Skip to content
Publicly Available Published by De Gruyter June 26, 2015

Highly efficient electroluminescence from purely organic donor–acceptor systems

  • Katsuyuki Shizu , Jiyoung Lee , Hiroyuki Tanaka , Hiroko Nomura , Takuma Yasuda , Hironori Kaji and Chihaya Adachi EMAIL logo

Abstract

Thermally activated delayed fluorescence (TADF) emitters are third-generation electroluminescent materials that realize highly efficient organic light-emitting diodes (OLEDs) without using rare metals. Here, after briefly reviewing the principles of TADF and its use in OLEDs, we report a sky-blue TADF emitter, 9-(4-(benzo[d]thiazol-2-yl)phenyl)-N3,N3,N6,N6-tetraphenyl-9H-carbazole-3,6-diamine (DAC-BTZ). DAC-BTZ is a purely organic donor–acceptor-type molecule with a small energy difference between its lowest excited singlet state and lowest triplet state of 0.18–0.22 eV according to fluorescence and phosphorescence spectra of a DAC-BTZ-doped film. In addition, the doped film exhibits a high photoluminescence quantum yield of 0.82. Time-resolved photoluminescence measurements of the doped film confirm that DAC-BTZ emits TADF. An OLED containing DAC-BTZ as an emitter exhibits a maximum external quantum efficiency (EQE) of 10.3%, which exceeds those obtained with conventional fluorescent emitters (5–7.5%). TADF from DAC-BTZ makes a large contribution to the high EQE of its OLED.

Thermally activated delayed fluorescence and its application in organic light-emitting diodes

Gilbert N. Lewis, one of the most influential chemists in the 20th century, first measured thermally activated delayed fluorescence (TADF) [1]. Before Lewis, most dyes were known to exhibit two phosphorescent bands associated with different photophysical processes. The band lying at higher energy was called the alpha emission band, while the other was called the beta emission band. Alpha and beta emissions are now called TADF and phosphorescence, respectively. Jabłoński proposed that alpha emission involved thermal activation from a phosphorescent state to a normal (fluorescent) state, while beta emission was spontaneous emission from a phosphorescent state. His interpretation is supported by several experimental observations: the alpha emission band is indistinguishable from the fluorescence band of a material, and lies at higher energy than the beta emission band; the rate of alpha emission increases with increasing temperature, while that of beta emission is much less dependent on temperature; and only beta emission is observed at low temperatures. Lewis showed that the alpha emission of fluorescein in boric acid obeyed the Arrhenius equation, and experimentally determined the thermal activation energy to be 8 ± 1 kcal mol–1 [1]. This was the beginning of experimental study on TADF. Since the 1960s, TADF has been observed from various compounds such as eosin [2], benzophenone [3], aromatic thiones [4], and fullerenes [5, 6], suggesting that TADF emitters have diverse chemical structures.

In 2009, an attempt was made to use TADF emitters in organic light-emitting diodes (OLEDs) [7]. OLEDs have attracted considerable attention because of their potential application as flexible flat-panel displays and next-generation solid-state lighting sources [8–10]. In an OLED, excitons are generated by charge recombination, and singlet and triplet excitons are generated in a ratio of 1:3. Therefore, singlet and triplet excitons account for 25% and 75% of the total excitons, respectively. As a result, the electroluminescence (EL) efficiency of an OLED depends largely on the triplet-to-light conversion efficiency. TADF emitters can convert triplet states into light via thermal activation from triplet to singlet states, so they hold promise as potential emitters and singlet-exciton harvesters for OLEDs. In 2011, TADF-based OLEDs were shown to be able to outperform conventional fluorescent OLEDs [11]. Since this discovery, we have markedly improved the luminescence efficiency of TADF emitters and external quantum efficiency (ηEQE) of TADF-based OLEDs. In 2012, we developed the first purely organic TADF emitter that realized highly efficient OLEDs [12] with ηEQE of 19%, comparable to those of phosphorescent OLEDs. TADF is now a promising alternative to fluorescence (used in first-generation devices) and phosphorescence (used in second-generation devices), and is considered to be the third-generation technique to obtain highly efficient OLEDs [13]. To date, TADF emitters with wide ranges of both emission colors and chemical structures have been designed and synthesized [7, 11, 12, 14–55].

Recently, the use of TADF has greatly expanded, with developments including OLEDs with promising operational stability [56], white OLEDs for application as a solid-state lighting source [31, 57], assist dopants in highly efficient fluorescent OLEDs [58, 59], controlling emitter orientation to enhance light-outcoupling efficiency [60–62], solution-processed OLEDs [63, 64], a roll-to-roll process for low-cost OLEDs [65], highly efficient electrogenerated chemiluminescence [66], and time-resolved fluorescence cell imaging [67]. The recent progress in the application of TADF suggests that a paradigm shift from OLED dopants that exhibit normal fluorescence and phosphorescence to those that show TADF is currently occurring.

Photoluminescence and electroluminescence processes involving TADF

Figure 1 shows photoluminescence (PL) and EL processes involving TADF. We assume here that TADF involves the lowest excited singlet state (S1), the lowest triplet state (T1), and the ground state (S0). TADF can be viewed as involving S1 ← T1 reverse intersystem crossing (RISC) and S1 → S0 radiative decay (Fig. 1c,f). In the PL process, S1 excitons are generated after the photoexcitation of S0 (Fig. 1a). Of these S1 excitons, Φp is converted into light (S1 → S0 radiative decay), and 1 – Φp is converted to T1 via S1 → T1 intersystem crossing (Fig. 1b). Of the T1 excitons, Φd is finally converted into light as TADF (Fig. 1c). The ratio Φd/(1 – Φp) is the T1-to-light conversion efficiency. PL quantum yield (PLQY, ΦPL) is given by ΦPL = Φp + Φd, where Φp and Φd are prompt and delayed components of ΦPL, respectively, and can be determined from transient PL decay measurements [15]. For fluorescent compounds that do not emit TADF, Φd = 0.

Fig. 1: 
          Photoluminescence and electroluminescence processes involving TADF. S1 is the lowest excited singlet state, T1 is the lowest triplet state, and S0 is the ground state. Φp and Φd are prompt and delayed components of PLQY (ΦPL), respectively.
Fig. 1:

Photoluminescence and electroluminescence processes involving TADF. S1 is the lowest excited singlet state, T1 is the lowest triplet state, and S0 is the ground state. Φp and Φd are prompt and delayed components of PLQY (ΦPL), respectively.

In the EL process, S1 and T1 are generated in a ratio of 1:3, so S1 and T1 account for 25% and 75% of the total excitons, respectively (Fig. 1d). Next, 0.25 Φp decays radiatively, and 0.25(1 – Φp) is converted to T1, resulting in T1 of

(1) 0.75 + 0.25 ( 1 Φ p )  (1)

(Fig. 1e). Of the T1 excitons,

(2) [ 0.75 + 0.25 ( 1 Φ p ) ] Φ d 1 Φ p  (2)

is finally extracted as TADF. When Φd = 0, only 0.25 Φp can be extracted as light. This situation is true for conventional fluorescent OLEDs.

η EQE of a TADF-based OLED depends on charge recombination factor (γ) and light outcoupling efficiency (ηout) as well as Φp and Φd. γ is between 0 and 1, while ηout is considered to be 0.2–0.3. ηEQE can be written as

(3) η EQE = { 0.25 Φ p + [ 0.75 + 0.25 ( 1 Φ p ) ] Φ d 1 Φ p } γ η out .  (3)

The corresponding internal quantum efficiency (IQE) is expressed as ηEQE/ηout, so it depends on γ, Φp, and Φd:

(4) ( IQE ) = { 0.25 Φ p + [ 0.75 + 0.25 ( 1 Φ p ) ] Φ d 1 Φ p } γ .  (4)

When perfect charge balance is achieved (γ = 1) and nonradiative decay is completely suppressed (Φd = 1 – Φp), the IQE is 100% and ηEQE reaches 20–30%. In fact, ηEQE of 20–30% has been realized by optimizing device structure [61, 68–70]. For conventional fluorescent OLEDs, because Φd = 0, IQE and ηEQE are limited to 25% and 5–7.5%, respectively, even if Φp of nearly 1 is achieved using a highly efficient fluorescent emitter. Thus, TADF emitters can substantially improve IQE and ηEQE by effectively converting triplet excitons into light. Eqs. 3 and 4 are useful to evaluate the potential of TADF emitters in OLEDs. Once Φp and Φd of a TADF emitter are determined experimentally, by setting γ = 1 and using eqs. 3 and 4, we can calculate the maximum IQE and range of maximum ηEQE of TADF-based OLEDs. The contributions from prompt fluorescence (ηp) and TADF (ηd) to ηEQE can be calculated from

(5) η p = 0.25 Φ p γ η out ,  (5)

(6) η d = [ 0.75 + 0.25 ( 1 Φ p ) ] Φ d 1 Φ p γ η out .  (6)

η p can be interpreted as the EQE for conventional fluorescent OLEDs. For example, for a green TADF emitter with Φp = 0.34 and Φd = 0.34, the maximum IQE and ηEQE are calculated to be 56% and 11–17%, respectively [37]. ηp and ηd are calculated to be 2–3% and 9–14%, respectively. ηd is larger than ηp, suggesting that TADF contributes considerably to ηEQE.

Molecular design of TADF emitters

As stated above, TADF involves S1 ← T1 RISC and S1 → S0 radiative decay. Therefore, the efficiency of TADF depends largely on the rates of S1 ← T1 RISC (kRISC) and S1 → S0 radiative decay (kr). Increasing kRISC and kr is an effective approach to improve the luminescence efficiency of TADF emitters. Here, for simplicity, S1 and T1 are assumed to be dominated by one-electron excitation from the highest occupied molecular orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO). In such cases, ΔEST is approximately expressed as 2K, where K is the exchange integral between two electrons occupying the HOMO and LUMO. K is small when the HOMO and LUMO are spatially separated. Because kRISC increases as the energy difference between S1 and T1EST) decreases [5], compounds with well-separated HOMO and LUMO are expected to exhibit fast kRISC. In contrast, compounds with strongly overlapping HOMO and LUMO are expected to exhibit fast kr [40, 41]. This is because kr increases with increasing transition dipole moment between S0 and S1 (μ10), and μ10 increases with increasing HOMO-LUMO spatial overlap; more specifically, the overlap density between the HOMO and LUMO [40, 41]. Thus, there is a trade-off between fast kRISC and fast kr. However, by carefully controlling HOMO and LUMO distributions and the resultant spatial overlap between them, moderate kRISC and kr are compatible, so we can obtain highly efficient TADF emitters.

Most TADF emitters contain electron-donating and -accepting units. Intramolecular donor–acceptor systems have been reported to efficiently convert electricity into light and used in electrochemiluminescence systems [71–73]. In a donor–acceptor system, the HOMO is predominantly distributed on an electron-donating unit, while the LUMO is predominantly distributed on an electron-accepting unit. Therefore, a donor–acceptor system has spatially separated HOMO and LUMO, and consequently, a small ΔEST. When suitable electron-donating and -accepting units are combined, ΔEST is sufficiently small to facilitate efficient TADF. Simple and effective approaches to further separate the spatial overlap between HOMO and LUMO are to induce steric hindrance between electron-donating and -accepting units and increase the twist angle between them [11, 12, 21]. Both of these approaches can effectively decrease the HOMO-LUMO spatial overlap. Examples of electron-donating units that can induce large twist angles include phenoxazine [18–20, 30, 31, 34, 36, 38, 44], phenothiazine [35], phenazine [30, 37], and 9,9-dimethylacridane [30, 34]. Donor–acceptor systems bridged by a spiro structure also possess good HOMO-LUMO separation and small ΔEST [14, 15, 24, 33]. Recently, TADF processes involving RISC from higher excited triplet states to excited singlet states have been suggested [74–76]. This finding opens a new direction for molecular design of TADF emitters.

Sky-blue TADF emitter 9-(4-(benzo[d]thiazol-2-yl)phenyl)-N3,N3,N6,N6-tetraphenyl-9H-carbazole-3,6-diamine (DAC-BTZ)

Here, we report a novel sky-blue TADF emitter, 9-(4-(benzo[d]thiazol-2-yl)phenyl)-N3,N3,N6,N6-tetraphenyl-9H-carbazole-3,6-diamine (DAC-BTZ, Fig. 2a). The synthetic route to DAC-BTZ is outlined in scheme 1. DAC-BTZ is a purely organic donor–acceptor-type molecule, in which N3,N3,N6,N6-tetraphenyl-9H-carbazole-3,6-diamine (DAC-II) serves as an electron-donating unit and 2-phenylbenzo[d]thiazole (BTZ) acts as an electron-accepting unit. BTZ has good electron-accepting ability and appears in other TADF emitters [36, 37]. DAC-II has great potential as an electron-donating unit to realize highly efficient TADF emitters because it improves the compatibility between fast kRISC and kr, and enhances TADF [41]. By combining DAC-II and BTZ, we obtained the efficient sky-blue TADF emitter DAC-BTZ.

Fig. 2: 
          (a) Chemical structure, (b) optimized S0 geometry, (c) HOMO, and (d) LUMO of DAC-BTZ. The S0 geometry, HOMO, and LUMO were calculated at the M06-2X/cc-pVDZ level of theory.
Fig. 2:

(a) Chemical structure, (b) optimized S0 geometry, (c) HOMO, and (d) LUMO of DAC-BTZ. The S0 geometry, HOMO, and LUMO were calculated at the M06-2X/cc-pVDZ level of theory.

Figure 2b–d show structures of DAC-BTZ obtained from density functional theory calculations. All calculations were performed using the Gaussian 09 program package [77]. The S0 geometry of DAC-BTZ was optimized at the M06-2X/cc-pVDZ level of theory [78, 79]. BTZ is almost coplanar, while the outmost diphenylamino groups of DAC-II have a propeller-like structure (Fig. 2b). The calculated twist angle between DAC-II and BTZ is 50.6°, which is small compared with those of twisted donor–acceptor-type TADF emitters [14, 15, 18–20, 24, 30, 31, 33–38, 44]. Excited states of DAC-BTZ were computed by time-dependent density functional theory [80, 81]. S1 of DAC-BTZ mainly consists of HOMO-LUMO excitation and has charge-transfer (CT) character. The HOMO and LUMO of DAC-BTZ are predominantly distributed on DAC-II and BTZ, respectively (Fig. 2c,d), and are spatially well separated. Consequently, ΔEST is sufficiently small to facilitate TADF. The outer N atoms of DAC-II both decrease ΔEST and increase μ10, making fast kRISC and kr compatible. By shifting the HOMO in the opposite direction to BTZ as indicated by green arrows in Fig. 2c, DAC-II decreases the HOMO-LUMO spatial overlap particularly on the central phenyl ring, leading to small ΔEST. At the same time, the overlap density between the HOMO and LUMO extends over the entire molecule, leading to a large μ10. The effects of DAC-II on the photophysical properties of TADF emitters have been discussed extensively elsewhere [41].

Synthesis and characterization of DAC-BTZ

9-(4-(Benzothiazol-2-yl)phenyl)-9H-carbazole

To a solution of 3.00 g of 2-(4-bromophenyl)-benzothiazole (10.3 mmol), 2.07 g of 9H-carbozole (12.4 mmol), 1.43 g of sodium tert-butoxide (14.9 mmol) and 0.19 g of tris(dibenzylideneacetone)dipalladium(0) (0.21 mmol) in 50 mL of toluene was added, with stirring, a solution of 2.0 M tri-tert-butylphosphine in 0.50 mL of toluene. The stirred mixture was heated under reflux for 7 h, and then 10 mL of water and 30 mL of chloroform were added to the cooled mixture. The resulting mixture was stirred for 30 min and then filtered through a pad of Celite. The filtrate was extracted with chloroform. The combined organic layers were washed with brine, dried over Mg2SO4, and concentrated in vacuo. Column chromatography of the residue solid (eluent: toluene/ethyl acetate = 5/1) afforded 3.45 g of the target compound (yield: 88%). 1H NMR (CDCl3, 500 MHz) δ = 7.32 (t, 2H, J = 7.50 Hz), 7.44 (t, 1H, J = 7.50 Hz), 7.45 (t, 2H, J = 7.75 Hz), 7.51 (d, 2H, J = 8.50 Hz), 7.54 (t, 1H, J = 7.00 Hz), 7.73 (d, 2H, J = 8.50 Hz), 7.95 (d, 1H, J = 8.00 Hz), 8.12 (d, 1H, J = 8.00 Hz), 8.15 (d, 2H, J = 8.00 Hz), 8.34 (d, 2H, J = 8.50 Hz). Matrix-assisted laser desorption/ionization mass spectrometry (MALDI-MS) m/z Calcd for C25H16N2S: 376; found: 376.

Scheme 1: 
            Synthesis of DAC-BTZ.
Scheme 1:

Synthesis of DAC-BTZ.

9-(4-(Benzothiazol-2-yl)phenyl)-3,6-dibromo-9H-carbazole

A mixture of 2.40 g of 9-(4-(benzothiazol-2-yl)phenyl)-9H-carbazole (6.38 mmol), 20 mL of tetrahydrofuran and 20 mL of ethyl acetate was stirred at 0–5 °C for 20 min. Then, 2.72 g of N-bromosuccinimide (15.3 mmol) was added in small portions, and the mixture was stirred at ambient temperature for 20 h. The resulting mixture was washed with an aqueous solution of sodium thiosulfate and extracted with ethyl acetate. The combined organic layers were washed with brine, dried over Mg2SO4, and concentrated in vacuo. The solid residue was washed with water and collected by filtration. The resulting solid was added to a mixture of acetone and methanol, and processed by sonication. The insoluble solid was collected by filtration to afford 3.30 g of the target compound (yield: 96%). 1H NMR (CDCl3, 500 MHz) δ = 7.34 (d, 2H, J = 8.50 Hz), 7.45 (t, 1H, J = 7.50 Hz), 7.54 (d, 2H, J = 8.50 Hz), 7.55 (t, 1H, J = 7.50 Hz), 7.66 (d, 2H, J = 6.50 Hz), 7.95 (d, 1H, J = 8.00 Hz), 8.12 (d, 1H, J = 8.50 Hz), 8.21 (s, 2H), 8.33 (d, 2H, J = 8.50 Hz). MALDI-MS m/z Calcd for C25H14Br2N2S: 531; found: 531.

9-(4-(Benzo[d]thiazol-2-yl)phenyl)-N3,N3,N6,N6-tetraphenyl-9H-carbazole-3, 6-diamine (DAC-BTZ)

To a solution of 2.00 g of 9-(4-(benzothiazol-2-yl)phenyl)-3,6-dibromo-9H-carbazole (3.74 mmol), 1.58 g of diphenylamine (9.36 mmol), 1.08 g of sodium tert-butoxide (11.2 mmol) and 68.4 mg of tris(dibenzylideneacetone)dipalladium (0.07 mmol) in 20 mL of toluene was added, with stirring, a solution of 2.0 M tri-tert-butylphosphine in 0.30 mL of toluene. The stirred mixture was heated under reflux for 20 h. Then, 100 mL of chloroform was added to the cooled mixture, which was subsequently stirred for 30 min and filtered through a pad of Celite and silica gel. The filtrate was extracted with chloroform. The combined organic layers were washed with brine, dried over Mg2SO4, and concentrated in vacuo. Column chromatography of the solid residue (eluent: toluene) afforded 2.48 g of the target compound (yield: 93%). This material was further purified by sublimation under reduced pressure before OLED fabrication. 1H NMR (DMSO, 500 MHz) δ = 6.94 (t, 4H, J = 7.25 Hz), 6.97 (d, 8H, J = 8.00 Hz), 7.24 (m, 10H), 7.51 (t, 1H, J = 7.75 Hz), 7.54 (d, 2H, J = 9.00 Hz), 7.59 (t, 1H, J = 7.75 Hz), 7.89 (d, 2H, J = 8.50 Hz), 8.02 (s, 2H), 8.11 (d, 1H, J = 8.00 Hz), 8.20 (d, 1H, J = 8.00 Hz), 8.38 (d, 2H, J = 8.50 Hz); 13C NMR (DMSO, 500 MHz) d = 111.3, 119.2, 121.6, 122.1, 122.4, 123.0, 124.0, 125.7, 126.3, 126.7, 127.1, 128.9, 129.2, 131.6, 134.6, 137.5, 139.3, 140.3, 147.9, 153.6, 166.3. MALDI-MS m/z Calcd for C49H34N4S: 710; found: 710. Anal. Calcd for C49H34N4S: C, 82.79; H, 4.82; N, 7.88; found: C, 82.69; H, 4.78; N, 7.87.

Experimental methods

NMR spectra were recorded on a Bruker Biospin Avance-III 500 NMR spectrometer at ambient temperature. All NMR spectra were referenced to residual solvent peaks. Matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF-MS) data were recorded on a Bruker Daltonics Autoflex III spectrometer in positive mode. UV-vis spectra were obtained using a UV-vis spectrophotometer (UV-2550, Shimadzu, Japan). PL spectra were recorded using a spectrofluorometer (Fluoromax-4, Horiba Jobin Yvon). PLQYs were measured using an absolute PLQY spectrometer (C11347 Quantaurus-QY, Hamamatsu, Japan). Transient PL decays of solution were obtained using a fluorescence lifetime spectrometer (C11367 Quantaurus-tau, Hamamatsu Photonics, Japan). Transient PL decays of doped films were measured using a streak camera (C4334, Hamamatsu Photonics, Japan). A N2 gas laser with a wavelength of 337 nm (Ken-X, Usho Optical Systems, Japan) was used as the excitation source. The HOMO energy level of DAC-BTZ was determined from photoelectron spectrometry (Riken Keiki, AC-3) measurements of a neat DAC-BTZ film fabricated by vacuum deposition. The LUMO energy level of DAC-BTZ was determined from its HOMO energy level and the UV-vis absorption spectrum of its neat film.

TADF behavior of DAC-BTZ in toluene solution

Figure 3a shows UV-vis absorption and PL spectra measured for a solution of DAC-BTZ (10–5 M) in toluene at room temperature. The peak maxima of the first absorption and emission bands are at 383 and 496 nm, respectively. The broad absorption and emission bands reflect the CT character of S1 of DAC-BTZ. In air, the PLQY of DAC-BTZ in toluene solution was 0.45. This value increased to 0.65 after bubbling N2 gas through the solution for 10 min. The increase in PLQY probably originates from TADF. In air, TADF is suppressed because triplet states of DAC-BTZ are quenched by oxygen dissolved in the solution, so the PLQY is low. Bubbling N2 gas through the solution removes oxygen from it, allowing triplet states to survive. As a result, TADF is observed from DAC-BTZ, causing the increase in PLQY. Figure 3b shows the transient PL decays for DAC-BTZ in toluene solution. After N2 gas bubbling, a long-lived component was observed in the transient PL decay curve, reflecting the generation of TADF. This oxygen-dependent quenching of TADF provides a quick and convenient method to screen the potential of TADF emitters. As stated below, the long-lived component caused by TADF is more clearly observed when DAC-BTZ is doped into a solid-state host layer than in solution.

Fig. 3: 
          (a) UV-vis absorption and PL spectra of DAC-BTZ in toluene solution (10–5 M). (b) Transient photoluminescence decays for DAC-BTZ in toluene solution measured in air and after bubbling N2 gas through the solution.
Fig. 3:

(a) UV-vis absorption and PL spectra of DAC-BTZ in toluene solution (10–5 M). (b) Transient photoluminescence decays for DAC-BTZ in toluene solution measured in air and after bubbling N2 gas through the solution.

TADF behavior of DAC-BTZ in a solid-state host layer

To evaluate DAC-BTZ as a potential emitter for OLEDs, we investigated its TADF behavior in a layer of bis(2-(diphenylphosphino)phenyl)ether oxide (DPEPO) [82]. DPEPO has high T1 energy and has been widely used as a host material for blue TADF emitters [16, 17, 20, 21, 23, 24, 29–31, 43]. A 6 wt% DAC-BTZ-doped DPEPO thin film was fabricated by vacuum deposition. ΦPL of the doped film was 0.82. Figure 4a shows transient PL decays of the film measured in the time range from 0 to 10 ms at temperatures of 100, 200, and 300 K. The decay curves consist of an intense peak and long tail. The intense peak corresponds to prompt fluorescence, while the long tail corresponds to delayed fluorescence. The delayed fluorescence originates from TADF because its intensity increases with increasing temperature. The decay curve at 300 K was fitted using a triexponential function, indicating superimposition of 1*CT → S0 and 3*CT → S0 emissions resulting from efficient spin-orbit coupling induced by the heavy sulfur atom. The decay times of the fitted function were 52 μs (prompt), 660 μs (delayed), and 4.6 ms (delayed). The respective amplitudes connected with the decay times were 1.900, 0.012, and 0.003. The decay time of 52 μs is long compared with conventional fluorescence lifetime. This is because when transient PL decay is measured in 0–10 milliseconds time range, time resolution is insufficient to measure decay time for prompt fluorescence. To achieve sufficiently high time resolution, PL decays should be measured in 0–100 ns as described below. From the fitted function, Φp and Φd were calculated to be 0.67 and 0.15, respectively. Therefore, the T1-to-light conversion efficiency of DAC-BTZ is 45%. Figure 4b shows transient PL decays measured for the same film in the time range of 0–100 ns at temperatures of 100, 200, and 300 K. The decay curve at 300 K in Fig. 4b was fitted using a triexponential function with decay times of 5.5 ns, 660 μs, and 4.6 ms. Thus, the lifetime of the prompt fluorescence was determined to be 5.5 ns.

Fig. 4: 
          Transient photoluminescence decays for a 6 wt% DAC-BTZ-doped DPEPO thin film measured at temperatures of 100, 200, and 300 K for time ranges of (a) 0–10 ms and (b) 0–100 ns. (c) Photoluminescence spectrum of the doped film measured at 300 K. (d) Prompt fluorescence and phosphorescence spectra of the doped film measured at 9 K.
Fig. 4:

Transient photoluminescence decays for a 6 wt% DAC-BTZ-doped DPEPO thin film measured at temperatures of 100, 200, and 300 K for time ranges of (a) 0–10 ms and (b) 0–100 ns. (c) Photoluminescence spectrum of the doped film measured at 300 K. (d) Prompt fluorescence and phosphorescence spectra of the doped film measured at 9 K.

Figure 4c shows a PL spectrum of the doped film. The broad emission band is attributed to fluorescence from DAC-BTZ. The peak position of the emission band is consistent with that measured for the toluene solution (Fig. 3a). ΔEST of TADF emitters have been elucidated by various methods including the energy difference between the onsets of fluorescence and phosphorescence spectra [17, 18, 20, 23–25, 29–31, 34, 37, 38, 41, 43, 44], the difference between peak wavelengths of fluorescence and phosphorescence spectra [11, 40, 42, 58], the temperature dependence of kRISC [7, 12, 21, 36], and Berberan-Santos plots [14, 15, 32]. The method based on shift between fluorescence and phosphorescence spectra is valid when the energies of the Franck-Condon states reached in 1*CT → S0 and 3*CT → S0 emissions are same. Otherwise, it underestimates ΔEST. Figure 4d shows prompt fluorescence and phosphorescence spectra of the doped film measured at 9 K. ΔEST of DAC-BTZ was estimated to be 0.18–0.22 eV from the energy difference between the onsets of these spectra.

Figure 5 illustrates the PL process of the doped film and proposed EL process of a DAC-BTZ-based OLED based on the above transient PL decay measurements. For the PL process, of S1 excitons generated by photoexcitation, 67% decay radiatively, and the remaining 33% are converted to T1 excitons (Fig. 5a,b). Because 45 % of the T1 excitons are converted into light as TADF (Fig. 5c), 82% of the total excitons are extracted as light. For the EL process, out of the 25% of S1 excitons generated by charge recombination, 67% (17% of the total excitons) decay radiatively, and the remaining 33% (8% of the total excitons) are converted to T1 excitons. Out of the 83% (75% + 8% of the total excitons) of T1 excitons, 45% are converted into light as TADF. As a result, 54% of the total excitons are extracted as light. From eq. 4, the maximum IQE (γ = 1) for the doped film is calculated to be 54%, and ηEQE of a DAC-BTZ-based OLED should be 11–16%. The calculated ηEQE is higher than those obtained using conventional fluorescent OLEDs (5–7.5%), suggesting that DAC-BTZ is promising as an EL material.

Fig. 5: 
          (a) Photoluminescence process for the 6 wt% DAC-BTZ:DPEPO thin film. (b) Electroluminescence process for a DAC-BTZ-based OLED.
Fig. 5:

(a) Photoluminescence process for the 6 wt% DAC-BTZ:DPEPO thin film. (b) Electroluminescence process for a DAC-BTZ-based OLED.

Electroluminescence from a DAC-BTZ-based OLED

An OLED containing DAC-BTZ as a sky-blue emitter was fabricated by vacuum deposition. The structure of the OLED was indium tin oxide (ITO)/4,4′-bis[N-(1-naphthyl)-N-phenyl]biphenyl diamine (α-NPD) (30 nm)/ 9,9′-biphenyl-3,3′-diyl-bis-9H-carbazole (m-CBP) (10 nm)/6 wt% DAC-BTZ:DPEPO (15 nm)/DPEPO (10 nm)/ 1,3,5-tris(N-phenylbenzimidizol-2-yl)benzene (TPBi) (35 nm)/LiF/Al (Fig. 6a). The HOMO and LUMO energy levels of DAC-BTZ were determined to be 5.4 and 2.7 eV, respectively, which are inside the HOMO-LUMO energy gap of DPEPO. Figure 6b shows the EQE-current density characteristics and EL spectrum of the OLED. The OLED exhibits sky-blue emission with a peak wavelength of 493 nm. The maximum EQE of the OLED was 10.3% at a current density of 3 × 10–3 mA/cm2. Although the maximum EQE is high compared with those of conventional fluorescent OLEDs (5–7.5%), it is lower than that predicted theoretically from Φp and Φd (11–16%). This suggests that there is room to further increase the maximum EQE, e.g. by improving carrier balance and controlling emitter orientation [60–62]. Assuming that γηout = 0.19, eq. 3 gives an experimental maximum EQE of 10.3%. By using γηout = 0.19 in eqs. 5 and 6, ηp and ηd are calculated to be 3.2% and 7.1%, respectively. Therefore, without TADF, the EQE of the OLED would be 3.2%. ηd is more than twice as large as ηp, suggesting that TADF makes a considerable contribution to the high EQE of this device.

Fig. 6: 
          (a) Energy levels (in eV) of the components in the DAC-BTZ-based OLED. (b) External quantum efficiency-current density characteristics of the OLED. The inset shows the electroluminescence spectrum of the device at a current density of 100 mA/cm2.
Fig. 6:

(a) Energy levels (in eV) of the components in the DAC-BTZ-based OLED. (b) External quantum efficiency-current density characteristics of the OLED. The inset shows the electroluminescence spectrum of the device at a current density of 100 mA/cm2.

Summary

We briefly reviewed the recent progress in TADF research and reported the novel sky-blue TADF emitter, DAC-BTZ. DAC-BTZ is a purely organic donor–acceptor system exhibiting high luminescence efficiency. When doped into the host DPEPO, DAC-BTZ showed a high PLQY of 0.82. Transient PL decay measurements confirmed that DAC-BTZ emitted TADF in the DPEPO host layer. An OLED containing DAC-BTZ as a sky-blue emitter exhibited a high EQE of 10.3%, outperforming conventional fluorescent OLEDs. This high EQE results from the efficient generation of TADF in DAC-BTZ.


Article note

A collection of invited papers based on presentations at the XXVth IUPAC Symposium on Photochemistry, Bordeaux, France, July 13–18, 2014.



Corresponding author: Chihaya Adachi, Center for Organic Photonics and Electronics Research (OPERA), Kyushu University, 744 Motooka, Nishi, Fukuoka 819-0395, Japan; Japan Science and Technology Agency (JST), ERATO, Adachi Molecular Exciton Engineering Project, 744 Motooka, Nishi, Fukuoka 819-0395, Japan; and International Institute for Carbon Neutral Energy Research (WPI-I2CNER), Kyushu University, 744 Motooka, Nishi, Fukuoka 819-0395, Japan, e-mail:

Acknowledgments

This research was funded by the Japan Society for the Promotion of Science (JSPS) through the “Funding Program for World-Leading Innovative R&D on Science and Technology (FIRST Program)”. Computations were partly carried out using computer facilities at the Research Institute for Information Technology, Kyushu University, and the Academic Center for Computing and Media Studies (ACCMS), Kyoto University. Computation time was provided by the Super Computer System, Institute for Chemical Research, Kyoto University.

References

[1] G. N. Lewis, D. Lipkin, T. T. Magel. J. Am. Chem. Soc.63, 3005 (1941).Search in Google Scholar

[2] C. A. Parker, C. G. Hatchard. J. Phys. Chem.66, 2506 (1962).Search in Google Scholar

[3] P. F. Jones, A. R. Calloway. Chem. Phys. Lett.10, 438 (1971).Search in Google Scholar

[4] A. Maciejewski, M. Szymanski, R. P. Steer. J. Phys. Chem.90, 6314 (1986).Search in Google Scholar

[5] M. N. Berberan-Santos, J. M. M. Garcia. J. Am. Chem. Soc.118, 9391 (1996).Search in Google Scholar

[6] F. A. Salazar, A. Fedorov, M. N. Berberan-Santos. Chem. Phys. Lett.271, 361 (1997).Search in Google Scholar

[7] A. Endo, M. Ogasawara, A. Takahashi, D. Yokoyama, Y. Kato, C. Adachi. Adv. Mater.21, 4802 (2009).Search in Google Scholar

[8] S. R. Forrest. Nature428, 911 (2004).10.1038/nature02498Search in Google Scholar PubMed

[9] S. Reineke, M. Thomschke, B. Lüssem, K. Leo. Rev. Mod. Phys.85, 1245 (2013).Search in Google Scholar

[10] H. Sasabe, J. Kido. J. Mater. Chem. C1, 1699 (2013).10.1039/c2tc00584kSearch in Google Scholar

[11] A. Endo, K. Sato, K. Yoshimura, T. Kai, A. Kawada, H. Miyazaki, C. Adachi. Appl. Phys. Lett.98, 083302 (2011).Search in Google Scholar

[12] H. Uoyama, K. Goushi, K. Shizu, H. Nomura, C. Adachi. Nature492, 234 (2012).10.1038/nature11687Search in Google Scholar PubMed

[13] C. Adachi. Jpn. J. Appl. Phys.53, 060101 (2014).10.7567/JJAP.53.060101Search in Google Scholar

[14] T. Nakagawa, S.-Y. Ku, K.-T. Wong, C. Adachi. Chem. Commun.48, 9580 (2012).Search in Google Scholar

[15] G. Méhes, H. Nomura, Q. Zhang, T. Nakagawa, C. Adachi. Angew. Chem. Int. Ed.51, 11311 (2012).Search in Google Scholar

[16] S. Y. Lee, T. Yasuda, H. Nomura, C. Adachi. Appl. Phys. Lett.101, 093306 (2012).Search in Google Scholar

[17] Q. Zhang, J. Li, K. Shizu, S. Huang, S. Hirata, H. Miyazaki, C. Adachi. J. Am. Chem. Soc.134, 14706 (2012).Search in Google Scholar

[18] H. Tanaka, K. Shizu, H. Miyazaki, C. Adachi. Chem. Commun.48, 11392 (2012).Search in Google Scholar

[19] H. Tanaka, K. Shizu, H. Nakanotani, C. Adachi. Chem. Mater.25, 3766 (2013).Search in Google Scholar

[20] J. Lee, K. Shizu, H. Tanaka, H. Nomura, T. Yasuda, C. Adachi. J. Mater. Chem. C1, 4599 (2013).10.1039/c3tc30699bSearch in Google Scholar

[21] K. Sato, K. Shizu, K. Yoshimura, A. Kawada, H. Miyazaki, C. Adachi. Phys. Rev. Lett.110, 247401 (2013).Search in Google Scholar

[22] G. Valchanov, A. Ivanova, A. Tadjer, D. Chercka, M. Baumgarten. Org. Electron.14, 2727 (2013).Search in Google Scholar

[23] T. Serevicius, T. Nakagawa, M.-C. Kuo, S.-H. Cheng, K.-T. Wong, C.-H. Chang, R. C. Kwong, S. Xia, C. Adachi. Phys. Chem. Chem. Phys.15, 15850 (2013).Search in Google Scholar

[24] K. Nasu, T. Nakagawa, H. Nomura, C.-J. Lin, C.-H. Cheng, M.-R. Tseng, T. Yasuda, C. Adachi. Chem. Commun.49, 10385 (2013).Search in Google Scholar

[25] J. Li, T. Nakagawa, J. MacDonald, Q. Zhang, H. Nomura, H. Miyazaki, C. Adachi. Adv. Mater.25, 3319 (2013).Search in Google Scholar

[26] J. Li, Q. Zhang, H. Nomura, H. Miyazaki, C. Adachi. Appl. Phys. Lett.105, 013301 (2014).Search in Google Scholar

[27] J. Li, H. Nomura, H. Miyazaki, C. Adachi. Chem. Commun.50, 6174 (2014).Search in Google Scholar

[28] Q. Zhang, H. Kuwabara, W. J. Potscavage, S. Huang, Y. Hatae, T. Shibata, C. Adachi. J. Am. Chem. Soc.136, 18070 (2014).Search in Google Scholar

[29] S. Wu, M. Aonuma, Q. Zhang, S. Huang, T. Nakagawa, K. Kuwabara, C. Adachi. J. Mater. Chem. C2, 421 (2014).10.1039/C3TC31936ASearch in Google Scholar

[30] Q. Zhang, B. Li, S. Huang, H. Nomura, H. Tanaka, C. Adachi. Nat. Photon.8, 326 (2014).Search in Google Scholar

[31] S. Y. Lee, T. Yasuda, Y. S. Yang, Q. Zhang, C. Adachi. Angew. Chem. Int. Ed.126, 6520 (2014).Search in Google Scholar

[32] H. Wang, L. Xie, Q. Peng, L. Meng, Y. Wang, Y. Yi, P. Wang. Adv. Mater.26, 5198 (2014).Search in Google Scholar

[33] H. Ohkuma, T. Nakagawa, K. Shizu, T. Yasuda, C. Adachi. Chem. Lett.43, 1017 (2014).Search in Google Scholar

[34] T. Takahashi, K. Shizu, T. Yasuda, K. Togashi, C. Adachi. Sci. Tech. Adv. Mater.15, 034202 (2014).Search in Google Scholar

[35] H. Tanaka, K. Shizu, H. Nakanotani, C. Adachi. J. Phys. Chem. C118, 15985 (2014).10.1021/jp501017fSearch in Google Scholar

[36] Y. Sagara, K. Shizu, H. Tanaka, H. Miyazaki, K. Goushi, H. Kaji, C. Adachi. Chem. Lett.44, 360 (2015).Search in Google Scholar

[37] J. Lee, K. Shizu, H. Tanaka, H. Nakanotani, T. Yasuda, H. Kaji, C. Adachi. J. Mater. Chem. C3, 2175 (2015).10.1039/C4TC02530JSearch in Google Scholar

[38] H. Tanaka, K. Shizu, J. Lee, C. Adachi. J. Phys. Chem. C119, 2948 (2015).10.1021/jp510751nSearch in Google Scholar

[39] M. Taneda, K. Shizu, H. Tanaka, C. Adachi. Chem. Commun.51, 5028 (2015).Search in Google Scholar

[40] K. Shizu, M. Uejima, H. Nomura, T. Sato, K. Tanaka, H. Kaji, C. Adachi. Phys. Rev. Applied3, 014001 (2015).10.1103/PhysRevApplied.3.014001Search in Google Scholar

[41] K. Shizu, H. Tanaka, M. Uejima, T. Sato, K. Tanaka, H. Kaji, C. Adachi. J. Phys. Chem. C119, 1291 (2015).10.1021/jp511061tSearch in Google Scholar

[42] K. Shizu, Y. Sakai, H. Tanaka, S. Hirata, C. Adachi, H. Kaji. ITE Trans. on MTA3, 108 (2015).10.3169/mta.3.108Search in Google Scholar

[43] S. Hirata, Y. Sakai, K. Masui, H. Tanaka, S. Y. Lee, H. Nomura, N. Nakamura, M. Yasumatsu, H. Nakanotani, Q. Zhang, K. Shizu, H. Miyazaki, C. Adachi. Nat. Mater.14, 330 (2015).Search in Google Scholar

[44] Y. Sakai, Y. Sagara, H. Nomura, N. Nakamura, Y. Suzuki, H. Miyazaki, C. Adachi. Chem. Commun.51, 3181 (2015).Search in Google Scholar

[45] J. C. Deaton, S. C. Switalski, D. Y. Kondakov, R. H. Young, T. D. Pawlik, D. J. Giesen, S. B. Harkins, A. J. M. Miller, S. F. Mickenberg, J. C. Peters. J. Am. Chem. Soc.132, 9499 (2010).Search in Google Scholar

[46] R. Czerwieniec, J. Yu, H. Yersin. Inorg. Chem.50, 8293 (2011).Search in Google Scholar

[47] R. Czerwieniec, K. Kowalski, H. Yersin. Dalton Trans.42, 9826 (2013).Search in Google Scholar

[48] X.-L. Chen, R. Yu, Q.-K. Zhang, L.-J. Zhou, X.-Y. Wu, Q. Zhang, C.-Z. Lu. Chem. Mater.25, 3910 (2013).Search in Google Scholar

[49] M. Osawa, I. Kawata, R. Ishii, S. Igawa, M. Hashimoto, M. Hoshino. J. Mater. Chem. C1, 4375 (2013).10.1039/c3tc30524dSearch in Google Scholar

[50] S. Igawa, M. Hashimoto, I. Kawata, M. Yashima, M. Hoshino, M. Osawa. J. Mater. Chem. C1, 542 (2013).10.1039/C2TC00263ASearch in Google Scholar

[51] T. Hofbeck, U. Monkowius, H. Yersin. J. Am. Chem. Soc.137, 399 (2014).Search in Google Scholar

[52] M. Osawa. Chem. Commun.50, 1801 (2014).10.1039/c3cc47871hSearch in Google Scholar PubMed

[53] C. L. Linfoot, M. J. Leitl, P. Richardson, A. F. Rausch, O. Chepelin, F. J. White, H. Yersin, N. Robertson. Inorg. Chem.53, 10854 (2014).Search in Google Scholar

[54] M. Osawa, M. Hoshino, M. Hashimoto, I. Kawata, S. Igawa, M. Yashima. Dalton Trans.44, 8369 (2015).Search in Google Scholar

[55] T. Gneuß, M. J. Leitl, L. H. Finger, N. Rau, H. Yersin, J. Sundermeyer. Dalton Trans.44, 8506 (2015).Search in Google Scholar

[56] H. Nakanotani, K. Masui, J. Nishide, T. Shibata, C. Adachi. Sci. Rep.3, 2127 (2013).Search in Google Scholar

[57] J.-i. Nishide, H. Nakanotani, Y. Hiraga, C. Adachi. Appl. Phys. Lett.104, 233304 (2014).Search in Google Scholar

[58] H. Nakanotani, T. Higuchi, T. Furukawa, K. Masui, K. Morimoto, M. Numata, H. Tanaka, Y. Sagara, T. Yasuda, C. Adachi. Nat. Commun.5 (2014).10.1038/ncomms5016Search in Google Scholar PubMed

[59] D. Zhang, L. Duan, C. Li, Y. Li, H. Li, D. Zhang, Y. Qiu. Adv. Mater.26, 5050 (2014).Search in Google Scholar

[60] T. Komino, H. Tanaka, C. Adachi. Chem. Mater.26, 3665 (2014).Search in Google Scholar

[61] J. W. Sun, J.-H. Lee, C.-K. Moon, K.-H. Kim, H. Shin, J.-J. Kim. Adv. Mater.26, 5684 (2014).Search in Google Scholar

[62] C. Mayr, S. Y. Lee, T. D. Schmidt, T. Yasuda, C. Adachi, W. Brütting. Adv. Func. Mater.24, 5232 (2014).Search in Google Scholar

[63] Y. J. Cho, K. S. Yook, J. Y. Lee. Adv. Mater.26, 6642 (2014).Search in Google Scholar

[64] Y. Suzuki, Q. Zhang, C. Adachi. J. Mater. Chem. C3, 1700 (2015).10.1039/C4TC02211DSearch in Google Scholar

[65] K. Kawano, K. Nagayoshi, T. Yamaki, C. Adachi. Org. Electron.15, 1695 (2014).Search in Google Scholar

[66] R. Ishimatsu, S. Matsunami, T. Kasahara, J. Mizuno, T. Edura, C. Adachi, K. Nakano, T. Imato. Angew. Chem. Int. Ed.53, 6993 (2014).Search in Google Scholar

[67] X. Xiong, F. Song, J. Wang, Y. Zhang, Y. Xue, L. Sun, N. Jiang, P. Gao, L. Tian, X. Peng. J. Am. Chem. Soc.136, 9590 (2014).Search in Google Scholar

[68] B. S. Kim, J. Y. Lee. Adv. Func. Mater.24, 3970 (2014).Search in Google Scholar

[69] Y. J. Cho, K. S. Yook, J. Y. Lee. Adv. Mater.26, 4050 (2014).Search in Google Scholar

[70] M. P. Gaj, C. Fuentes-Hernandez, Y. Zhang, S. R. Marder, B. Kippelen. Org. Electron.16, 109 (2015).Search in Google Scholar

[71] J. Herbich, A. Kapturkiewicz, J. Nowacki. Chem. Phys. Lett.262, 633 (1996).Search in Google Scholar

[72] A. Kapturkiewicz, J. Herbich, J. Nowacki. Chem. Phys. Lett.275, 355 (1997).Search in Google Scholar

[73] P. Borowicz, J. Herbich, A. Kapturkiewicz, R. Anulewicz-Ostrowska, J. Nowacki, G. Grampp. Phys. Chem. Chem. Phys.2, 4275 (2000).Search in Google Scholar

[74] L. Yao, S. Zhang, R. Wang, W. Li, F. Shen, B. Yang, Y. Ma. Angew. Chem.126, 2151 (2014).Search in Google Scholar

[75] L. Yao, S. Zhang, R. Wang, W. Li, F. Shen, B. Yang, Y. Ma. Angew. Chem. Int. Ed.53, 2119 (2014).Search in Google Scholar

[76] T. Sato, M. Uejima, K. Tanaka, H. Kaji, C. Adachi. J. Mater. Chem. C3, 870 (2015).10.1039/C4TC02320JSearch in Google Scholar

[77] M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski, D. J. Fox, (GAUSSIAN 09, Revision C.01, Gaussian, Inc., Wallingford, CT, 2009).Search in Google Scholar

[78] Y. Zhao, D. Truhlar. Theor. Chem. Acc.120, 215 (2008).Search in Google Scholar

[79] J. T. H. Dunning. J. Chem. Phys.90, 1007 (1989).10.1063/1.456153Search in Google Scholar

[80] R. Bauernschmitt, R. Ahlrichs. Chem. Phys. Lett.256, 454 (1996).Search in Google Scholar

[81] M. E. Casida, C. Jamorski, K. C. Casida, D. R. Salahub. J. Chem. Phys.108, 4439 (1998).Search in Google Scholar

[82] H. Xu, L.-H. Wang, X.-H. Zhu, K. Yin, G.-Y. Zhong, X.-Y. Hou, W. Huang. J. Phys. Chem. B110, 3023 (2006).10.1021/jp055355pSearch in Google Scholar PubMed

Published Online: 2015-06-26
Published in Print: 2015-07-01

©2015 IUPAC & De Gruyter

Downloaded on 1.5.2024 from https://www.degruyter.com/document/doi/10.1515/pac-2015-0301/html
Scroll to top button