.1 | 0.85 | 2.1 × 105 | 0.4 × 105 | 1.00 | — | — | — | 91 |
20 | 453, 482 | 3.4 | 0.60 | 1.7 × 105 | 1.2 × 105 | — | — | — | — | 92 |
21 | 486, 518 | 2.8 | 0.40 | 1.4 × 105 | 2.1 × 105 | — | — | — | — | 92 |
22 | 461, 492 | 3.3 | 0.53 | 1.6 × 105 | 1.4 × 105 | — | 464, 494 | (0.23, 0.35) | — | 92 |
23 | 463, 487 | — | 0.38 | — | — | — | — | (0.17, 0.30) | 19.2 | 93 |
24 | 459, 487 | — | 0.45 | — | — | — | — | (0.17, 0.28) | 21.1 | 93 |
25 | 462, 487 | — | 0.50 | — | — | — | — | (0.17, 0.29) | 21.3 | 93 |
26 | 454, 484 | 1.1 | 0.87 | 7.9 × 105 | 1.2 × 105 | — | 455, 484 | (0.17, 0.30) | 18.9 | 94 |
27 | 462, 494 | 1.8 | 0.99 | 5.5 × 105 | 0.5 × 104 | — | 462, 494 | (0.15, 0.28) | 22.5 | 94 |
28 | 468 | 1.6 | 0.05 | 2.7 × 105 | 3.6 × 105 | 0.78 | — | — | — | 9 |
29 | 510 | 2.0 | 0.50 | 2.5 × 105 | 2.5 × 105 | 0.30 | — | — | — | 9 |
30 | 414a | 14a | — | — | — | 0.39 | — | — | — | 9 |
31 | 390a | 27a | — | — | — | 0.80 | — | — | — | 9 |
32 | 428 | 0.05 | — | — | — | 0.73 | — | — | — | 9 |
33 | 449 | 1.08 | 0.66 | 6.1 × 105 | 3.1 × 105 | 0.28 | — | — | — | 96 |
34 | 443 | 0.15 | 0.06 | 4.0 × 105 | 6.3 × 105 | 0.50 | — | — | — | 96 |
35 | 428 | 1.25 | 0.27 | 2.2 × 105 | 5.8 × 106 | 0.50 | — | — | — | 96 |
36 | 425 | 0.15 | 0.03 | 2.0 × 105 | 6.5 × 106 | 0.72 | — | — | — | 96 |
37 | 468, 495 | 2.8 | 0.76 | 2.7 × 105 | 0.9 × 105 | 0.31 | — | (0.18, 0.35) | 7.9 | 98 |
38 | 441, 468 | 22.0 | 0.59 | 0.3 × 105 | 0.2 × 105 | 0.53 | — | — | — | 97 and 99 |
39 | 441, 468 | 1.7 | 0.46 | — | — | 0.47 | — | — | — | 98 and 99 |
40 | 441, 470 | 1.7 | 0.45 | 2.7 × 105 | 3.3 × 105 | 0.45 | — | — | — | 98 and 99 |
41 | 435, 465 | 3.6 | 0.94 | 2.6 × 105 | 1.7 × 104 | 0.61 | 438, 466 | (0.16, 0.16) | 3.9 | 99 |
Thompson et al. reported photophysical and electrochemical properties of nineteen cyclometalated Ir(III) complexes with various ancillary ligands and three types of main ligand (dfppy, ppy, and terpyridine (tpy)).106 The ancillary ligands were chosen to be ‘non-chromophoric’, so as to drive the excited-state properties dominated by the main ligands and Ir. For this approach, the 3LC state energy was expected to be relatively constant for all complexes, whereas the 1MLCT state energies could be altered by varying the ancillary ligand. As expected, the F-substituted main ligand (dfppy) showed a blue-shifted emission relative to the tpy-based Ir(III) complex. The effects of the ancillary ligands on the excited-state properties of the cyclometalated Ir(III) complexes were independent of the choice of main ligand. That is, the cyclometalated Ir(III) complexes with the same ancillary ligands but different main ligands exhibited different kr values. Thompson et al. explained that the degree of 1MLCT character mixed into the T1 state decreases when the Ir(III) complex has a deeper HOMO energy level. This yields an increased energy gap between the singlet and triplet energies and consequently, lowers the kr values.
De Cola et al. studied triazole-based ancillary ligand systems, 42–48, as described in Fig. 11.107 Those researchers designed ppy main ligands while varying the phenyl-ring substitution positions at 4/6 or 5 with different substituents, i.e., H, F, and CF3. The ancillary ligands were based on phenyltriazole with substituent variations at the triazole ring positions. They performed a detailed theoretical analysis of the effects of the substituents in the main ligand to study the emissive state properties. Calculation results showed that the substituents not only affect the emission energy but, also, change the ordering of the lowest excited triplet states. The same group reported a series of heteroleptic Ir(III) complexes, 45, 48, and 49–54 (Fig. 11), in which the 1,2,4-triazole ancillary ligands are varied with different substituents while the main ligand is fixed as dfppy.108 By increasing the electron-withdrawing ability, the HOMO energy level is lowered, and consequently, the HOMO–LUMO gap is widened. This yields blue-shifted emission with narrower full width at half maximum. In this series, complexes 51 and 52 exhibit lower PLQY, which may be due to the torsional angle between the phenyl and triazole rings arising from F atoms in the ortho position. In these cases, the lowest MLCT states are shifted from the pyridyl-triazole to the ppy. Two preliminary devices were constructed using 49 and 51, with both exhibiting EQEs exceeding 7% together with a blue colour (CIEx,y = 0.17, 0.27). De Cola et al.'s work was extended to 1,2,3-triazole-based Ir(III) complexes, 55–57 (Fig. 11).109 The emission spectra in the solution for 1,2,3-triazole-based complexes are slightly blue-shifted with respect to their analogous complexes with 1,2,4-triazole moieties, 45, 48, and 49–54, having enhanced colour purity to blue.
Bryce et al. introduced F-substituted 2,3′-bipyridine derivatives rather than ppy derivatives as the main ligand, to lower the HOMO level and develop efficient deep-blue phosphors, i.e., 58 and 59 (Fig. 11).110 Both of these complexes show intense blue phosphorescence emission at 437 and 435 nm, respectively, with high PLQYs exceeding 0.65. When implemented in OLED devices, the 59-based device exhibited 13% EQE with a deep blue colour (CIEx,y = 0.16, 0.17).
Recently, our group investigated the role of the ancillary ligand on the emission behaviours.111 A series of heteroleptic Ir(III) complexes, 60–62, were prepared, comprised of 2-(2,4-difluoro-3-(trifluoromethyl)phenyl)-4-methylpyridine (dfCF3) as the main ligand and the following different ancillary ligands: acetylacetonate (60), picolinate (61), and tetrakis-pyrazolyl borate (62). The heteroleptic Ir(III) complexes of 60–62 exhibit emission peaks at 470, 455, and 450 nm, respectively, in dichloromethane solution. For 60–62, the Huang–Rhys factors are estimated to be 0.76, 0.87, and 0.97, respectively, as shown in Fig. 12a. Interestingly, however, the NR rate constants are in the order 60 (4.89 × 105 s−1) > 61 (1.17 × 105 s−1) > 62 (0.28 × 105 s−1). To explain this phenomenon, we measured the temperature-dependent τem to estimate the activation barrier from the radiative triplet state to NR state, 3MC. The activation barriers were calculated as 46, 61, and >100 meV for 60–62, respectively. Theoretical quantum chemical calculations were also conducted to support the experimental data. According to these calculations, the activation barrier for 62 (7.9 kcal mol−1) is significantly higher than those of 60 (1.7 kcal mol−1) and 61 (2.9 kcal mol−1), as depicted in Fig. 12b. We suggested that reorganisation to a trigonal bipyramidal geometry may be difficult because of the steric demands of the borate ligand. Further, the bulkiness of the borate ligand may restrict the free rotation of the pyrazolyl group. Thus, the activation barrier to 3MC for 62 is much higher than those of 60 and 61. Accordingly, a 62-based blue phosphorescent OLED device exhibited the best performance among the series, with high current and power efficiencies of 32.9 cd A−1 and 25.4 lm W−1, respectively. Available photophysical- and electrochemical data of 49–62 with their device performances are summarised in Table 3.
Fig. 12 (a) Energy potential curves with Huang–Rhys factors for 60–62 and (b) reaction profiles for Ir–N bond cleavage to form NR states for 62. Adapted with permission from ref. 111. Copyright 2009, The Royal Society of Chemistry. |
Photophysical properties | Oxidation (V) | Device performances | Ref. | |||||||
---|---|---|---|---|---|---|---|---|---|---|
λ em (nm) | τ em (μs) | PLQY | k r (s−1) | k nr (s−1) | EL (λmax, nm) | CIE (x, y) | EQEmax (%) | |||
49 | 463, 492 | 0.1 | 0.03 | 2.0 × 105 | 2.8 × 105 | 0.91 | — | — | — | 108 |
50 | 464, 492 | 0.1 | 0.01 | 1.2 × 105 | 0.8 × 105 | 0.91 | — | — | — | 108 |
51 | 458, 487 | 0.2 | 0.04 | 3.0 × 105 | 1.6 × 106 | 1.02 | 461, 490 | (0.17, 0.27) | 7.4 | 108 |
52 | 460, 490 | 0.1 | 0.04 | 2.7 × 105 | 1.4 × 106 | 0.99 | — | — | — | 108 |
53 | 459, 488 | 0.2 | 0.04 | 2.9 × 105 | 7.1 × 105 | 0.99 | — | — | — | 108 |
54 | 459, 489 | 0.1 | 0.03 | 2.1 × 105 | 4.2 × 105 | 0.97 | — | — | — | 108 |
55 | 460, 489 | 1.2 | 0.32 | 2.6 × 105 | 5.6 × 105 | 0.98 | 495 | (0.18, 0.40) | 0.4 | 109 |
56 | 457, 487 | 1.3 | 0.32 | 2.3 × 105 | 5.3 × 105 | 1.02 | — | — | — | 109 |
57 | 458, 487 | 1.1 | 0.27 | 2.5 × 105 | 6.8 × 105 | 0.99 | 495 | (0.17, 0.40) | 0.48 | 109 |
58 | 437, 466 | 3.0 | 0.65 | 2.2 × 105 | 1.2 × 105 | — | 430–440 | (0.15, 0.13) | 11.2 | 110 |
59 | 435, 464 | 3.0 | 0.70 | 2.3 × 105 | 1.0 × 105 | — | 460–470 | (0.14, 0.11) | 13.0 | 110 |
60 | 470, 494 | 0.9 | 0.56 | 6.2 × 105 | 4.9 × 105 | 0.88 | — | (0.14, 0.26) | 19.9 | 111 |
61 | 455, 484 | 1.8 | 0.79 | 4.4 × 105 | 1.2 × 105 | 1.12 | — | (0.14, 0.18) | 15.5 | 111 |
62 | 450, 478 | 4.6 | 0.87 | 1.9 × 105 | 0.3 × 105 | 1.18 | — | (0.14, 0.20) | 22.6 | 111 |
Later, Da Como et al. prepared CN-substituted pmb-based Rh, Pt, and Ir carbene complexes to investigate the role of the SOC and ΔEST in controlling the radiative phosphorescence rate.132 The CN-substituted Ir(III) complex, 67 (Fig. 14), exhibits enhanced PLQY up to 0.78, much higher than 66 (note that the PLQY of 66 was later corrected from 0.04 to 0.37).57 This idea was extended to heteroleptic Ir(III) complexes with F-coordinated benzyl carbene main ligands and a 2-pyridyl triazolate ancillary ligand by Wu et al., who reported complexes 68–70 (Fig. 14).133 Those researchers observed significant blue-shifted emission through F-atom attachment, along with higher PLQY with insertion of a saturated methylene spacer in the main ligand. The enhanced PLQY was rationalised by estimating different NR decay rate constants for the reported complexes. Among the considered complexes, 70 was applied as a dopant in a blue OLED device. CIE coordinates of (0.158, 0.128) with an EQE of up to 6.0% were obtained. However, this value dropped to 2.7% at a practical brightness of 100 cd m−2. The results of a further theoretical investigation revealed remarkably destabilised HOMO energies through introduction of a methylene spacer in 69 and 70; thus, the HOMO–LUMO energy gaps in these complexes were increased.134
Kido et al. reported another carbene complex, 71, that exhibited blue emission with λmax at 445 nm.135 When 71 was doped with 3,6-bis(diphenylphosphoryl)-9-phenylcarbazole at 10 wt% (PO9, host) in film, high ηPL values of 0.70 were obtained with a τp of 19.6 μs at RT. This behaviour was compared with 1 and it was concluded that the strong ligand field effect in the carbene complex induces significant shifts in the d-orbital energies, which then facilitates high PLQY and longer lifetime. Additionally, 71 was combined with other red and green phosphors to fabricate a white OLED; the device exhibited ηp,max and ηp,1000 values of 59.9 and 43.3 lm W−1, respectively. Karatsu et al. systematically studied substitution effects on Ir(pmb)3 (66) moieties, 72–75 (Fig. 14), using electron-donating (–OCH3) or withdrawing (–CF3 and –CN) groups on the phenyl ring.136 The luminescent properties were found to be greatly affected by the functional group on the phenyl moiety, whereas the geometries and electrochemical properties remained within a similar range.
In 2013, De Cola et al. first reported efficient deep-blue-emissive cationic bi-pincer Ir(III) carbene complexes, 76 and 77 (Fig. 14), which were obtained using a pincer-type ligand, (4,6-dimethyl-1,3-phenylene-κC2)bis(1-butylimidazol-2-ylidene), with I− and PF6− as counter anions.137 Both complexes exhibit vibronic progression from the 3LC/3MLCT excited states with two main emission maxima at 394 and 406 nm. The PLQYs of 76 and 77 in solution are higher than those of the other reported bis-tridentate Ir(III) complexes at 0.41 and 0.38, respectively, because of the presence of strong-field ligands and a more rigid tridentate system. Interesting features are found in the solid state. The crystals of 76 exhibit a large, dominant redshifted emission at 500 nm, whereas those of 77 exhibit dual emission with the original high-energy emission and an additional low-energy emission at approximately 500 nm. The effect of the counter anion is remarkably apparent in the amorphous film. At a 50% doping ratio, significant low-energy emission appears for 76; however, 77 exhibits only a minor band in this region. Although the fabricated preliminary device exhibited very low efficiency (<1%), the EL spectrum showed saturated blue emission with maxima at 386 and 406 nm.
A series of N-heterocyclic carbene Ir(III) complexes, 78–85 (Fig. 14), having dfppy as the main ligand and NHCs as ancillary ligands, were systematically studied by Zuo et al. in 2015.138 By modifying the phenyl moiety in the NHCs with electron-withdrawing substituents or replacing the phenyl ring with N-heteroaromatic rings, the HOMO–LUMO gaps were increased and the emissions were blue-shifted accordingly. Among this series of carbene-based Ir(III) complexes, the 82-based blue phosphorescent OLED device exhibits good performance, with a maximum ηc of 37.83 cd A−1, an EQE of 10.3%, and an Lmax of 8709 cd m−2. More recently, Wong et al., reported heteroleptic iridium(III) complexes with NHC and acidic pyrazolyl-pyridine (fppz) moieties as ancillary ligands, 86 and 87, respectively.139 These ancillary ligands successfully maintained the relatively high LUMO energy levels, because the LUMOs of these Ir(III) complexes are almost completely located on the antibonding π* orbital of the pyridyl ring of the main ligand, F4ppy. Thus, sky-blue emissions at a wavelength of around 465 nm with PLQYs of over 0.60. Accordingly, OLED devices were fabricated which afford high luminance values of over 10000 cd m−2 at a driving voltage of 10 V with peak current efficiencies of 47.6 and 45.5 cd A−1, corresponding to EQEs of 20.6% and 19.6% for 86-based and 87-based devices, respectively.
In 2016, impressive device efficiency enhancement using carbene-based Ir(III) complexes was reported by the Forrest group.140N-Phenyl, N-methylpyridoimidazole was used as a cyclometalating ligand and both fac- (88) and mer- (89) isomers of Ir(pmp)3 were prepared (Fig. 14). Although their absorption and emission a slightly red-shifted compared to 66, from 380 to 418 nm for 88 and to 465 nm for 89, both complexes exhibit significantly higher PLQYs (≈0.77) compared to 66 (0.37).57 This enhancement is due to the N atom in the pyridoimidazol-2-yl moiety, which modulates the LUMO energy level. As a result, the emissive triplet state is stabilised, and a concomitant increase is obtained in the energy gap between the emissive T1 state and the non-emissive 3MC state. Notably, both complexes show enhanced device performance. That is, the 88-based device achieves a remarkable reduction in efficiency roll-off at high current density with high luminance (EQE = 10.1%), along with a deep-blue colour coordinate of (0.16, 0.09). The 89-based device shows even higher efficiency with a high EQE of 14.4% and high luminance of >22000 cd m−2. In particular, 88 can serve as the HTL and EBL material simultaneously. Available photophysical- and electrochemical data of 63–103 with their device performances are summarised in Table 4.
Photophysical properties | Oxidation (V) | Device performances | Ref. | |||||||
---|---|---|---|---|---|---|---|---|---|---|
λ em (nm) | τ em (μs) | PLQY | k r (s−1) | k nr (s−1) | EL (λmax, nm) | CIE (x, y) | EQEmax (%) | |||
a Measured at 77 K. b Doped film. | ||||||||||
63 | 414a | 14.0 | — | — | — | 0.41 | — | — | — | 131 |
64 | 480 | 37.0 | 0.38 | 0.1 × 105 | 0.2 × 105 | 0.31 | — | — | — | 131 |
65 | 380a | 0.4 | 0.02 | 0.5 × 105 | 2.0 × 106 | 0.22 | — | — | — | 131 |
66 | 389 | 0.22 | 0.04 | 1.8 × 105 | 4.3 × 106 | — | — | — | — | 131 |
67 | 380 | — | 0.78 | 4.9 × 104 | 1.3 × 104 | — | — | — | — | 132 |
68 | 392, 461 | 0.001 | <0.001 | 6.0 × 105 | 1.2 × 109 | — | — | — | — | 133 |
69 | 460 | 0.22 | 0.22 | 1.0 × 106 | 3.5 × 106 | — | — | — | — | 133 |
70 | 458 | 0.38 | 0.73 | 1.9 × 106 | 7.0 × 105 | — | 434, 460 | (0.16, 0.13) | 6.0 | 133 |
71 | 445 | 19.6b | 0.70b | 0.4 × 106 | 0.2 × 105 | — | — | — | — | 135 |
72 | 390, 407 | 1.3 | 0.44 | 3.4 × 10−5 | 4.3 × 10−5 | 0.45 | — | — | — | 136 |
73 | 396, 416 | 6.1 | 0.84 | 1.4 × 10−5 | 0.3 × 10−5 | 0.74 | — | — | — | 136 |
74 | 521, 445 | 14.0 | 0.71 | 0.5 × 10−5 | 0.2 × 10−5 | 0.84 | — | — | — | 136 |
75 | 403, 415 | 5.0 | 0.76 | 1.5 × 10−5 | 0.5 × 10−5 | 0037 | — | — | — | 136 |
76 | 384, 406 | 8.9 | 0.41 | 4.6 × 10−5 | 6.6 × 10−5 | 0.68 | — | — | — | 137 |
77 | 384, 406 | 9.4 | 0.38 | 4.1 × 10−5 | 6.5 × 10−5 | 0.65 | 386, 406 | — | — | 137 |
78 | 483 | 2.1 | 0.14 | 0.6 × 105 | 3.7 × 105 | — | — | — | — | 138 |
79 | 483 | 1.8 | 0.65 | 3.6 × 105 | 1.9 × 105 | — | — | — | — | 138 |
80 | 473 | 1.8 | 0.73 | 4.1 × 105 | 1.5 × 105 | — | — | — | — | 138 |
81 | 469 | 1.9 | 0.57 | 3.0 × 105 | 0.2 × 105 | — | — | — | — | 138 |
82 | 469 | 1.8 | 0.69 | 3.8 × 105 | 1.7 × 105 | — | 470 | — | 10.3 | 138 |
83 | 473, 498 | 1.8 | 0.61 | 3.4 × 105 | 2.2 × 105 | — | — | — | — | 138 |
84 | 471, 497 | 1.7 | 0.33 | 1.9 × 105 | 3.9 × 105 | — | — | — | — | 138 |
85 | 455, 479 | 1.9 | 0.32 | 1.7 × 105 | 3.6 × 105 | — | — | — | — | 138 |
86 | 469, 462 | 17.6 | 0.60 | 0.3 × 105 | 0.2 × 105 | 465 | (0.19, 0.39) | 20.6 | 139 | |
87 | 462, 459 | 16.6 | 0.68 | 0.4 × 105 | 0.2 × 105 | — | (0.19, 0.37) | 19.6 | 139 | |
88 | 418 | 1.2 | 0.76 | 6.4 × 105 | 2.0 × 105 | 0.23 | — | — | — | 140 |
89 | 465 | 0.8 | 0.78 | 1.0 × 106 | 2.7 × 105 | — | — | — | — | 140 |
90 | 430 | 3.8 | 0.98 | 2.6 × 105 | 0.5 × 104 | — | — | — | — | 142 |
91 | 454 | 4.5 | 0.85 | 1.9 × 105 | 0.3 × 105 | — | — | (0.15, 0.19) | 7.6 | 142 |
92 | 468 | 4.3 | 0.99 | 2.3 × 105 | 0.2 × 104 | — | — | (0.15, 0.19) | 10.8 | 142 |
93 | 469 | 4.7 | 0.45 | 1.0 × 105 | 1.2 × 105 | — | 450 | (0.15, 0.19) | 15.2 | 142 |
94 | 422 | 5.3 | 0.33 | 0.6 × 105 | 1.2 × 105 | — | — | — | — | 142 |
95 | 515 | 0.17 | 0.03 | 1.8 × 105 | 5.9 × 106 | 0.43 | — | — | — | 143 |
96 | 555 | 0.0037 | 0.001 | 2.7 × 105 | 2.7 × 108 | 0.29 | — | — | — | 143 |
97 | 444, 472 | 11.2 | 0.68 | 6.1 × 104 | 2.9 × 104 | (0.14, 0.11) | 18.5 | 144 | ||
98 | 450, 477 | 11.0 | 0.53 | 4.8 × 104 | 4.1 × 104 | (0.14, 0.14) | 18.2 | 144 | ||
99 | 414, 424 | 0.3, 1.8 | 0.25 | — | — | 0.88 | — | — | — | 121 |
100 | 412, 427 | 0.7, 1.8 | 0.72 | — | — | 0.80 | 431 | (0.15, 0.08) | 7.2 | 121 |
101 | 418b | 6.1b | 0.13b | 0.2 × 10−5 | 1.4 × 10−5 | 0.16 | — | — | — | 145 |
102 | 418b | 1.8b | 0.31b | 1.7 × 10−5 | 3.8 × 10−5 | 0.21 | — | — | — | 145 |
103 | 459b | 0.9b | 0.48b | 5.6 × 10−5 | 6.1 × 10−5 | 0.25 | — | — | — | 145 |
Zhou and Powell later calculated the possible reaction pathways of 88 and 66 to the NR 3MC states, the correlation between the lowest activation barrier to the 3MC states, and the experimental NR rate constants.141 According to their calculations, the Ir–C bond of the NHC ligand is mostly elongated in the excited state for both isomers. In addition, those authors calculated the key parameters of the T1 states: (1) the Franck–Condon state in the ground-state, (2) the optimised 3MLCT state, (3) an intermediate 3MC state, and (4) the optimised 3MC state. When a benzannulated component of the NHC ligands is replaced with a fused pyridyl ring, transitioning from 66 to 88, the energy barrier to the NR state is increased. Zhou and Powell explained that the presence of the N atom with greater electronegativity may strengthen the Ir–C (NHC) bond and generate a higher energy barrier in the 3MC state, yielding a slower knr for 88. More importantly, they found that the metal–ligand bond in the 3MC state does not break away and is reversible. The PLQY enhancement of 88 may be due to this unique property.
In 2017, Wong et al. prepared a series of N-heterocyclic carbene Ir(III) complexes, 90–94, to study the effect of electron-withdrawing/donating nature of the substituent on the phenyl ring by modifying 89.142 The device fabricated using 94 (Fig. 14) exhibited the maximum EQE, at 19.0% with restricted efficiency roll-off. Recently, Kang et al. modified the phenyl ring of 88 with a bulkier substituent, a xylyl ring at the ortho positions of pmb, to investigate the NR decay pathway according to the ligand bulkiness.143 The detailed photo-dynamics in the excited states were studied using transient absorption and time-resolved emission techniques in a series of Ir(III) complexes. When the bulkier ligand was attached, the PLQYs of both the fac- (95, Fig. 14) and mer- (96) isomers were anomalously quenched in the solution at 300 K. They found new, broad TA bands for 95 and 96 at approximately 720 nm with increasing delay time; this band was associated with structural changes in the excited triplet state which yielded fast localisation migration via inter-ligand charge transfer and influenced the deactivation pathway to quench the emission intensity.
N-Dibenzofuranyl-N-methylimidazole-based Ir(III) carbene complexes with fac- (97) and mer- (98, Fig. 14) isomers was also reported by our group.144 For the previously reported Ir(III) carbene complexes, the mer-isomer is equally or more efficiently luminescent than the fac-isomer in solution and in the solid state;131,140 however, the fac-isomer labelled 97 exhibits better luminescence properties in solution and device application than the mer-isomer labelled 98. We rationalised this difference using DFT calculations, which revealed that the energy barrier from T1 to 3MC of 98 is lower than that of 97. Among the two OLED devices, the 97-based device showed a higher EQE value (18.5%) than the 98-based device (18.2%). Furthermore, the CIEx,y for the 97-based device showed a deeper blue coordination of (0.14, 0.11) compared to the 98-based device with (0.14, 0.14).
Zysman-Colman et al. reported two kinds of –CF3 functionalised mer-Ir(pmi)3 complexes, 99 and 100 (Fig. 14), in which the substitution position of –CF3 on the phenyl ring was varied to obtain a high PLQY through sufficient stabilisation of the HOMO energy level.121 Both complexes show structured deep-blue emission (λmax ≈ 425 nm) in degassed dichloromethane at RT. However, the PLQYs are significantly affected depending on the substitution position. For 99, the PLQY is relatively low at 0.25, whereas that of 100 is considerably higher at 0.72, being comparable to 89. An optimised device with 100 as a blue dopant and 99 as an efficient exciton/electron blocker exhibited deep-blue CIE coordinates of (0.154, 0.052), with an EQEmax of 13.4% and an EQE of 12.5% at 100 cd m−2.
Most recently, Teets et al. reported a new type of Ir(III) carbene complex with an acyclic diaminocarbene (ADC) as an ancillary ligand, 101–103 (Fig. 14), referred to as a ‘mixed-carbene complex’.145 These mixed-complexes were prepared through a cascade reaction with nucleophilic addition reaction followed by base-assisted cyclometalation reaction of the ADC intermediate. Compared to homoleptic Ir(III) carbene complexes, the mixed-carbene complexes show the same or even better photoluminescence characteristics; for example, higher PLQY. As ADC is an even stronger σ-donor ligand than NHCs, because of its greater 2p in its s orbital, the molecular design strategy for saturated blue emissive material may be extended following this approach.
Following the above works, extensive studies were subsequently performed to develop a blue-emissive bis-tridentate Ir(III) complex. In 2016, the Chi group reported cooperation between bis(imidazolylidene) benzenes and functionalised 2-pyrazolyl-6-phenyl pyridine (or isoquinoline) chelates ligands to form bis-tridentated Ir(III) complexes, 111–118 (Fig. 15).155 Although the most blue-shifted emission among this series is at 467 nm for solution 111, the methyl-substituted complexes exhibit almost unitary PLQY whereas the isopropyl-substituted complexes exhibit lower PLQY. This difference was rationalised by considering the reduction in the torsional vibration degrees of freedom for the methyl groups. The researchers also noted that increases in the rigidity and multidentate coordination mode are key factors yielding the excellent PL and EL efficiency obtained for these complexes. Depending on the π-conjugation and/or electronic characteristics of the chelates, the emission colour can be finely tuned. Accordingly, the authors successfully fabricated efficient, independent red, green, and blue OLED devices. They also obtained a white OLED device by combining these dopants in the emitting layer. For example, 95% doping in a 9-(3-(9H-carbazol-9-yl)phenyl)-9H-carbazole-3-carbonitrile (mCPCN) host yielded blue-emitting OLEDs with an EQE of 27% and CIEx,y of (0.18, 0.40). To enlarge the HOMO–LUMO energy band gap of this sky-blue dopant, the same group designed new Ir(III) complexes in which the 6-pyrazolyl-2-ppy was replaced with 6-pyrazolyl-2-phenoxylpyridine to break the π-conjugation between the pyridyl and fluorophenyl units.156 As expected, all emission spectra of these 6-pyrazolyl-2-phenoxylpyridine based complexes, 119–121(Fig. 15), are clearly blue-shifted versus the emission spectra for the 6-pyrazolyl-2-ppy-based Ir(III) complexes, 111–116 (Fig. 15). Note that deep-blue EL with an EQE of >20% and CIEx,y = (0.15, 0.17) were achieved using 121 as a dopant material. Available photophysical- and electrochemical data of 104–132 with their device performances are summarised in Table 5.
Photophysical properties | Oxidation (V) | Device performances | Ref. | |||||||
---|---|---|---|---|---|---|---|---|---|---|
λ em (nm) | τ em (μs) | PLQY | k r (s−1) | k nr (s−1) | EL (λmax, nm) | CIE (x, y) | EQEmax (%) | |||
104 | 506 | 9.5 | 0.11 | 0.1 × 105 | 1.0 × 105 | >1.7 | — | — | — | 153 |
105 | 585 | 3.9 | 0.21 | 3.4 × 105 | 1.2 × 106 | — | — | — | — | 152 |
106 | 560 | — | — | — | — | — | — | — | — | 152 |
107 | 550, 592 | 1.0 | 0.04 | 0.4 × 105 | 8.9 × 106 | 1.65 | — | — | — | 153 |
108 | 547, 582 | 5.7 | 0.19 | 0.3 × 105 | 1.4 × 106 | 1.64 | — | — | — | 153 |
109 | 593, 623 | 1.6 | 0.10 | 0.6 × 105 | 5.7 × 105 | 1.18 | — | — | — | 153 |
110 | 509, 541 | 1.2 | 0.04 | 0.4 × 105 | 9.1 × 105 | >1.7 | — | — | — | 153 |
111 | 467, 501 | 5.4 | 0.99 | 1.8 × 105 | 0.1 × 104 | — | — | (0.18, 0.40) | 27.0 | 155 |
112 | 468, 503 | 6.7 | 0.86 | 1.3 × 105 | 0.2 × 105 | — | — | (0.19, 0.39) | 20.8 | 155 |
113 | 490, 526 | 4.7 | 0.92 | 2.0 × 105 | 0.2 × 105 | — | — | (0.29, 0.57) | 30.0 | 155 |
114 | 492, 529 | 4.4 | 0.80 | 1.8 × 105 | 0.5 × 105 | — | — | (0.28, 0.57) | 28.3 | 155 |
115 | 501, 535 | 2.8 | 1.00 | 3.6 × 105 | — | — | — | (0.34, 0.60) | 31.4 | 155 |
116 | 503, 537 | 3.0 | 0.98 | 3.3 × 105 | 0.7 × 104 | — | — | (0.34, 0.60) | 30.0 | 155 |
117 | 593, 643 | 8.2 | 1.00 | 1.2 × 105 | — | — | — | (0.63, 0.38) | 27.4 | 155 |
118 | 595, 647 | 6.9 | 0.98 | 1.4 × 105 | 0.3 × 104 | — | — | (0.63, 0.38) | 20.0 | 155 |
119 | 471 | 25.1 | 0.81 | 0.3 × 105 | 0.7 × 104 | 0.67 | — | — | — | 156 |
120 | 478 | 4.42 | 0.82 | 1.9 × 105 | 0.4 × 105 | 0.53 | — | (0.15, 0.24) | 19.7 | 156 |
121 | 472 | 8.66 | 0.72 | 0.8 × 105 | 0.3 × 105 | 0.52 | — | (0.15, 0.17) | 20.7 | 156 |
122 | 473 | 61.2 | 0.68 | 0.1 × 105 | 0.5 × 104 | 0.80 | — | — | — | 156 |
123 | 473 | 18.6 | 0.79 | 0.4 × 105 | 0.1 × 105 | 0.62 | — | — | — | 156 |
124 | 486 | 2.8 | 1.00 | 3.6 × 105 | — | 0.55 | 484 | (0.19, 0.34) | 19.6 | 157 |
125 | 473 | 3.2 | 0.84 | 2.6 × 105 | 0.5 × 105 | 0.69 | 468 | (0.17, 0.25) | 21.6 | 157 |
126 | 476 | 2.7 | 0.83 | 3.1 × 105 | 0.6 × 105 | 0.68 | 472 | (0.17, 0.26) | 19.6 | 157 |
127 | 506 | 2.59 | 0.92 | 7.7 × 10−5 | — | 0.48 | — | — | — | 158 |
128 | 489 | 4.43 | 0.94 | 8.2 × 10−5 | — | 0.75 | — | — | — | 158 |
129 | 491 | 3.89 | 1.00 | 9.2 × 10−5 | — | 0.59 | — | — | — | 158 |
130 | 515 | 0.93 | 0.96 | 2.8 × 10−5 | — | 0.37 | — | — | — | 158 |
131 | 473 | 1.52 | 0.97 | 3.3 × 10−5 | — | 0.67 | — | — | — | 158 |
132 | 477 | 1.53 | 1.00 | 3.7 × 10−5 | — | 0.62 | — | — | — | 158 |
Next, the Chi group employed a carbazolyl unit to replace the phenoxy group.157 The resultant carbazole-based Ir(III) complexes, 124–126 (Fig. 15), exhibit structureless emission bands indicating the dominant charge transfer contribution in the emissive process, with maxima at 486, 473, and 476 nm, respectively. The PLQY is almost unitary for 124, and slightly lower, at approximately 83%, for both 125 and 126 in solution at RT. Surprisingly, the radiative lifetimes (τrad) of 124–126 are significantly shorter (2.77–3.80 μs) than those of the previously reported Ir(III) complexes, 111 and 121, (τrad = 5.41 and 12.0 μs). This may be advantageous for application of these materials in device structures, because it may reduce the population density of the long-lived triplet excitons. Importantly, the Chi group also conducted photo-degradation experiments on the complexes in deaerated toluene under irradiation. The carbazole-based bis-tridentate Ir(III) complexes exhibits superior photostability to the tri-bisdentate deep-blue Ir(III) carbene complexes, 88 and 89. Furthermore, the rate of photodegradation of the carbazole-based Ir(III) complexes is lower than those of 111 and 121, indicating that the six-membered N-containing metallacycle in these complexes is more stable than the corresponding five-membered metallacycle system as a result of the robust chelating framework. More recently, improved photo-stabilities with shortened τem were obtained by replacing the central pyridine unit with a pyrimidine unit, 127–132 (Fig. 15).158 Various methodologies for fine-tuning the electronic transitions of bis-tridentate Ir(III) complexes for high performance and durable phosphorescent OLEDs are being developed at present.159–162
Recently, Bryce et al., reported a series of diarylhydrazide-bridged diiridium complexes functionalised with ppy-based cyclometalating ligands, 133–141 (Fig. 16).171 Among them, 141 exhibits sky-blue emission in doped film with λmax = 460 nm; this value indicates 10 nm hypsochromic shifting compared to 1. This complex also exhibits high PLQY (0.69), high kr (4.26 × 105 s−1), and relatively short τp (2.24 μs) in doped film. A combination of X-ray molecular structure analyses and NMR studies have revealed that the intramolecular π–π interactions between the bridging and cyclometalating ligands rigidify the complexes and play an important role in the observed, excellent photophysical properties. Bryce et al. continued their work using hydrazide-bridged diiridium complexes, 142–145 (Fig. 16).172 Complexes 143–145 are strongly emissive (PLQYs = 0.47–0.55) when doped into PMMA, whereas complex 142 exhibits a relatively small PLQY of 0.11. Those researchers interpreted these findings as indicating that the intramolecular π–π interactions are promoted by introduction of F substituents to the phenyl rings of the bridging ligand. In 2018, Chi et al. reported sublimable diiridium complexes bearing both functional 2-pyrazolyl-6-phenyl pyridine chelate and bidentate phenyl imidazolylidene chelate, 146–149 (Fig. 16).173 Among them, 149 exhibits blue-shifted emission peaks at 468, 500, and 536 nm due to the combined inductive and mesomeric effects. Importantly, very high PLQYs of >0.90 were observed for all diiridium complexes in both solution and film states at RT. Accordingly, the resulting OLED device using 149 showed impressive blue-emitting characteristics with an EQE of 14.2%. It should be noted OLED devices with 146 and 147 showed much higher EQEs of 18.3% (yellowish green colour) and 27.6% (green colour), respectively. These investigations of diiridium complexes offer a new method for future development of deeper-blue emitting Ir(III) complexes. Available photophysical- and electrochemical data of 133–149 with their device performances are summarised in Table 6.
Photophysical properties | Oxidation (V) | Device performances | Ref. | |||||||
---|---|---|---|---|---|---|---|---|---|---|
λ emmax (nm) | τ em (μs) | PLQY | k r (s−1) | k nr (s−1) | EL (λmax, nm) | CIE (x, y) | EQEmax (%) | |||
a Doped film. | ||||||||||
133 | 516a | 1.8a | 0.61a | 3.4 × 105 | 2.2 × 105 | 0.42, 0.67 | — | — | — | 171 |
134 | 503a | 2.0a | 0.59a | 3.0 × 105 | 2.1 × 105 | 0.52, 0.77 | — | — | — | 171 |
135 | 503a | 2.1a | 0.71a | 3.4 × 105 | 1.4 × 105 | 0.56, 0.81 | — | — | — | 171 |
136 | 507a | 2.0a | 0.66a | 3.3 × 105 | 1.7 × 105 | 0.50, 0.76 | — | — | — | 171 |
137 | 507a | 11.4a | 0.72a | 5.2 × 105 | 2.0 × 105 | 0.58, 0.90 | — | — | — | 171 |
138 | 504a | 1.1a | 0.66a | 5.8 × 105 | 3.0 × 106 | 0.62, 0.78 | — | — | — | 171 |
139 | 470 | 1.2 | 0.65a | 5.5 × 105 | 3.0 × 105 | 0.93, 1.28 | — | — | — | 171 |
472a | 1.2a | 0.60a | 5.5 × 105 | 3.4 × 105 | 0.97, 1.33 | |||||
140 | 471a | 1.1a | 0.46a | 4.1 × 105 | 4.8 × 105 | 0.95, 1.12 | — | — | — | 171 |
141 | 460a | 1.6a | 0.69a | 4.3 × 105 | 2.0 × 105 | 0.81, 1.18 | — | — | — | 171 |
142 | 500a | 1.8a | 0.11a | 0.6 × 105 | 4.9 × 105 | 0.34, 0.60 | — | — | — | 172 |
143 | 501a | 2.8a | 0.55a | 2.0 × 105 | 1.6 × 105 | 0.30, 0.67 | — | — | — | 172 |
144 | 486a | 4.2a | 0.47a | 1.1 × 105 | 1.3 × 105 | 0.49, 0.89 | — | — | — | 172 |
145 | 480a | 4.6a | 0.52a | 1.1 × 105 | 1.1 × 105 | 0.68, 1.12 | — | — | — | 172 |
146 | 531 | 3.0 | 0.98 | 3.3 × 105 | 0.1 × 105 | 0.64, 0.94 | 530 | 0.36, 0.61 | 18.3 | 173 |
147 | 487, 521, 559 | 4.6 | 0.95 | 2.1 × 105 | 0.1 × 105 | 0.50, 0.78 | 489 | 0.19, 0.53 | 27.6 | 173 |
148 | 485, 519, 554 | 7.0 | 0.64 | 0.9 × 105 | 0.5 × 105 | 0.58, 0.89 | 173 | |||
149 | 531 | 2.9 | 0.91 | 3.1 × 105 | 0.3 × 105 | 0.61, 0.92 | — | — | — | 173 |
Photophysical properties | Oxidation (V) | Device performances | Ref. | |||||||
---|---|---|---|---|---|---|---|---|---|---|
λ em (nm) | τ em (μs) | PLQY | k r (s−1) | k nr (s−1) | EL (λmax, nm) | CIE (x, y) | EQEmax (%) | |||
150 | 468, 494 | 3.2 | 0.94 | 3.0 × 10−5 | 0.3 × 10−5 | — | — | (0.15, 0.35) | 18.1 | 174 |
151 | 501, 526 | 0.2 | 0.20 | 9.1 × 10−5 | 3.6 × 10−4 | — | — | — | — | 174 |
152 | 469, 495 | 0.9 | 0.72 | 7.7 × 10−5 | 3.0 × 10−5 | — | — | — | — | 174 |
153 | 506, 529 | 0.2 | 0.15 | 9.4 × 105 | 5.3 × 10−4 | — | — | — | — | 174 |
154 | 472 | 14.9 | 0.70 | 4.7 × 104 | 2.0 × 104 | — | — | — | 175 | |
155 | 533 | 18.3 | 0.77 | 4.2 × 104 | 1.3 × 104 | — | — | — | 175 |
This journal is © the Partner Organisations 2020 |