Next Article in Journal
Computational Modeling of DNA 3D Structures: From Dynamics and Mechanics to Folding
Next Article in Special Issue
Trapping an Ester Hydrate Intermediate in a π-Stacked Macrocycle with Multiple Hydrogen Bonds
Previous Article in Journal
A Low-Cost Colorimetric Assay for the Analytical Determination of Copper Ions with Consumer Electronic Imaging Devices in Natural Water Samples
Previous Article in Special Issue
Surfactant Additives Containing Hydrophobic Fluorocarbon Chains and Hydrophilic Sulfonate Anion for Highly Reversible Zn Anode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tröger’s Base-Derived Thermally Activated Delayed Fluorescence Dopant for Efficient Deep-Blue Organic Light-Emitting Diodes

1
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, China
2
Fujian Science & Technology Innovation Laboratory for Optoelectronic Information of China, Fuzhou 350108, China
3
Xiamen Key Laboratory of Rare Earth Photoelectric Functional Materials, Xiamen Institute of Rare Earth Materials, Haixi Institutes, Chinese Academy of Sciences, Xiamen 361021, China
4
University of Chinese Academy of Sciences, Beijing 100049, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(12), 4832; https://doi.org/10.3390/molecules28124832
Submission received: 29 April 2023 / Revised: 2 June 2023 / Accepted: 13 June 2023 / Published: 17 June 2023
(This article belongs to the Special Issue Multifunctional Metal Oxides: Synthesis and Applications)

Abstract

:
The development of efficient deep-blue emitters with thermally activated delayed fluorescence (TADF) properties is a highly significant but challenging task in the field of organic light-emitting diode (OLED) applications. Herein, we report the design and synthesis of two new 4,10-dimethyl-6H,12H-5,11-methanodibenzo[b,f][1,5]diazocine (TB)-derived TADF emitters, TB-BP-DMAC and TB-DMAC, which feature distinct benzophenone (BP)-derived acceptors but share the same dimethylacridin (DMAC) donors. Our comparative study reveals that the amide acceptor in TB-DMAC exhibits a significantly weaker electron-withdrawing ability in comparison to that of the typical benzophenone acceptor employed in TB-BP-DMAC. This disparity not only causes a noticeable blue shift in the emission from green to deep blue but also enhances the emission efficiency and the reverse intersystem crossing (RISC) process. As a result, TB-DMAC emits efficient deep-blue delay fluorescence with a photoluminescence quantum yield (PLQY) of 50.4% and a short lifetime of 2.28 μs in doped film. The doped and non-doped OLEDs based on TB-DMAC display efficient deep-blue electroluminescence with spectral peaks at 449 and 453 nm and maximum external quantum efficiencies (EQEs) of 6.1% and 5.7%, respectively. These findings indicate that substituted amide acceptors are a viable option for the design of high-performance deep-blue TADF materials.

1. Introduction

Over the past few decades, organic light-emitting diodes (OLEDs) have attracted extensive academic and industrial attention in solid-state lighting and flexible ultra-thin displays due to their intrinsic merits of being ultra thin and lightweight as well as having high brightness and flexibility with infinite design possibilities [1,2,3]. Electroluminescence (EL) materials which determine the main device performance, such as emission color, EL efficiency, and operating lifetime, are the core part of OLEDs [4,5,6]. Thermally activated delayed fluorescent (TADF) materials, which can harvest triplet (T1) and singlet (S1) excitons for light emission via efficient reverse intersystem crossing (RISC), have emerged as promising fluorescent dyes for the construction of efficient OLEDs [7]. The thermally activatable RISC process requires a small amount of singlet–triplet energy splitting (∆EST), which can be achieved in donor–acceptor (D–A) molecules in which the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) distributions are well separated [8,9,10,11,12,13].
The development of deep-blue emitters is crucial in realizing full-color OLED displays with a high performance [14,15,16,17,18,19,20,21,22]. When designing deep-blue TADF emitters with D–A features, a weak CT character and a confined conjugation length are necessary to restrict the emission color to the deep-blue region [23]. However, the weakened CT character generally leads to a relatively large ΔEST and a small rate constant for reverse intersystem crossing (kRISC), which would negatively affect device efficiency and stability in OLEDs [24]. Hence, it is challenging, and strict design principles must be followed to balance the trade-offs between ΔEST, the RISC rate, the photoluminescence quantum yield (PLQY), and color purity of deep-blue TADF emitters [25,26,27].
Benzophenone-derived units are well known for their ability to undergo efficient intersystem crossing (ISC) due to their strong spin–orbit coupling [28]. This property makes them a highly attractive option as acceptor blocks for constructing TADF emitters. Although many benzophenone-based TADF materials have been reported to emit green to red light, achieving blue TADF materials has been challenging due to their strong electron-withdrawing ability [29,30,31,32,33,34,35,36]. Recently, Tang et al. modified the para-position of the carbonyl group of benzophenone with N to obtain sky blue TADF materials with satisfactory performances [37,38,39]. Duan and Adachi et al. proposed that minimizing the distance between heteroatoms in donors and acceptors can significantly enhance the contribution of heteroatoms to the natural orbitals of singlet and triplet states. This, in turn, can promote the proportion of the n-π* transition and increase the SOC between them, resulting in an improved RISC rate [40].
Tröger’s base (TB) is a C2-symmetric, unique, V-shaped twisted scaffold [41] that has been utilized to create amorphous host materials with a high thermal stability for OLEDs [42]. In this work, we designed and synthesized two novel 4,10-dimethyl-6H,12H-5,11-methanodibenzo[b,f][1,5]diazocine (TB)-derived TADF molecules with symmetrical D–A features, namely TB-BP-DMAC and TB-DMAC, bearing different benzophenone-derived acceptors and the same dimethylacridin (DMAC) donors. Compared to the typical benzophenone acceptor used in TB-BP-DMAC, the amide acceptor in TB-DMAC has a much weaker electron-withdrawing ability, resulting in a noticeable blue shift in the emission from green (with a spectrum peak at 517 nm for TB-BP-DMAC) to deep blue (with a spectrum peak at 467 nm for TB-DMAC). Additionally, TB-DMAC exhibits a higher photoluminescence quantum yield (PLQY) of 50.4% and a faster RISC rate of 6.13 × 105 s−1 compared to those of TB-BP-DMAC (PLQY = 31.5%, kRISC = 5.83 × 105 s−1). In a doped OLED employing TB-DMAC as the emitter, the deep-blue emission had a maximum external quantum efficiency (EQE) of 6.1%, which is higher than the 3.9% EQE of the TB-BP-DMAC-based green-emitting device.

2. Results and Discussion

2.1. Synthesis and Characterization

Synthetic routes and detailed synthetic procedures of the target compounds are described in Scheme 1 and the Section 3, respectively. The intermediate TB-BP-Br was synthesized in three steps, including condensation, Miyaura borylation reaction, and Suzuki coupling. Meanwhile, the intermediate TB-PhBr was obtained via two steps, namely condensation and N-acylation. The Pd-catalyzed Buchwald–Hartwig cross-coupling of the compounds TB-PhBr and TB-BP-Br with 9,9-dimethyl-9,10-dihydroacridine (DMAC) resulted in good yields of TB-BP-DMAC and TB-DMAC, respectively. All the newly synthesized materials were purified by column chromatography. These compounds are quite air stable in both the solid state and in solution. The chemical structures of the newly synthesized compounds were characterized using 1H NMR spectra, as shown in Supporting Information. The target compounds TB-BP-DMAC and TB-DMAC were further purified by vacuum sublimation before characterization and device fabrication were performed. The thermal stability of these compounds was examined by a thermal gravimetric analysis (TGA) under a nitrogen atmosphere. As depicted in Figure S7, both TB-BP-DMAC and TB-DMAC exhibit a high thermal stability, with decomposition temperatures (Td, corresponding to a 5% weight loss) of 270 and 299 °C, respectively. These values are sufficiently high to enable film and device fabrication through high-vacuum thermal deposition.

2.2. Theoretical Investigation

The geometric structures and electronic properties of these molecules have been investigated by DFT and time-dependent DFT (TD-DFT) calculations at the BMK/6-31G** level. The optimized ground-state molecular geometries revealed symmetrical twisted structures and nearly perpendicular donor–acceptor (D–A) conformations of these two molecules, with dihedral angles of 89.6–89.8° between the DMAC units and the bridging phenyl rings, respectively (Figure 1a). Such large D–A torsion angles typically promote effective spatial separation between the frontier molecular orbitals, thereby facilitating a small ΔEST, which is crucial for the reverse intersystem crossing (RISC) process. The highest occupied molecular orbitals (HOMOs) of these compounds are primarily confined at electron-donating DMAC moieties, while the LUMOs are mainly distributed over the electron-deficient benzophenone moiety, implying a good separation of HOMO and LUMO distributions (Figure 1b). The TD-DFT calculations revealed that the S1 states of TB-BP-DMAC and TB-DMAC have predominant CT characteristics, while their T1 states are both locally excited states (3LE). Compared with that of TB-DMAC, the molecular conjugation length of TB-BP-DMAC is significantly extended by adding a bridging phenyl ring and adjusting the connection position, resulting in a much deeper LUMO level than that of TB-DMAC (−1.26 versus −0.53 eV). Based on the TD-DFT calculations, the SOC constants for S1 and T1 states were found to be 0.45 and 0.86 cm−1 for TB-BP-DMAC and TB-DMAC, respectively. These significant SOC interactions may be attributed to the different excitation natures of these lowest excited states, which could play a crucial role in the RISC processes.

2.3. Electrochemical Properties

The electrochemical properties of these compounds were characterized by cyclic voltammetry (CV) (Figure S8). From the onsets of the oxidation curves, the HOMO energy levels of TB-BP-DMAC and TB-DMAC were calculated to be −5.26 and −5.28 eV, respectively. The close HOMO levels of two compounds can be ascribed to the use of identical donors. The LUMO energy levels were then determined to be −2.40 and −2.08 eV based on the HOMO levels and the optical band gaps (2.86 and 3.19 eV, vide infra) which were evaluated from the onsets of the absorption spectra. The LUMO energy level of TB-BP-DMAC was much deeper than that of TB-DMAC due to the significantly extended π-conjugation, which is consistent with the results of the quantum chemical calculations.

2.4. Photophysical Properties

The photophysical properties of these emitters were investigated in toluene (c = 10−5 M). Figure 2a displays the absorption spectra of TB-BP-DMAC and TB-DMAC in toluene. The intense higher-energy absorption bands observed at wavelengths below 380 nm and 335 nm, respectively, are primarily attributed to the spin-allowed π-π* transitions. Additionally, these molecules exhibit weak and broad absorption bands in a lower-energy region, which can be ascribed to the ICT transitions from the donor to the acceptor units. The absorption edges for these transitions are located at 434 nm for TB-BP-DMAC and 389 nm for TB-DMAC. Considering that they have the same donor units, the significant difference in the absorption edges between these compounds should originate from their obviously different acceptor strengths. TB-BP-DMAC and TB-DMAC exhibit bright green and deep-blue photoluminescence (PL) in toluene with broad and structureless spectra peaking at 511 and 454 nm, and corresponding Commission Internationale de L’Eclairage (CIE) color coordinates of (0.28, 0.50) and (0.16, 0.16), respectively. The broad and structureless PL spectra confirm the ICT characteristics of the emissive states. The significant blue shift observed in the PL spectrum of TB-DMAC, when compared to that of TB-BP-DMAC, is likely attributed to the significantly weaker acceptor strength of the former.
To be consistent with the situation in the optimized OLEDs, the photophysical properties of these two emitters were further investigated in 20 wt%-doped 2,8-bis(diphenylphosphinyl)dibenzo[b,d]furan (PPF) films, in which PPF served as a host. Figure S10 demonstrates that the PL characteristics of these emitters in doped films are similar to those observed in toluene solutions. The steady-state PL spectra show a minor redshift with peak wavelengths at 517 and 467 nm and corresponding CIE color coordinates of (0.30, 0.49) and (0.18, 0.21) for TB-BP-DMAC and TB-DMAC, respectively. To gain a deeper understanding of the emissive states, we investigated the transient photoluminescence (PL) decay behavior of these emitters across various temperatures (Figure S13). Figure 2b illustrates that at room temperature, each emitter exhibits a transient PL curve comprising two distinct components: a prompt component and a delayed component. The prompt (τPF) and delayed (τDF) component lifetimes were fitted to be 12.1 ns and 2.55 μs for TB-BP-DMAC and 11.6 ns and 2.28 μs for TB-DMAC, respectively (Figures S11 and S12). The proportion of the DF component increases progressively as the temperature is elevated from 77 K to 300 K, indicating the typical TADF properties of both emitters (Figure S13). Therefore, it can be concluded that the delayed components observed in these emitters originate from TADF involving T1 → S1 → S0 transitions. To experimentally determine the S1 and T1 energy levels, as well as the energy gap (ΔEST) between them, the time-resolved PL spectra of the doped film samples were acquired at 77 K (Figure 2c,d). The S1 energy, T1 energy, and ΔEST were calculated based on the onset wavelengths of the time-resolved PL spectra, yielding values of 2.81, 2.68, and 0.128 eV for TB-BP-DMAC and 3.12, 2.96, and 0.158 eV for TB-DMAC, respectively. The deep-blue emitter TB-DMAC demonstrates a relatively smaller ΔEST compared to the majority of previously reported deep-blue TADF emitters, while the green emitter TB-BP-DMAC exhibits a relatively larger ΔEST among green TADF emitters (Table S5). The small ΔEST values of these emitters allow an efficient up-conversion of excitons from the T1 state to the S1 state through RISC. Apparently, both the S1 and T1 states were effectively tuned by the substitution strategy. Directly linking the carbanyl groups with the nitrogen atoms on the Tröger’s base unit resulted in the formation of an amide acceptor, which exhibited a significantly weaker electron-deficient strength than that of the benzophenone acceptor and was found to be responsible for the observed changes in PL behaviors. The photoluminescence quantum yields (ΦPL) of TB-BP-DMAC and TB-DMAC were measured to be 31.5% and 50.4%, respectively. From ΦPL and the respective proportions of the prompt and delayed fluorescence components in the transient PL decay curves, the quantum yields of prompt fluorescence (ΦPF) and delayed fluorescence (ΦDF) in doped films were calculated to be 21.2% and 10.3% for TB-BP-DMAC and 36.1% and 14.3% for TB-DMAC, respectively (Table 1). The rate constants of the key photophysical processes were calculated based on the efficiency and lifetime data by using a reported method (see Supporting Information for details). For TB-BP-DMAC, the rate constants follow the order of the non-radiative rate constant ( k n r s ) > intersystem-crossing (ISC) rate constants (kISC) > the radiation rate constant ( k r s ). In contrast, for TB-DMAC, the order is k r s > k n r s > kISC (Table 1). The RISC rate constants (kRISC) of TB-BP-DMAC and TB-DMAC were calculated to be 5.83 × 105 and 6.13 × 105 s−1, respectively. These values suggest fast RISC processes in these molecules, which may be attributed to the small ΔEST values and the significant SOC interactions between the excited states involved in the RISC processes.

2.5. Electroluminescence Performance

To explore the electroluminescence (EL) properties of these TADF emitters, we fabricated multilayer doped OLEDs with a device configuration of ITO/HATCN (5 nm)/TAPC (30 nm)/TCTA (5 nm)/mCP (5 nm)/PPF: Emitter (30 nm)/PPF (10 nm)/Bphen (30 nm)/Liq (1 nm)/Al (100 nm), in which HATCN, TAPC, TCTA, mCP, PPF, Bphen, and Liq serve as the hole-injection layer, hole-transporting layer, electron-blocking layer, host, hole-blocking layer, electron-transporting layer, and electron-injecting layer, respectively. PPF was selected as a host material due to its high triplet energy (T1 = 3.1 eV), which is conducive to confine triplet excitons of the investigated emitters (T1 = 2.68–2.96 eV). The chemical structures of the functional organic materials used in the devices are shown in Figure S15.
At first, we investigated the dependence of EL performance on the doping concentration for TB-DMAC, using emitter doping concentrations of 10 wt%, 20 wt%, 30 wt%, and 100% (non-doped). The results are summarized in Figure 3 and Table 2. A slight spectral redshift in EL was observed as the doping concentration increased, with emission peaks ranging from 447 to 453 nm for TB-DMAC based devices when the doping concentration increased from 10% to 100% (Table 2). The TB-DMAC based device exhibited the highest EQEmax of 6.1% at a doping concentration of 20 wt%. Notably, while the doping concentration continued to increase to 30 wt% and 100% (non-doped), the EQEs did not decrease significantly. The non-doped device still maintained an EQEmax of 5.7%.
The 20 wt% doped OLEDs employing TB-BP-DMAC as an emitter were also fabricated and characterized for comparison (Figure 4 and Table 2). The 20 wt% doped OLEDs based on TB-BP-DMAC and TB-DAMC turned on at 5.3 and 2.9 V and displayed green and deep-blue emissions with spectral peaks at 508 nm and 449 nm and CIE coordinates of (0.28, 0.51) and (0.17, 0.15), respectively. The TB-DMAC-based device exhibited a higher EQEmax of 6.1%, while the TB-BP-DMAC-based device achieved an EQEmax of only 3.9%, probably due to the lower PLQY.

3. Experimental Section

3.1. General Procedures

All air- and moisture-sensitive reactions were carried out under an argon atmosphere. All reagents and solvents were obtained commercially and used without further purification except as noted. Dry 1,4-dioxane, dry trifluoroacetic acid (TFA), dry tetrahydrofuran (THF), dry toluene, and dry dichloromethane (DCM) were purchased from Adamas Reagent, Ltd. (Shanghai, China). The three intermediates TB-Br, TB-B, and TB were prepared by procedures from the literature [43,44]. 1H NMR (500 MHz) and 13C NMR (126 MHz) spectra were recorded on a Bruker Avance III NMR spectrometer, in deuterated chloroform (CDCl3).

3.2. Synthesis of Materials

3.2.1. Synthesis of ((4,10-Dimethyl-6H,12H-5,11-methanodibenzo[b,f][1,5]diazocine-2,8-diyl)bis(4,1-phenylene))bis((4-bromophenyl)methanone) (TB-BP-Br)

A mixture of TB-B (2.2 g, 4.5 mmol), bis(4-bromophenyl)methanone (4.59 g, 13.5 mmol), Pd(PPh3)4 (0.52 g, 0.45 mmol), THF (112.5 mL), and aqueous K2CO3 (1 M, 45 mL) was placed under vacuum and then backfilled with nitrogen three times before being heated in an oil bath at 60 °C for 16 h. After cooling to room temperature, the layers were separated. The aqueous phase was extracted with DCM (3 × 50 mL), the combined organic phases were dried over sodium sulfate and filtered, and then the solvent was removed. The residue was purified by column chromatography over silica using ethyl acetate/petroleum ether as eluent to afford TB-BP-Br as a white solid (1.15 g, yield: 33%). 1H NMR was as follows: (500 MHz, CDCl3) δ 7.81 (d, J = 8.3 Hz, 4H), 7.68–7.54 (m, 12H), 7.38 (s, 2H), 7.10 (s, 2H), 4.74 (d, J = 13.8 Hz, 2H), 4.45 (s, 2H), 4.16 (d, J = 16.8 Hz, 2H), and 2.53 (s, 6H).

3.2.2. Synthesis of (4,10-Dimethyldibenzo[b,f][1,5]diazocine-5,11(6H,12H)-diyl)bis((4-bromophenyl)methanone) (TB-PhBr)

Aluminum trichloride (1.60 g, 12 mmol), TB (1.25 g, 5 mmol), and DCM (50 mL) were added to the flask under vacuum and then backfilled with nitrogen three times at 0 °C. After stirring for 15 min, a solution of 4-bromobenzoyl chloride (2.42 g, 11 mmol) was added dropwise under stirring. The reaction was performed for three hours at 40 °C. Dilute hydrochloric acid was slowly added dropwise to the flask to terminate the reaction. The reaction was extracted with dichloromethane (3 × 40 mL). After collecting organic-phase, it was dried with anhydrous Na2SO4, the solvent was removed to give the crude product. Using petroleum ether and dichloromethane-mixed solvent as eluent, silica gel column separation and purification were performed, followed by vacuum-drying to obtain a white solid (1.51 g, yield: 50%). 1H NMR was as follows: (500 MHz, CDCl3) δ 7.29 (ddd, J = 10.7, 5.3, 2.1 Hz, 7H), 7.26 (d, J = 2.0 Hz, 1H), 6.96–6.90 (m, 4H), 6.83 (dd, J = 6.7, 2.2 Hz, 2H), 5.70 (d, J = 14.9 Hz, 2H), 4.39 (d, J = 15.0 Hz, 2H), and 2.13 (s, 6H).

3.2.3. Synthesis of ((4,10-dimethyl-6H,12H-5,11-methanodibenzo[b,f][1,5]diazocine-2,8-diyl)bis(4,1-phenylene))bis((4-(9,9-dimethylacridin-10(9H)-yl)phenyl)methanone) (TB-BP-DMAC)

A mixture of TB-BP-Br (1.15 g, 1.5 mmol), 9,9-dimethyl-9,10-dihydroacridine (0.94 g, 4.5 mmol), tri-tert-butylphosphine tetrafluoroborate (0.087 g, 0.3 mmol), sodium tert-butoxide (0.43 g, 4.5 mmol), palladium acetate (0.035 g, 0.15 mmol), and dry toluene was placed under vacuum and then backfilled with nitrogen three times before being heated in an oil bath at 110 °C for 18 h. After cooling to room temperature, the solvent was removed by vacuum rotary evaporation. The mixture was extracted with dichloromethane (3 × 40 mL). After collecting organic-phase, it was dried with anhydrous Na2SO4, the solvent was removed to give the crude product. Using petroleum ether and dichloromethane-mixed solvent as eluent, silica gel column separation and purification were performed, followed by vacuum drying to obtain a yellow-green solid (1.38 g, yield 90%). 1H NMR was as follows: (500 MHz, CDCl3) δ 8.11–8.04 (m, 4H), 7.95 (d, J = 8.4 Hz, 4H), 7.67 (d, J = 8.4 Hz, 4H), 7.54–7.45 (m, 8H), 7.41 (s, 2H), 7.14 (s, 2H), 6.99 (dtd, J = 21.2, 7.3, 1.4 Hz, 8H), 6.34 (dd, J = 8.0, 1.3 Hz, 4H), 4.75 (s, 2H), 4.46 (s, 2H), 4.19 (d, J = 16.6 Hz, 2H), 2.55 (s, 6H), and 1.70 (s, 12H). 13C NMR was as follows: (126 MHz, CDCl3) 195.42, 145.30, 140.49, 137.23, 135.64, 133.75, 132.60, 130.95, 130.74, 130.60, 129.98, 128.30, 126.77, 126.43, 125.37, 123.43, 121.09, 114.33, 55.20, 36.09, 31.13, and 17.45.

3.2.4. Synthesis of (4,10-dimethyldibenzo[b,f][1,5]diazocine-5,11(6H,12H)-diyl)bis((4-(9,9-dimethylacridin-10(9H)-yl)phenyl)methanone) (TB-DMAC)

It was prepared by the same procedure as that of TB-BP-DMAC, except TB-PhBr was used instead of TB-BP-Br. Vacuum drying was performed to obtain a white solid (yield: 85%). 1H NMR was as follows: (500 MHz, CDCl3) δ 7.83 (d, J = 7.3 Hz, 2H), 7.47 (d, J = 8.4 Hz, 4H), 7.41 (dd, J = 5.9, 3.3 Hz, 4H), 7.29 (t, J = 7.6 Hz, 2H), 7.11 (t, J = 7.9 Hz, 6H), 6.95–6.82 (m, 8H), 6.01 (dd, J = 6.0, 3.4 Hz, 4H), 5.74 (d, J = 13.3 Hz, 2H), 4.44 (d, J = 13.3 Hz, 2H), 2.16 (s, 6H), and 1.61 (d, J = 9.2 Hz, 12H). 13C NMR was as follows: (126 MHz, CDCl3) 168.84, 144.27, 143.08, 140.40, 135.02, 134.89, 134.80, 131.74, 131.08, 130.74, 130.44, 130.27, 128.37, 126.27, 125.18, 120.85, 113.86, 54.77, 35.97, 30.94, and 18.36.

3.3. Thermogravimetric Analyses

Thermogravimetric analyses were performed on a METTLER TOLEDO system (METTLER-TOLEDO International Inc., Columbus, Switzerland) with a heating rate of 10 °C/min under nitrogen.

3.4. Electrochemical Measurement

Cyclic voltammetry was performed at room temperature in anhydrous and argon-saturated dichloromethane solutions of 0.1 M tetrabutylammonium hexafluorophosphate and 1.0 mM investigated compounds with a CHI840D electrochemical analyzer (CH Instruments, Inc., Austin, TX, USA). Glassy carbon, platinum wire, and Ag/Ag+ (0.01 M of AgNO3 in acetonitrile) were selected as the working electrode, auxiliary electrode, and reference electrode, respectively. The ferrocenium/ferrocene couple was used as an internal standard.

3.5. Photophysical Measurements

UV–Vis absorption spectra were recorded with an Agilent Cary 5000 (Agilent Technologies, Santa Clara, CA, USA) UV–Vis spectrophotometer under ambient conditions. The absolute PLQYs were measured on an Edinburgh F900 spectrophotometer (Edinburgh Instruments Limited, Edinburgh, United Kingdom) equipped with an integrating sphere with the followsing properties: reference sample: BaSO4; film samples: 20 wt% of TB-BP-DMAC or TB-DMAC doped into PPF host, prepared by vacuum deposition; and excitation wavelength: 300 nm. The excitation spectra and emission spectra were collected in the wavelength ranges of 285–315 nm and 360–750 nm, respectively. Steady-state PL spectra were measured on an Edinburgh FLS980 (Edinburgh Instruments Limited, Edinburgh, UK) using a Xenon lamp as an excitation light source. The transient PL decay curves were measured on the same spectrophotometer (FLS980) in time-correlated single-photon counting (TCSPC) mode with an NT242-1K OPO laser (Ekspla, Vilnius, Lithuania) as an excitation light source. The time-resolved PL spectra were recorded on an Edinburgh LP980 spectrophotometer with an NT242-1K OPO laser excitation source.

3.6. Computational Methodology

All the calculations were carried out using the Gaussian 09 program package. The density functional theory (DFT) calculations at the BMK/6-31G** level were used to optimize the ground state geometries of the investigated compounds. Time-dependent density functional theory (TD-DFT) calculations were performed at the same level using the optimized ground state geometries. The spin–orbit couplings were calculated at B3LYP/6-31G** by ORCA. The electron density diagrams of molecular orbitals were generated using GaussView program. The partition orbital composition was analyzed by using the Multiwfn 2.4 program.

3.7. Device Fabrication and Characterization

Indium tin oxide (ITO) glass substrates were cleaned successively in an ultrasonic bath containing deionized water, acetone, and isopropanol. They were then blown dry with N2 and treated in UV/ozone ambient conditions for 15 min prior to film deposition. The organic materials were deposited onto the ITO-coated substrates at a rate of 1 Å s−1 under high vacuum (<1 × 10−4) by thermal evaporation in vacuum chamber. Then Liq and Al were successively deposited at a rate of 0.1 Å s−1 and 5 Å s−1, respectively. The electroluminescence (EL) spectra and Commission Internatinale de L’Eclairage coordinates (CIE) of the OLEDs were recorded by an integrated optoelectronic performance test system with a calibrated spectra radiometer (TOPCON SR-UL1R) (Topcon Corporation, Tokyo, Japan). The current efficiency and power efficiency were measured using Keithley 2400 source meter (Tektronix, Inc., Beaverton, OR, USA).

4. Conclusions

In summary, we have designed and synthesized two TADF emitters with D-A-D molecular configuration, named TB-BP-DMAC and TB-DMAC, by employing Tröger’s-base-fused benzophenone and amide acceptors as the electron acceptors, respectively. Our theoretical and experimental investigations revealed that the amide acceptor in TB-DMAC exhibits a much weaker strength compared to the benzophenone acceptor used in TB-BP-DMAC, which results in a significantly blue-shifted emission from green to deep blue, a higher emission efficiency, and a faster RISC process. TB-BP-DMAC and TB-DMAC exhibit green and deep-blue photoluminescence with PLQYs of 31.5% and 50.4% as well as decay fluorescence lifetimes of 2.55 and 2.28 μs in doped films, respectively. By using TB-DMAC as emitter, the doped and non-doped OLEDs have efficient deep-blue electroluminescence peaks at 449 and 453 nm with maximum EQEs of 6.1% and 5.7%, respectively, which far exceed the device performances of TB-BP-DMAC-based green-emitting devices. This indicates that substituted amide acceptors are a viable option for the design of high-performance deep-blue TADF materials. These results demonstrate the high potential of substituted amide acceptors in the construction of high-performance deep-blue TADF materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28124832/s1, Figure S1: 1H-NMR spectrum of TB-BP-Br (500 MHz, CDCl3); Figure S2. 1H-NMR spectrum of TB-PhBr (500 MHz, CDCl3); Figure S3. 1H-NMR spectrum of TB-BP-DMAC (500 MHz, CDCl3); Figure S4. 13C-NMR spectrum of TB-BP-DMAC (126 MHz, CDCl3); Figure S5. 1H-NMR spectrum of TB-DMAC (500 MHz, CDCl3); Figure S6. 13C-NMR spectrum of TB-DMAC (126 MHz, CDCl3); Figure S7. TGA curves of TB-BP-DMAC and TB-DMAC (The red dashed line marks 95% of the original sample weight); Figure S8. Cyclic Voltammograms of TB-BP-DMAC and TB-DMAC. Dashed vertical lines and intersections of tangent lines indicate the Eox values; Figure S9. Electron-density distribution of the excited states (the purple area denotes an increase in charge density, while the blue area denotes a decrease in charge density) and character; Figure S10. PL spectra (20 wt%-doped PPF films) measured at room temperature; Figure S11. Transient PL decay curves (0–125 ns time-range, prompt decay) of (a) TB-BP-DMAC and (b) TB-DMAC in 20 wt%-doped PPF film at room temperature. Inset shows the fitting function along with the fitting results for the lifetimes of the prompt component, as well as the instrumental response function (IRF); Figure S12. Transient PL decay curves of (a) TB-BP-DMAC and (b) TB-DMAC in 20 wt%-doped PPF film at room temperature. Inset shows the fitting function and fitting results of delayed component lifetimes; Figure S13. Transient PL decay curves of (a) TB-BP-DMAC and (b) TB-DMAC in 20 wt%-doped PPF film at different temperatures; Figure S14. The comparison of EL (20 wt%-doped OLEDs) and PL spectra (20 wt%-doped films) of TB-BP-DMAC and TB-DAMC. The EL spectra were recorded at 8 V. The PL spectra were obtained using 300 nm excitation. All spectra were obtained at room temperature; Figure S15. Molecular structures of the functional materials used in the OLEDs; Table S1: Summary of CV data and energy levels; Table S2. Calculated spin-orbit coupling (SOC) constants; Table S3. Fitting results of the transient PL decay curves of prompt fluorescence; Table S4. Fitting results of the transient PL decay curves of delayed fluorescence; Table S5. Summary of ΔEST values for some deep blue and green emitters. References [38,45,46,47,48,49,50,51] are cited in the supplementary materials.

Author Contributions

Conceptualization, Z.-L.W., X.-L.C. and C.-Z.L.; Methodology, Z.-L.W. and X.-L.C.; Software, Z.-L.W., X.L. and L.-Y.M.; Validation, Z.-L.W. and C.-Z.L.; Investigation, Z.-L.W.; Data curation, Z.-L.W.; Writing—original draft, Z.-L.W.; Writing—review & editing, X.-L.C. and C.-Z.L.; Supervision, X.-L.C. and C.-Z.L.; Project administration, X.-L.C. and C.-Z.L.; Funding acquisition, X.-L.C. and C.-Z.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant Nos. 52073286 and 21805281), the Natural Science Foundation of Fujian Province (Grant No. 2006L2005), the Fujian Science and Technology Innovation Laboratory for Optoelectronic Information of China (Grant Nos. 2021ZR132 and 2021ZZ115), and the Youth Innovation Foundation of Xiamen City (Grant Nos. 3502Z20206082 and 3502Z20206083).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data generated by this research are included in the article.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Tang, C.W.; Vanslyke, S.A. Organic electroluminescent diodes. Appl. Phys. Lett. 1987, 51, 913–915. [Google Scholar] [CrossRef]
  2. Baldo, M.A.; O’Brien, D.F.; You, Y.; Shoustikov, A.; Sibley, S.; Thompson, M.E.; Forrest, S.R. Highly efficient phosphorescent emission from organic electroluminescent devices. Nature 1998, 395, 151–154. [Google Scholar] [CrossRef]
  3. Uoyama, H.; Goushi, K.; Shizu, K.; Nomura, H.; Adachi, C. Highly efficient organic light-emitting diodes from delayed fluorescence. Nature 2012, 492, 234–238. [Google Scholar] [CrossRef]
  4. Song, W.; Lee, J.Y. Degradation Mechanism and Lifetime Improvement Strategy for Blue Phosphorescent Organic Light-Emitting Diodes. Adv. Opt. Mater. 2017, 5, 1600901. [Google Scholar] [CrossRef]
  5. Sivasubramaniam, V.; Brodkorb, F.; Hanning, S.; Loebl, H.P.; van Elsbergen, V.; Boerner, H.; Scherf, U.; Kreyenschmidt, M. Fluorine cleavage of the light blue heteroleptic triplet emitter FIrpic. J. Fluor. Chem. 2009, 130, 640–649. [Google Scholar] [CrossRef]
  6. Udagawa, K.; Sasabe, H.; Cai, C.; Kido, J. Low-driving-voltage blue phosphorescent organic light-emitting devices with external quantum efficiency of 30%. Adv. Mater. 2014, 26, 5062–5066. [Google Scholar] [CrossRef] [PubMed]
  7. Chen, X.K.; Kim, D.; Bredas, J.L. Thermally Activated Delayed Fluorescence (TADF) Path toward Efficient Electroluminescence in Purely Organic Materials: Molecular Level Insight. Acc. Chem. Res. 2018, 51, 2215–2224. [Google Scholar] [CrossRef]
  8. Adachi, C. Third-generation organic electroluminescence materials. Jpn. J. Appl. Phys. 2014, 53, 060101. [Google Scholar] [CrossRef]
  9. Wong, M.Y.; Zysman-Colman, E. Purely Organic Thermally Activated Delayed Fluorescence Materials for Organic Light-Emitting Diodes. Adv. Mater. 2017, 29, 1605444. [Google Scholar] [CrossRef] [Green Version]
  10. Dias, F.B.; Penfold, T.J.; Monkman, A.P. Photophysics of thermally activated delayed fluorescence molecules. Methods Appl. Fluoresc. 2017, 5, 012001. [Google Scholar] [CrossRef]
  11. Cui, L.-S.; Gillett, A.J.; Zhang, S.-F.; Ye, H.; Liu, Y.; Chen, X.-K.; Lin, Z.-S.; Evans, E.W.; Myers, W.K.; Ronson, T.K.; et al. Fast spin-flip enables efficient and stable organic electroluminescence from charge-transfer states. Nat. Photonics 2020, 14, 636–642. [Google Scholar] [CrossRef]
  12. Ji, S.-C.; Zhao, T.; Wei, Z.; Meng, L.; Tao, X.-D.; Yang, M.; Chen, X.-L.; Lu, C.-Z. Manipulating excited states via Lock/Unlock strategy for realizing efficient thermally activated delayed fluorescence emitters. Chem. Eng. J. 2022, 435, 134868. [Google Scholar] [CrossRef]
  13. He, J.-L.; Kong, F.-C.; Sun, B.; Wang, X.-J.; Tian, Q.-S.; Fan, J.; Liao, L.-S. Highly efficient deep-red TADF organic light-emitting diodes via increasing the acceptor strength of fused polycyclic aromatics. Chem. Eng. J. 2021, 424, 130470. [Google Scholar] [CrossRef]
  14. Godumala, M.; Choi, S.; Cho, M.J.; Choi, D.H. Thermally activated delayed fluorescence blue dopants and hosts: From the design strategy to organic light-emitting diode applications. J. Mater. Chem. C 2016, 4, 11355–11381. [Google Scholar] [CrossRef]
  15. Cai, X.; Su, S.-J. Marching Toward Highly Efficient, Pure-Blue, and Stable Thermally Activated Delayed Fluorescent Organic Light-Emitting Diodes. Adv. Funct. Mater. 2018, 28, 1802558. [Google Scholar] [CrossRef]
  16. Im, Y.; Byun, S.Y.; Kim, J.H.; Lee, D.R.; Oh, C.S.; Yook, K.S.; Lee, J.Y. Recent Progress in High-Efficiency Blue-Light-Emitting Materials for Organic Light-Emitting Diodes. Adv. Funct. Mater. 2017, 27, 1603007. [Google Scholar] [CrossRef]
  17. Bui, T.T.; Goubard, F.; Ibrahim-Ouali, M.; Gigmes, D.; Dumur, F. Recent advances on organic blue thermally activated delayed fluorescence (TADF) emitters for organic light-emitting diodes (OLEDs). Beilstein J. Org. Chem. 2018, 14, 282–308. [Google Scholar] [CrossRef] [Green Version]
  18. Bui, T.-T.; Goubard, F.; Ibrahim-Ouali, M.; Gigmes, D.; Dumur, F. Thermally Activated Delayed Fluorescence Emitters for Deep Blue Organic Light Emitting Diodes: A Review of Recent Advances. Appl. Sci. 2018, 8, 494. [Google Scholar] [CrossRef] [Green Version]
  19. Lee, J.-H.; Chen, C.-H.; Lee, P.-H.; Lin, H.-Y.; Leung, M.-k.; Chiu, T.-L.; Lin, C.-F. Blue organic light-emitting diodes: Current status, challenges, and future outlook. J. Mater. Chem. C 2019, 7, 5874–5888. [Google Scholar] [CrossRef]
  20. Liu, Y.; Li, C.; Ren, Z.; Yan, S.; Bryce, M.R. All-organic thermally activated delayed fluorescence materials for organic light-emitting diodes. Nat. Rev. Mater. 2018, 3, 18020. [Google Scholar] [CrossRef]
  21. Jeon, S.K.; Lee, H.L.; Yook, K.S.; Lee, J.Y. Recent Progress of the Lifetime of Organic Light-Emitting Diodes Based on Thermally Activated Delayed Fluorescent Material. Adv. Mater. 2019, 31, e1803524. [Google Scholar] [CrossRef]
  22. Jang, H.J.; Lee, J.Y.; Kim, J.; Kwak, J.; Park, J.-H. Progress of display performances: AR, VR, QLED, and OLED. J. Inf. Disp. 2020, 21, 1–9. [Google Scholar] [CrossRef] [Green Version]
  23. Gao, Y.; Wu, S.; Shan, G.; Cheng, G. Recent Progress in Blue Thermally Activated Delayed Fluorescence Emitters and Their Applications in OLEDs: Beyond Pure Organic Molecules with Twist D-π-A Structures. Micromachines 2022, 13, 2150. [Google Scholar] [CrossRef]
  24. Shi, C.; Liu, D.; Li, J.; He, Z.; Song, K.; Liu, B.; Wu, Q.; Xu, M. tert-Butyltriazine-Diphenylaminocarbazole based TADF materials: π-Bridge modification for enhanced kRISC and efficiency stability. Dyes Pigm. 2022, 204, 110430. [Google Scholar] [CrossRef]
  25. Konidena, R.K.; Lee, J.Y. Molecular Design Tactics for Highly Efficient Thermally Activated Delayed Fluorescence Emitters for Organic Light Emitting Diodes. Chem. Rec. 2019, 19, 1499–1517. [Google Scholar] [CrossRef]
  26. Che, W.; Xie, Y.; Li, Z. Structural Design of Blue-to-Red Thermally-Activated Delayed Fluorescence Molecules by Adjusting the Strength between Donor and Acceptor. Asian J. Org. Chem. 2020, 9, 1262–1276. [Google Scholar] [CrossRef]
  27. Huang, W.; Xie, L.; Feng, Q.; Wang, H.; Qian, Y.; Zhou, T. Recent Advances in Substituent Effects of Blue Thermally Activated Delayed Fluorescence Small Molecules. Acta Chim. Sin. 2021, 79, 557–574. [Google Scholar] [CrossRef]
  28. Zhao, W.; He, Z.; Lam, J.W.Y.; Peng, Q.; Ma, H.; Shuai, Z.; Bai, G.; Hao, J.; Tang, B.Z. Rational Molecular Design for Achieving Persistent and Efficient Pure Organic Room-Temperature Phosphorescence. Chem 2016, 1, 592–602. [Google Scholar] [CrossRef] [Green Version]
  29. Yun, J.H.; Lee, K.H.; Lee, J.Y. Benzoylphenyltriazine as a new acceptor of donor–acceptor type thermally-activated delayed-fluorescent emitters. J. Ind. Eng. Chem. 2021, 102, 226–232. [Google Scholar] [CrossRef]
  30. Hu, J.; Zhang, X.; Zhang, D.; Cao, X.; Jiang, T.; Zhang, X.; Tao, Y. Linkage modes on phthaloyl/triphenylamine hybrid compounds: Multi-functional AIE luminogens, non-doped emitters and organic hosts for highly efficient solution-processed delayed fluorescence OLEDs. Dyes Pigm. 2017, 137, 480–489. [Google Scholar] [CrossRef]
  31. Matsuoka, K.; Albrecht, K.; Nakayama, A.; Yamamoto, K.; Fujita, K. Highly Efficient Thermally Activated Delayed Fluorescence Organic Light-Emitting Diodes with Fully Solution-Processed Organic Multilayered Architecture: Impact of Terminal Substitution on Carbazole-Benzophenone Dendrimer and Interfacial Engineering. ACS Appl. Mater. Interfaces 2018, 10, 33343–33352. [Google Scholar] [CrossRef] [PubMed]
  32. Yang, Y.; Wang, S.; Zhu, Y.; Wang, Y.; Zhan, H.; Cheng, Y. Thermally Activated Delayed Fluorescence Conjugated Polymers with Backbone-Donor/Pendant-Acceptor Architecture for Nondoped OLEDs with High External Quantum Efficiency and Low Roll-Off. Adv. Funct. Mater. 2018, 28, 1706916. [Google Scholar] [CrossRef]
  33. Wu, L.; Wang, K.; Wang, C.; Fan, X.C.; Shi, Y.Z.; Zhang, X.; Zhang, S.L.; Ye, J.; Zheng, C.J.; Li, Y.Q.; et al. Using fluorene to lock electronically active moieties in thermally activated delayed fluorescence emitters for high-performance non-doped organic light-emitting diodes with suppressed roll-off. Chem. Sci. 2020, 12, 1495–1502. [Google Scholar] [CrossRef]
  34. Zou, S.N.; Peng, C.C.; Yang, S.Y.; Qu, Y.K.; Yu, Y.J.; Chen, X.; Jiang, Z.Q.; Liao, L.S. Fully Bridged Triphenylamine Derivatives as Color-Tunable Thermally Activated Delayed Fluorescence Emitters. Org. Lett. 2021, 23, 958–962. [Google Scholar] [CrossRef]
  35. Huang, F.; Wang, K.; Shi, Y.Z.; Fan, X.C.; Zhang, X.; Yu, J.; Lee, C.S.; Zhang, X.H. Approaching Efficient and Narrow RGB Electroluminescence from D-A-Type TADF Emitters Containing an Identical Multiple Resonance Backbone as the Acceptor. ACS Appl. Mater. Interfaces 2021, 13, 36089–36097. [Google Scholar] [CrossRef]
  36. Fan, X.C.; Wang, K.; Shi, Y.Z.; Chen, J.X.; Huang, F.; Wang, H.; Hu, Y.N.; Tsuchiya, Y.; Ou, X.M.; Yu, J.; et al. Managing Intersegmental Charge-Transfer and Multiple Resonance Alignments of D-A Typed TADF Emitters for Red OLEDs with Improved Efficiency and Color Purity. Adv. Opt. Mater. 2021, 10, 2101789. [Google Scholar] [CrossRef]
  37. Huang, J.; Nie, H.; Zeng, J.; Zhuang, Z.; Gan, S.; Cai, Y.; Guo, J.; Su, S.J.; Zhao, Z.; Tang, B.Z. Highly Efficient Nondoped OLEDs with Negligible Efficiency Roll-Off Fabricated from Aggregation-Induced Delayed Fluorescence Luminogens. Angew. Chem. Int. Ed. 2017, 56, 12971–12976. [Google Scholar] [CrossRef]
  38. Wu, X.; Zeng, J.; Peng, X.; Liu, H.; Tang, B.Z.; Zhao, Z. Robust sky-blue aggregation-induced delayed fluorescence materials for high-performance top-emitting OLEDs and single emissive layer white OLEDs. Chem. Eng. J. 2023, 451, 138919. [Google Scholar] [CrossRef]
  39. Liu, H.; Fu, Y.; Tang, B.Z.; Zhao, Z. All-fluorescence white organic light-emitting diodes with record-beating power efficiencies over 130 lm W−1 and small roll-offs. Nat. Commun. 2022, 13, 5154. [Google Scholar] [CrossRef] [PubMed]
  40. Cai, M.; Auffray, M.; Zhang, D.; Zhang, Y.; Nagata, R.; Lin, Z.; Tang, X.; Chan, C.-Y.; Lee, Y.-T.; Huang, T.; et al. Enhancing spin-orbital coupling in deep-blue/blue TADF emitters by minimizing the distance from the heteroatoms in donors to acceptors. Chem. Eng. J. 2021, 420, 127591. [Google Scholar] [CrossRef]
  41. Rúnarsson, Ö.V.; Artacho, J.; Wärnmark, K. The 125th Anniversary of the Tröger’s Base Molecule: Synthesis and Applications of Tröger’s Base Analogues. Eur. J. Org. Chem. 2012, 2012, 7015–7041. [Google Scholar] [CrossRef]
  42. Neogi, I.; Jhulki, S.; Ghosh, A.; Chow, T.J.; Moorthy, J.N. Amorphous host materials based on Troger’s base scaffold for application in phosphorescent organic light-emitting diodes. ACS Appl. Mater. Interfaces 2015, 7, 3298–3305. [Google Scholar] [CrossRef] [PubMed]
  43. Kiehne, U.; Bruhn, T.; Schnakenburg, G.; Frohlich, R.; Bringmann, G.; Lutzen, A. Synthesis, resolution, and absolute configuration of difunctionalized Troger’s base derivatives. Chemistry 2008, 14, 4246–4255. [Google Scholar] [CrossRef]
  44. Didier, D.; Tylleman, B.; Lambert, N.; Vande Velde, C.M.L.; Blockhuys, F.; Collas, A.; Sergeyev, S. Functionalized analogues of Tröger’s base: Scope and limitations of a general synthetic procedure and facile, predictable method for the separation of enantiomers. Tetrahedron 2008, 64, 6252–6262. [Google Scholar] [CrossRef]
  45. Wada, Y.; Nakagawa, H.; Matsumoto, S.; Wakisaka, Y.; Kaji, H. Organic light emitters exhibiting very fast reverse intersystem crossing. Nat. Photonics 2020, 14, 643–649. [Google Scholar] [CrossRef]
  46. Chen, X.-L.; Tao, X.-D.; Wei, Z.; Meng, L.; Lin, F.-L.; Zhang, D.-H.; Jing, Y.-Y.; Lu, C.-Z. Thermally Activated Delayed Fluorescence Amorphous Molecular Materials for High-Performance Organic Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2021, 13, 46909–46918. [Google Scholar] [CrossRef]
  47. Gao, H.; Shen, S.; Qin, Y.; Liu, G.; Gao, T.; Dong, X.; Pang, Z.; Xie, X.; Wang, P.; Wang, Y. Ultrapure Blue Thermally Activated Delayed Fluorescence (TADF) Emitters Based on Rigid Sulfur/Oxygen-Bridged Triarylboron Acceptor: MR TADF and D-A TADF. J. Phys. Chem. Lett. 2022, 13, 7561–7567. [Google Scholar] [CrossRef]
  48. Liu, B.; Li, J.; Liu, D.; Mei, Y.; Lan, Y.; Song, K.; Li, Y.; Wang, J. Electron-withdrawing bulky group substituted carbazoles for blue TADF emitters: Simultaneous improvement of blue color purity and RISC rate constants. Dyes Pigm. 2022, 203, 110329. [Google Scholar] [CrossRef]
  49. Lee, Y.H.; Park, S.; Oh, J.; Shin, J.W.; Jung, J.; Yoo, S.; Lee, M.H. Rigidity-Induced Delayed Fluorescence by Ortho Donor-Appended Triarylboron Compounds: Record-High Efficiency in Pure Blue Fluorescent Organic Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2017, 9, 24035–24042. [Google Scholar] [CrossRef] [PubMed]
  50. Xia, Y.; Li, J.; Chen, X.; Li, A.; Guo, K.; Chen, F.; Zhao, B.; Chen, Z.; Wang, H. Molecular Engineering of Push-Pull Diphenylsulfone Derivatives towards Aggregation-Induced Narrowband Deep Blue Thermally Activated Delayed Fluorescence (TADF) Emitters. Chemistry 2022, 28, e202202434. [Google Scholar] [CrossRef]
  51. Yoo, J.-Y.; Choi, Y.J.; Kim, K.W.; Ha, T.H.; Lee, C.W. Improved device efficiency and lifetime of green thermally activated delayed fluorescence materials with multiple donors and cyano substitution. Dyes Pigm. 2023, 214, 111200. [Google Scholar] [CrossRef]
Scheme 1. Synthetic route of TB-BP-DMAC and TB-DMAC.
Scheme 1. Synthetic route of TB-BP-DMAC and TB-DMAC.
Molecules 28 04832 sch001
Figure 1. (a) The structure and torsion angle of TB-BP-DMAC and TB-DMAC in ground state, respectively; (b) calculated HOMO, LUMO, S1 and T1 distributions, and SOC constants of TB-BP-DMAC and TB-DMAC.
Figure 1. (a) The structure and torsion angle of TB-BP-DMAC and TB-DMAC in ground state, respectively; (b) calculated HOMO, LUMO, S1 and T1 distributions, and SOC constants of TB-BP-DMAC and TB-DMAC.
Molecules 28 04832 g001
Figure 2. (a) Absorption and PL spectra (λex = 300 nm) of TB-BP-DMAC and TB-DMAC in toluene (c = 1 × 10−5 M) measured at room temperature (a dashed arrow and a solid arrow indicate the absorption and PL spectra, respectively). Inset: the photographs of TB-BP-DMAC (left) and TB-DMAC (right) in toluene (1 × 10−5 M) under excitation of 365 nm UV light, taken with an iPhone; (b) transient PL decay curves in 20 wt%-doped PPF films at room temperature; (c,d) time-resolved PL spectra in 20 wt%-doped PPF films at 77 K (PF: prompt fluorescence; Phos: phosphorescence).
Figure 2. (a) Absorption and PL spectra (λex = 300 nm) of TB-BP-DMAC and TB-DMAC in toluene (c = 1 × 10−5 M) measured at room temperature (a dashed arrow and a solid arrow indicate the absorption and PL spectra, respectively). Inset: the photographs of TB-BP-DMAC (left) and TB-DMAC (right) in toluene (1 × 10−5 M) under excitation of 365 nm UV light, taken with an iPhone; (b) transient PL decay curves in 20 wt%-doped PPF films at room temperature; (c,d) time-resolved PL spectra in 20 wt%-doped PPF films at 77 K (PF: prompt fluorescence; Phos: phosphorescence).
Molecules 28 04832 g002
Figure 3. OLEDs based on TB-DMAC with various doping concentrations (in weight): (a) current density/voltage/luminance characteristics; (b) external quantum efficiency (EQE), power efficiency (PE), and current efficiency (CE) versus luminance characteristics.
Figure 3. OLEDs based on TB-DMAC with various doping concentrations (in weight): (a) current density/voltage/luminance characteristics; (b) external quantum efficiency (EQE), power efficiency (PE), and current efficiency (CE) versus luminance characteristics.
Molecules 28 04832 g003
Figure 4. (a) Energy level diagram of the OLEDs; (b) current density/voltage/luminance characteristics; (c) current efficiency (CE)/luminance-power efficiency (PE) curves; (d) EL spectra (inset) of the devices and external quantum efficiency (EQE) vs. luminance relationships of devices.
Figure 4. (a) Energy level diagram of the OLEDs; (b) current density/voltage/luminance characteristics; (c) current efficiency (CE)/luminance-power efficiency (PE) curves; (d) EL spectra (inset) of the devices and external quantum efficiency (EQE) vs. luminance relationships of devices.
Molecules 28 04832 g004
Table 1. Photophysical data of the investigated compounds in 20 wt%-doped PPF film.
Table 1. Photophysical data of the investigated compounds in 20 wt%-doped PPF film.
CompoundλPL a
[nm]
ΦPL b
[%]
ΦPFDF c
[%]
τPFDF d
[ns/μs]
ES1/ET1/ΔEST e
[eV]
k r s / k n r s /kISCf
[107s−1]
kRISC g
[105s−1]
CIE1931
[x,y]
TB-BP-DMAC51731.521.2/10.312.1/2.552.81/2.68/0.1281.75/3.81/2.695.83(0.30, 0.49)
TB-DMAC46750.436.1/14.311.6/2.283.12/2.96/0.1583.11/3.06/2.436.13(0.18, 0.21)
a the wavelength at photoluminescence maximum; b overall PLQY; c ΦPF and ΦDF are the quantum yields of prompt fluorescence and delayed fluorescence, respectively; d τPF and τDF are the lifetimes of prompt fluorescence and delayed fluorescence, respectively; e energy levels of S1 and T1 state were estimated from the onsets of time-resolved PL spectra at 77 K (Figure 2c,d); f  k r s , k n r s , and kISC represent the radiative, nonradiative rate constant, and intersystem-crossing (ISC) rate constant of the S1 states, respectively; g kRISC refers to the rate constants of reverse ISC.
Table 2. Summary of device performances.
Table 2. Summary of device performances.
DeviceDoping ConcentrationλEL a
[nm]
Von b
[V]
Lmax c
[cd/m2]
EQEmax d
[%]
CEmax e
[cd/A]
PEmax f
[lm/W]
CIE1931 g
[x,y]
TB-DMAC10%4473.28065.46.66.6(0.17, 0.13)
20%4492.916846.19.19.9(0.17, 0.15)
30%4523.014746.09.09.5(0.17, 0.16)
100% (non-doped)4533.15525.710.411.2(0.18, 0.19)
TB-BP-DMAC20%5085.335243.911.06.9(0.28, 0.51)
a the wavelength at electroluminescence maximum (recorded at 8 V); b turn-on voltage at 1 cd/m2; c maximum luminance; d EQE maximum value; e current efficiency maximum value; f power efficiency maximum value; g Commission Internatinale de L’Eclairage coordinates measured at 8 V.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wu, Z.-L.; Lv, X.; Meng, L.-Y.; Chen, X.-L.; Lu, C.-Z. Tröger’s Base-Derived Thermally Activated Delayed Fluorescence Dopant for Efficient Deep-Blue Organic Light-Emitting Diodes. Molecules 2023, 28, 4832. https://doi.org/10.3390/molecules28124832

AMA Style

Wu Z-L, Lv X, Meng L-Y, Chen X-L, Lu C-Z. Tröger’s Base-Derived Thermally Activated Delayed Fluorescence Dopant for Efficient Deep-Blue Organic Light-Emitting Diodes. Molecules. 2023; 28(12):4832. https://doi.org/10.3390/molecules28124832

Chicago/Turabian Style

Wu, Ze-Ling, Xin Lv, Ling-Yi Meng, Xu-Lin Chen, and Can-Zhong Lu. 2023. "Tröger’s Base-Derived Thermally Activated Delayed Fluorescence Dopant for Efficient Deep-Blue Organic Light-Emitting Diodes" Molecules 28, no. 12: 4832. https://doi.org/10.3390/molecules28124832

Article Metrics

Back to TopTop