Next Article in Journal
Study on the Manipulation Strategy of Metallic Microstructures Based on Electrochemical-Assisted Method
Next Article in Special Issue
Effects of Thermal Treatment on DC Voltage-Driven Color Conversion in Organic Light-Emitting Diode
Previous Article in Journal
Application of Ternary Nanoparticles in the Heat Transfer of an MHD Non-Newtonian Fluid Flow
Previous Article in Special Issue
Optoelectronic Properties of MAPbBr3 Perovskite Light-Emitting Diodes Using Anti-Solvent and PEDOT:PSS/PVK Double-Layer Hole Transport Layers
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Blue Thermally Activated Delayed Fluorescence Emitters and Their Applications in OLEDs: Beyond Pure Organic Molecules with Twist D-π-A Structures

1
State Key Laboratory of Synthetic Chemistry, Department of Chemistry, The University of Hong Kong, Hong Kong, China
2
Institute of Functional Material Chemistry and National & Local United Engineering Lab for Power Battery, Faculty of Chemistry, Northeast Normal University, Changchun 130024, China
3
HKU Shenzhen Institute of Research and Innovation, Shenzhen 518053, China
*
Authors to whom correspondence should be addressed.
Micromachines 2022, 13(12), 2150; https://doi.org/10.3390/mi13122150
Submission received: 22 October 2022 / Revised: 26 November 2022 / Accepted: 28 November 2022 / Published: 5 December 2022
(This article belongs to the Special Issue Organic Light Emitting Diodes (OLEDs))

Abstract

:
Thermally activated delayed fluorescence (TADF) materials, which can harvest all excitons and emit light without the use of noble metals, are an appealing class of functional materials emerging as next-generation organic electroluminescent materials. Triplet excitons can be upconverted to the singlet state with the aid of ambient thermal energy under the reverse inter-system crossing owing to the small singlet–triplet splitting energy (ΔEST). This results from a specific molecular design consisting of minimal overlap between the highest occupied molecular orbital and the lowest unoccupied molecular orbital, due to the spatial separation of the electron-donating and electron-releasing part. When a well-designed device structure is applied, high-performance blue-emitting TADF organic light-emitting diodes can be realized with an appropriate molecular design. Unlike the previous literature that has reviewed general blue-emitting TADF materials, in this paper, we focus on materials other than pure organic molecules with twist D-π-A structures, including multi-resonance TADF, through-space charge transfer TADF, and metal-TADF materials. Cutting-edge molecules with extremely small and even negative ΔEST values are also introduced as candidates for next-generation TADF materials. In addition, OLED structures used to exploit the merits of the abovementioned TADF emitters are also described in this review.

1. Introduction

Organic light-emitting diodes (OLEDs) and the use of organic materials in light-emitting technology have caught the attention of researchers since Tang and VanSlyke’s pioneering work in 1987 [1] owing to their extraordinary merits, such as high efficiency, light weight, flexibility, and fast response [2,3,4,5,6]. A typical bottom-emitting OLED has a sandwich device structure, in which organic functional layer(s) are sandwiched between a transparent anode and a metal cathode. For high-efficiency OLEDs, the sandwiched functional layers usually consist of a hole-transporting layer (HTL), an emissive layer (EML), and an electron-transporting layer (ETL). In some cases, especially in phosphorescent and thermally activated delayed fluorescent (TADF) OLEDs, one or more auxiliary layers, such as a hole-blocking layer (HBL), a hole-injection layer (HIL), an electron-blocking layer (EBL), and an electron-injection layer (EIL), are added in the stack to further improve device performance. HBLs and HILs are used to confine carriers and excitons in the EML to avoid leaking current and the unwanted recombination of excitons outside the EML, and HIL and EIL are used to facilitate the injection of holes from the anode and electrons from the cathode to lower the driving voltage of the OLED [7]. In some cases, inorganic materials can be employed as HILs and/or EILs, such as MoO3, V2O5, LiF, Cs2CO3, etc. [8]. Such a sophisticated device structure is designed to fully exploit the potential of the emitting material, whose photoluminescent quantum yield (PLQY) and decay lifetime determine the performance of a well-designed OLED [9,10,11,12,13,14].
Light-emitting materials used in OLEDs have advanced rapidly in the last three decades, from fluorescent, phosphorescent, and triplet–triplet annihilation (TTA) materials to TADF materials [15]. Organic electroluminescence (EL) results from the radiative deactivation of excitons from the excited state to the ground state, which can occur with different spin multiplicities such as singlet or triplet excited states, resulting in fluorescence or phosphorescence. When OLEDs work under electrical excitation, singlet and triplet excitons are formed in the EML at a ratio of 1:3 [16], according to spin statistics. The quantum efficiency (QE) of an OLED, which refers to the numerical ratio of photons to injected electron-hole pairs [17], is used to assess its electricity-photon conversion efficiency, which can be further subdivided into internal QE (IQE) and external QE (EQE). IQE, the numerical ratio of total photons generated within the EML to the charge pairs injected, is directly determined by the emitting material [18]. In contrast, EQE is the numerical ratio of total photons emitting out of the device to the charge pairs injected. The chemical structure of the emitter and the associated functional materials can have significant impacts on the device efficiency, since photons are created within and emitted from the EML. IQE of fluorescent OLEDs is limited to 25% in traditional fluorescent emitters, which can only exploit singlet excitons for light emission. On the other hand, phosphorescent emitters that contain noble metals such as platinum, iridium, or osmium can harvest both singlet and triplet excitons through strong spin–orbit coupling, leading to high IQEs of up to 100% in phosphorescent OLEDs [19]. Nonetheless, further application of phosphorescent emitters is limited by the low abundance and high price of noble metals [20]. TTA fluorescent emitters can only achieve a maximum IQE of 62.5% by converting two triplet excitons into one singlet exciton, and their emission is mostly blue [21].
Alternatively, Yersin et al. [22] and Adachi et al. [23] developed inexpensive metal complexes and purely organic molecules, respectively, that are capable of harvesting both singlet and triplet excitons for emission, known as TADF. By reducing the overlap between the molecule’s highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO), small singlet–triplet splitting (ΔEST) can be achieved for efficient TADF. After singlet and triplet excitons are formed in the EML, the accumulated triplet excitons in T1 are transferred to S1 via a reverse inter-system crossing (RISC) process enhanced by thermal activation in the TADF mechanism [24]. By activating an effective up-conversion process of RISC from an excited triplet state to an excited singlet state, the TADF mechanism can also potentially achieve an IQE of 100% [25,26,27]. Compounds with suitable donor–acceptor building blocks are usually designed as TADF emitters, in which intramolecular charge transfer (ICT) with a small ΔEST is achieved by donor and acceptor moieties in the same molecule, allowing an effective RISC process to collect singlet emissions by converting triplets to singlets. In addition, twisted molecules, such as phenyl linkers between donors and acceptors, and a clear separation between the HOMO and LUMO distribution benefit TADF [7,9,28,29].
Along with judicious molecular designs, the performance of TADF-OLEDs has made significant progress. Remarkably, researchers have committed their attention to pure blue TADF emitters, which promise to be the final hurdle blocking the achievement of OLEDs for low-cost, highly efficient, and stable display illumination devices [30]. As one of the three primary colors, emitting blue is necessary for both display and lighting applications. However, neither fluorescent nor phosphorescent emitters are suitable candidates for blue OLEDs due to the low efficiency of the former and unsatisfactory stability of the latter. In this regards, TADF emitters could be a possible solution for low-cost, highly efficient, and stable blue OLEDs due to their potential high efficiency and stability. Early reviews have introduced the design concepts, the relationship between molecular structure and photophysical properties, and PL and EL performances of blue TADF materials, most of which are pure organic molecules with twist D-π-A structures [9]. However, new TADF emitters designed with novel concepts reported very recently have been sparsely covered in reviews. In this contribution, we focus on blue TADF materials possessing molecular structures other than pure organic with twist D-π-A structures. Recent advances of multiple MR-TADF materials are reviewed in Section 2, especially blue-emitting ones, that potentially address the issues of efficiency roll-off, quenching at high concentrations, and limited operational lifetimes that exist in early MR-TADF emitters. Newly emerged blue TADF emitters from through-space charge transfer (TSCT) processes are introduced in Section 3. A recent development in high-performance blue-emitting metal-TADF materials, especially those based on gold and copper, is briefly presented in Section 4. The outlook and perspective of next-generation TADF emitters with extremely small and even negative ΔEST are predicted in Section 5. In addition, OLED structures that are used to fully exploit the merits of the above-mentioned TADF emitters are also emphasized in this review.

2. Blue MR-TADF Emitters

Unlike the continuous enhancement in their efficiency, the color purity of blue TADF emitters has barely been improved due to their broad emission spectrum caused by a large Stokes shift, which is a result of the twisted conformation of the D-A molecules and the structural relaxation in the excited states. Alternatively, TADF can also be achieved in molecules with an MR effect, such as P- and B-doped rigid polycyclic aromatic hydrocarbons (PAHs) [31,32,33], from which narrowband emission with high color purity can be achieved due to the suppressed structural relaxation and lower vibronic coupling between the S1-S0 transition and the stretching/scissoring vibrations. Since the first MR-TADF material was reported by Hatakeyama and co-workers in 2016 [31], significant progress has been made in terms of efficiency and color purity [34]. Typically, OLEDs with the blue MR-TADF emitter ν-DABNA have demonstrated an emission maximum at 469 nm with an FWHM of 18 nm and an EQEmax of 34.4% [35]. In addition, owing to their small Stokes shift, MR-TADF materials have been widely used as terminal emitters in hyper-fluorescent OLEDs to improve the color purity of devices with phosphorescent or TADF emitters only [36,37,38,39,40]. Nonetheless, pronounced efficiency roll-off, severe quenching at high concentrations, and limited device stabilities remain formidable challenges to developing efficient blue MR-TADF emitters. In this section, we focus on the reports that are likely to address these issues.
The efficiency roll-off of OLEDs based on MR-TADF emitters is usually severe when compared to devices with conventional TADF emitters, due to the longer decay lifetime of the former. The long decay lifetime of MR-TADF emitters is mainly caused by their moderate ΔEST and RISC rate constant, which leads to an increased triplet-involved annihilation process [41,42,43,44]. Cao, Yang and colleagues presented a series of narrowband deep blue MR-TADF emitters (BN1-BN3, Figure 1) possessing a gradually enlarged ring-fused structure, which could simultaneously reduce ΔEST, enhance fosc, and retain narrow linewidth of MR-TADF emitters [45,46]. By measuring the threshold energy of low-temperature fluorescence and phosphorescence spectra, the ΔESTs of BN1, BN2, and BN3 were deduced to be 0.20, 0.16, and 0.15 eV, along with gradually accelerated RISC rates of 1.3, 2.6, and 25.5 × 104 s−1 at room temperature. Together with the shorter lifetime (17.8 µs) of the delayed fluorescent and a higher PLQY of 0.98 in DBFPO thin film, BN3 displayed improved EL performance in hyper-fluorescent OLEDs with a structure of ITO/HAT-CN (5 nm)/TAPC (30 nm)/TCTA(15 nm)/mCBP (10 nm)/DBFPO: 25 wt% 3Cz2BN: 1 wt% BN1–BN3 (25 nm)/DBFPO (15 nm)/ANT-BIZ (30 nm)/Liq/Al. In these devices, BN1–BN3 were used as terminal emitters, while 3Cz2BN was used as a TADF sensitizer. The high EQEmax of 37.6% and EQE1000 of 26.2%, as well as the deep blue color with CIE coordinates of (0.14, 0.08) of the device with BN1, indicated that chromophore π-extension could decrease the decay lifetime of MR-TADF emitters without weakening its PLQY and broadening its FWHM, and may eventually address the efficiency roll-off issue.
For most MR-TADF molecules, the functional groups are mostly simple, rigid, and finitely extended to avoid spectral broadening. Such highly planar configurations may lead to intermolecular interaction-induced quenching, such as triplet–triplet annihilation (TTA) and triplet-polaron quenching (TPQ), at high dopant concentrations. Thus, extremely low dopant concentrations (usually less than 5 wt%) of MR-TADF emitters are used in OLEDs, which may lead to a relatively low maximum luminance and substantial efficiency roll-off at high luminance. To address this issue, host-featured segments have been introduced to form self-host MR-TADF emitters. In addition to averting intermolecular interactions, transporting abilities can also be readily tuned in such self-host emitters [44,47,48]. As shown in Figure 2, Xu and co-workers reported an ambipolar self-host featured MR-TADF emitter, tCBNDADPO, in which a B-N framework (tCBN) was substituted with an ambipolar A-D-A host segment (DADPO). The D-A type PO structures of DADPO significantly improved the TADF properties of tCBNDADPO, especially the accelerated singlet radiative rate constant of 2.11 × 108 s−1, and exponentially reduced nonradiative rate constants, leading to a high PLQY of 99% for thin films with a high doping concentration of 30%. At the same time, narrowband blue emission with an FWHM of ≈ 28 nm was preserved in such film conditions. Taking advantage of the self-host feature, high EL performance of tCBNDADPO can be achieved in the OLED with a rather simple device structure of ITO/MoO3 (6 nm)/mCP (50 nm)/tCBNDADPO: DBFDPO (25 nm)/DBFDPO (40 nm)/LiF (1 nm)/Al (100 nm). A high EQEmax of 30.8% with coordinates of (0.14, 0.22) was achieved at a high concentration of 30%. Despite the efficiency roll-off (34.7% at a luminance of 300 cd m−2) still being unsatisfactory, the high concentration tolerance and the simplified device structure utilized may pave a path to the practical application of self-host MR-TADF emitters.
In addition to the serious efficiency roll-off, the relatively larger ΔEST and lower RISC process of MR-TADF emitters also make it difficult to control the triplet exciton density in the device, leading to a short device lifetime. To address this issue, Park and colleagues demonstrated that by the introduction of an additional blocking group to manage intermolecular interaction and the concentration quenching effect, a high efficiency and stable MR-TADF OLED is possible [49]. With an additional di-tert-butylphenyl (dtB) substituent on 2,12-di-tert-butyl-5,9-bis(4-(tert-butyl)phenyl)-5,9-dihydro-5,9-diaza-13b-boranaphtho[3,2,1-de]anthracene (t-DABNA) [50], 2,12-di-tert-butyl-5,9-bis(4-(tert-butyl)phenyl)-7-(3,5-di-tert-butylphenyl)-5,9-dihydro-5,9-diaza-13b-boranaphtho[3,2,1-de]anthracene (t-DABNA-dtB) was synthesized, as shown in Figure 3. With the increase in doping concentration in mCBP films, the PLQY of t-DABNA gradually decreased from 87% at 3 wt% to 54% at 10 wt%. At the same doping concentrations, the PLQYs of t-DABNA-dtB were 97 and 78%, respectively. The high PLQY of t-DABNA-dtB at high concentrations is attributable to its larger HOMO–LUMO overlap and suppressed concentration quenching, which is a result of the bulky substituent. By using t-DABNA or t-DABNA-dtB as the emitting dopant in OLEDs with an architecture of ITO /BCFN:HATCN (40 nm, 30 wt%)/BCFN (10 nm)/mCBP (10 nm)/mCBP:mCBP-CN:Emitter (30 nm, 3 wt%)/DBFTRz (5 nm)/ZADN (20 nm)/LiF (1.5 nm)/Al (200 nm), high EQEmax of over 25% were achieved. To further improve the device performance, especially the operational lifetime and efficiency roll-off, a sophisticated tandem device structure was adopted. High EQEmax of 30.1%, which slightly decreased to 28.8% at high luminance of 1000 cd m−2, was achieved in the tandem device with t-DABNA-dtB, together with a pure blue emission color with CIE coordinates of (0.116, 0.116) at 10,000 cd m−2. As shown in Figure 4, a long device lifetime, LT95, of 502 h at an initial luminance of 1000 cd m−2 was achieved in the tandem device. This lifetime is one of the best results for blue OLED lifetimes as reported in the literature, showing a bright future for blue MR-TADF in practical applications.

3. Blue TSCT-TADF Emitters

In addition to emitters in a twisted D-π-A structure via through-bond charge transfer (TBCT), TADF can also be the result of TSCT processes that usually take place between a pair of exciplex-forming D and A molecules, which is known as inter-TSCT. On the other hand, TSCT-TADF emitters (intra-TSCT) are molecules where the D and A segments are linked in a V or U shape, in which TSCT between D and A segments is possible. When compared with TBCT emitters, TSCT emitters have reduced electronic communication and coupling between D and A moieties, which benefits the elevation of the singlet and triplet levels and thus blue emission. At the same time, high PLQYs could easily be achieved in the TSCT molecules with intrinsic intramolecular noncovalent interactions [51,52]. The distance and orientation between D and A segments, determined by the linker, is a crucial factor in the excited state dynamics of a TSCT-TADF emitter [53]. Of the various types of alignment of D and A segments in TSCT molecules, face-to-face alignment with short distance is the most favorable for efficient TADF emission. However, only a few blue examples have been reported among high-performance OLEDs with TSCT-TADF emitters [54,55,56,57,58,59,60].
As shown in Figure 5, Wang and co-workers demonstrated a TADF polymer in which the TSCT effect takes place between pendant D and A units on a nonconjugated polyethylene backbone [59]. Unlike conventional conjugated D-A polymers, there is no direct conjugation between D and A units in the TSCT-TADF polymer because of its nonconjugated architecture, which is favorable for blue emission. Two polymers with the same A unit of 2,4,6-triphenyl-1,3,5-triazine (TRZ) but different D units of 9,9-dimethyl-10-phenyl-acridan (Ac) or 9,9-bis(1,3-ditert-butylphenyl)-10-phenyl-acridan (TBAc) were synthesized and compared. As a result of its steric 1,3-ditert-butylphenyl groups, the acridan unit is separated from the triazine unit in the TBAc-based polymer, while acridan approaches triazine in the Ac-based polymer, which determines the different D-A distance in different polymers. As expected, since TSCT is highly sensitive to the D-A distance, the distinct TSCT effect and TADF feature were only displayed in the Ac-based polymer, where a small ΔEST of 0.019 eV and high PLQY up to 60% in the film state were achieved. On the other hand, no TSCT effect and only fluorescence emission were found in the TBAc-based counterpart. When the polymer with 95 mol % Ac and 5 mol % TRZ (P-Ac95-TRZ05) was used as the EML in the device with a structure of ITO/PEDOT: PSS (40 nm)/P-Ac95-TRZ05 (40 nm)/TSPO1 (8 nm)/TmPyPB(42 nm)/LiF (1 nm)/Al (100 nm), blue emission with CIE coordinates of (0.176, 0.269) and an EQEmax of 12.1% were achieved, which is the first example of blue TSCT-TADF polymer for solution-processed OLEDs.
Zhang, Duan and colleagues adopted a xanthene bridge to construct space-confined face-to-face D-A alignment and minimize the D-A distance to 2.7–2.8 Å to strengthen the electronic interaction between weak D and A, which is required for efficient blue emission [39]. The targeted TSCT-TADF emitters are shown in Figure 6, exhibiting peaks around 460 nm, PLQYs > 90%, and a krs of nearly 107 s−1. The EL of such blue emitters was investigated in OLEDs with a structure of ITO/HAT-CN (5 nm)/TAPC (30 nm)/TCTA (10 nm)/mCP (10 nm)/PPF: 30 wt% emitters (24 nm)/PPF (10 nm)/BPhen (30 nm)/LiF (0.5 nm)/Al (150 nm). Blue emission with an EQEmax of 27.8% and CIE coordinate of (0.17, 0.29) were achieved when dCz-Xo-TRZ was used as the emitter. dCz-Xo-TRZ was further used as a sensitizer in hyper-fluorescent OLEDs with the well-known MR-TADF emitter, v-DABNA, as the terminal emitter. A high EQEmax of 34.7% with a CIEy coordinate of 0.15 and FWHM of 19 nm was realized in the resulting device.
Despite the high efficiency achieved in blue TSCT-TADF emitters [39,55,57], their color purity is still far from meeting the ultrapure-blue requirements of CIE (0.14, 0.08) defined by the NTSC, as well as CIE (0.131, 0.046) for the standard of Rec. 2020 [60], due to the large Stokes shift and broad emission with large FWHM in conventional TADF emitters. Alternatively, as mentioned in Section 2, the FWHM can be effectively narrowed by using MR-TADF emitters due to their rigid polycyclic aromatic framework. Ren and co-workers designed TSCT-TADF emitters by combining acridine derivatives and electron-withdrawing boron/oxygen heterocycles [60]. As depicted in Figure 7, by controlling the intramolecular stacking of rigid heteroaromatic compounds, deep blue emission with a narrow FWHM could be achieved. By using AC-BO, QAC-BO, and Cz-BO as emitters in OLEDs with a structure of ITO/HAT-CN (20 nm)/TAPC (40 nm)/mCP (10 nm)/20% emitters: DPEPO (20 nm)/DPEPO (10 nm)/TmPyPB(50 nm)/LiF (1 nm)/Al (150 nm), the EL properties of AC-BO, QAC-BO, and Cz-BO were investigated. With decreasing stacking distance, emission maxima of AC-BO, QAC-BO, and Cz-BO were significantly blue-shifted from 456 to 412 nm, while the FWHM narrowed from 71 to 43 nm, leading to CIE coordinates of (0.147, 0.122), (0.145, 0.076), and (0.163, 0.034), respectively. The device with AC-BO shows an EQEmax of 19.3%, the best performance among all deep blue TSCT-TADF emitters so far [60]. The device with QAC-BO achieved an EQEmax of 15.8%, making it the first high-efficiency TSCT-TADF material to meet the ultrapure blue requirements of CIE (0.14, 0.08) defined by the NTSC. Nonetheless, the relatively low maximum luminance of less than 1000 cd m−2 and pronounced efficiency roll-off of more than 60% at 100 cd m−2 hinders the practical application of these deep blue emitters.

4. Blue Metal-TADF Emitters

Despite their rapid development, the long-term stability of OLEDs with organic TADF emitters is still inferior when compared with those with phosphorescent or fluorescent emitters, due to the long-lived triplet excitons of organic TADF molecules resulting from their inefficient spin–orbital coupling (SOC) [61,62]. Alternatively, by taking advantage of the strong SOC of phosphorescent organometallic complexes, metal-TADF emitters may achieve high emission quantum yields and fast radiative decay rates, and thereby eventually find practical applications in the OLED industry. Since Yersin and co-workers demonstrated the first TADF emission from a [(pop)Cu(N^N)] complex, in addition to Cu(I) emitters [63,64], second and third row transition metal complexes, such as those of Pd(II), Ag(I), Au(I), and Au(III), have been developed as TADF emitters [65,66]. Among the reported metal-TADF emitters, those based on two-coordinate carbene-metal-amide (CMA) emitters and tetradentate Au(III) complexes have displayed the best performance in terms of efficiency and stability [67,68]. Nonetheless, performances of blue-emitting OLEDs with metal-TADF were relatively inferior when compared to those of green, yellow, and red in terms of efficiency and color purity, until a recent publication presented blue-emitting metal-TADF emitters based on Cu(I) and Au(I) complexes [40,62,65,66,69,70,71,72].
As shown in Figure 8, by substituting CF3 and tBu for the carbazole substitution of the CMA(Au) emitter they reported earlier [67], Romanov, Credgington and colleagues developed a highly efficient blue-emitting metal-TADF emitter Au(I)-1 [71]. Since the HOMO is almost entirely located on the Cz donor of the (AdCAAC)AuCz archetype, the introduction of electron-withdrawing groups to this moiety influences the HOMO more than the LUMO. It widens the HOMO–LUMO gap, shifting it towards blue emission [67]. A high quantum yield of 0.96 with an emission maximum located at 495 nm was achieved in Au(I)-1 in toluene solution. Considering that the high polarity of CMA compounds makes their emission energies sensitive to their molecular environment, allowing the “tuning” of EL by suitable host media from green to sky blue, the emission maximum of Au(I)-1 can be tuned from 484 nm in neat film to 479 and 464 nm when doped in o-CBP and DPEPO hosts, respectively. The EL performance of Au(I)-1 was characterized by an OLED with a device structure of ITO/TAPC (40 nm)/o-CBP (5 nm)/EML (30 nm)/TSPO1 (40 nm)/LiF (1 nm)/Al (100 nm). In addition, since CMAs do not suffer from strong concentration quenching in the solid state, attributed to the lack of close metal–metal contacts, host-free OLEDs with neat Au(I)-1 as the EML have also been fabricated. Similar to the trend in the PL spectra, the device with the high-polarity DPEPO host shows the best color purity, with a maximum located at 450 nm, leading to CIE coordinates of (0.17, 0.17) along with a high EQE of up to 20.9%. Encouragingly, a high performance with an EQE of up to 17.3% and CIE coordinates of (0.18, 0.27) were also achieved in the host-free Au(I)-1-based OLED.
As depicted in Figure 9, by modifying a5, whose HOMO and LUMO are localized on the aryl donor triphenylamine (TPA) and mainly on the C^N^C ligand [73], respectively, two blue-emitting Au(III)-TADF complexes were developed by lowering the HOMO energy or raising the LUMO energy. Specifically, Au(III)-1 was prepared by replacing the OEt group in complex a5 with a more electron-donating dimethyl-amine group to raise the LUMO energy, while Au(III)-2 was prepared by further substituting the phenyl rings of the TPA with electron-withdrawing F substituents to lower the HOMO energy. By means of electrochemistry, the LUMO energy of Au(III)-1 and Au(III)-2 was measured to be 0.18 eV higher than that of complex a5, attributable to the replacement of the OEt group by the more electron-donating NMe2 group. For Au(III)-2, the fluorine substituents on the TPA ligand slightly lowered the HOMO energy by about 0.03 eV from that of Au(III)-1. Thus, PL maxima at 484 and 470 nm were achieved in Au(III)-1 and Au(III)-2, respectively, in PMMA thin film. Despite its less blue emission, the high PLQY of 0.82 for Au(III)-1 in PMMA thin film is much higher than that (0.34) of Au(III)-2. By their application in solution-processed OLEDs, EL properties of both TADF Au(III) complexes were investigated. The device structure was ITO/PEDOT:PSS/OTPD (4 nm)/PYD2: Au-emitter (60 nm)/DPEPO (10 nm)/TPBi (40 nm)/LiF (1.2 nm)/Al (100 nm). The Au(III)-1- and Au(III)-2-based devices show EL maxima at 473 and 465 nm, respectively, with a FWHM of 64–67 nm and CIE coordinates of (0.16, 0.25) for the former and (0.16, 0.23) for the latter. Due to its higher PLQY, the EQEmax of 15.25% shown by the Au(III)-1-based device is much higher than that of the Au(III)-2-based device (6.76%). By employing an MR-TADF blue emitter, v-DABNA, as the terminal emitter, a hyper-fluorescent solution-processed OLED was fabricated using the same device structure for Au(III)-1 (Figure 10). Efficient ET from Au(III)-1 to v-DABNA, with an estimated Förster distance of 2.83 nm and an ET rate of 2.35 × 107 s−1, was determined by photophysical investigations. The blue hyper-fluorescent OLED achieved a high EQEmax of 16.6%, a narrowband emission with an FWHM of 23 nm, and high color purity with CIE coordinates of (0.14, 0.18). When compared with that (30%) of the Au(III)-1 only device, the efficiency roll-off (13%) of the hyper-fluorescent device was significantly improved, attributable to the effective ET process in the latter that could lower the population of triplet excitons in Au(III)-1 at high luminance [40].
Since the reports of CMA emitters based on bulky NHC ligands, with emission quantum yields up to unity, a kr of up to 1 × 106 s−1, and EL efficiency up to 20%, Cu(I) complexes have been expected to replace expensive phosphorescent complexes in OLEDs as emitters [64,67,70,74]. Nonetheless, the operational stability of Cu(I)-based OLEDs, especially that of blue-emitting ones, has barely been reported in the literature [75]. With the use of heterocyclic π-annulated NHC ligands with appreciable π-acidity, we report a panel of thermally and air-stable CMA(Cu)-TADF emitters based on bulky pyrazine- and pyridine-fused NHC ligands [62]. As a blue-emitter, Cu-5 (Figure 11) shows PL emission with a maximum at 470 nm, a PLQY of 0.52, and a kr of 1.1 × 106 s−1 in thin films. EL performance of Cu-5 was investigated in a vacuum-deposited OLED with a device structure of ITO/HAT-CN (5 nm)/TAPC (50 nm)/TCTA (10 nm)/Cu-5: TCTA: DPEPO (20 nm)/DPEPO (10 nm)/TPBi (40 nm)/LiF (1 nm)/Al (100 nm). When the doping concentration was 4 wt%, the Cu5-based device displayed a blue emission with an EL maximum located at 474 nm and CIE coordinates of (0.14, 0.22), as well as a high EQEmax of 23.6%, which is among the best blue OLEDs reported in the literature [76,77]. Notably, the operational lifetime (LT90) of OLEDs based on Cu5 was 0.4 h at an L0 of 7600 cd m−2, corresponding to 11.6 h at an L0 of 1000 cd m−2 using the acceleration coefficient (n), obtained as shown in Figure 11. By using v-DABNA as a terminal emitter, hyper-fluorescent OLEDs with Cu5 as a sensitizer showed an improved color purity with CIE coordinates of (0.15, 0.20) and a longer LT90 of 12.2 h at an L0 of 1000 cd m−2.

5. Outlook and Perspective

Bimolecular annihilations, such as TTA and TPA, caused by the long decay time of TADF emitters are the main reason for the efficiency roll-off and unsatisfactory stability of TADF-OLEDs, particularly blue ones. To surmount such disadvantages, specific molecular structures with extremely small or even negative ΔEST have been designed to decrease the decay time [78,79,80,81,82].
Yersin and co-workers reported a strategy to design a new type of emitter based on a new exciton harvesting mechanism by significantly reducing ΔEST, rigidifying molecular structure, maximizing the Franck–Condon factors, shifting the charge transfer states, and providing energetically nearby lying states [78]. As shown in Figure 12, in the designed compound DSH, both D and A moieties are rigidly linked by two bridges at relatively large separations [81]. Thus, the low-lying CT states, 1CT(DA) and 3CT(DA), have a minimal splitting of ≈10 cm−1 (1.2 meV), while the fluorene-substituted bridge can further stabilize the molecular structure. At the same time, one more low-lying fluorene-localized 3LE(F) state is found in the fluorene-substituted bridge. Their design mitigated the forbidden ISC processes, enabling faster ISC and RISC rates for DSH than those of conventional TADF emitters. Thus, the long decay times of conventional TADF emitters can be avoided in DSH. By using DSH as an emitting dopant in OLEDs, with a structure of ITO/HAT-CN (5 nm)/TAPC (40 nm)/CCP (10 nm)/DSH: PPF (10 nm)/PPF (10 nm)/Tm3PyBPZ (40 nm)/LiF (1.2 nm)/Al (100 nm), an EQEmax of 18.7% with an EL maximum at 486 nm was achieved. Although the device performance and emission color are strongly dependent on the dielectric constant of the host material, which could cause difficulty in optimizing the device structure, the specific molecular design of DSH provides a guideline for next-generation TADF materials.
Despite the general agreement that ΔEST must be positive, some nitrogen-substituted phenalene analogues have the potential for a negative ΔEST, which can be attributed to the double-excitation configurations. Two electrons of occupied orbitals are promoted to virtual orbitals in double-excitation configurations [83,84,85,86]. Since the accessible double-excitation configurations in T1 are restricted (Pauli exclusion principle), an effective admixture of such configurations stabilizes S1, leading to a possibly negative ΔEST when the stabilization overcomes the exchange energy (Figure 13a). With the motivation of discovering molecules with negative ΔEST, from which delayed fluorescence from inverted singlet and triplet (DFIST) excited states can be observed, Miyajima and colleagues computationally screened 34,596 molecules, which were heptazine analogues with 186 different substituents [82]. Among the screening results, HzTFEX2 and HzPipX2 (Figure 13b), were compared. Theory (EOM-CCSD) calculations revealed that HzTFEX2 has a negative ΔEST of −12 meV, while HzPipX2 has a small positive ΔEST of 10 meV. Measurements of the photophysics and transient PL decay further confirmed the negative ΔEST of HzTFEX2, which exhibits DFIST. When it was used as the emitting dopant in the device with a structure of Glass/ITO (130 nm)/PEDOT: PSS (30 nm)/MoO3 (5 nm)/BST (3 nm)/DBFSiDBF (10 nm)/PPF: 10 wt% HzTFEX2 (15 nm)/PPF (10 nm)/Alq3 (40 nm)/Liq (1 nm)/Al (80 nm), intense blue EL originating from HzTFEX2 was observed with an EQEmax of 17% and EL maxima at 450 nm and 479 nm with CIE coordinates of (0.17, 0.24). The remarkable efficiency roll-off suggests that great efforts are still needed to fully exploit the benefits of negative ΔEST in EL devices.

Author Contributions

Conceptualization, Y.G. and S.W.; writing—original draft, Y.G. and S.W.; writing—review and editing, Y.G., G.S. and G.C.; supervision, G.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Guangdong Major Project of Basic and Applied Basic Research (2019B030302009), Hong Kong Quantum AI Lab Limited, Innovation and Technology Fund (PRP/071/19FX), Shunde Science and Technology Bureau (2030218000158), and Science Technology and Innovation Commission of Shenzhen Municipality (JCYJ20180508162429786 and JCYJ20200109150414471).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Full name, molecular structure and function of auxiliary materials used in OLEDs in this review.
Table A1. Full name, molecular structure and function of auxiliary materials used in OLEDs in this review.
MaterialFull NameStructureFunction
HAT-CNDipyrazino[2,3-f:2′,3′-h]quinoxaline-2,3,6,7,10,11-hexacarbonitrileMicromachines 13 02150 i001HIL
TAPCDi-[4-(N,N-ditolyl-amino)-phenyl]cyclohexanMicromachines 13 02150 i002HTL
TCTA4,4′,4″-Tris(carbazol-9-yl)-triphenylamineMicromachines 13 02150 i003HTL
mCBP3,3-Di(9H-carbazol-9-yl)biphenylMicromachines 13 02150 i004HTL
ANT-BIZ1-(4-(10-([1,1′-biphenyl]-4-yl)anthracen-9-yl)phenyl)-2-ethyl-1H-benzo[d]imidazoleMicromachines 13 02150 i005ETL
mCP1,3-Di(9H-carbazol-9-yl) benzene, N,N′-Dicarbazolyl-3,5-benzeneMicromachines 13 02150 i006HTL
BCFNN-([1,1′-biphenyl]-4-yl)-9,9-dimethyl-N-(4-(9-phenyl-9H-carbazol-3-yl)phenyl)-9H-fluoren-2-amineMicromachines 13 02150 i007HTL
DBFTRz2,8-Bis(4,6-diphenyl-1,3,5-triazin-2-yl)dibenzo[b,d]furanMicromachines 13 02150 i008HBL
ZADN2-(4-(9,10-Di(naphthalen-2-yl)anthracen-2-yl)phenyl)-1-phenyl-1H-benzo[d]imidazoleMicromachines 13 02150 i009ETL
TSPO1Diphenyl[4-(triphenylsilyl)phenyl]phosphine oxideMicromachines 13 02150 i010HBL
TmPyPB1,3,5-tri[(3-pyridyl)-phen-3-yl]benzeneMicromachines 13 02150 i011ETL
PPF2,8-Bis(diphenylphosphoryl)dibenzo[b,d]furanMicromachines 13 02150 i012HTL
BPhen4,7-diphenyl-1,10-phenanthrolineMicromachines 13 02150 i013HBL
OTPD[N,N′-bis(4-(6-((3-ethyloxetan-3-yl)methoxy))-hexylphenyl)-N,N′-diphenyl-4,4′-diamine]Micromachines 13 02150 i014HIL
PYD22,6-Bis(9H-carbazol-9-yl)pyridine, 9,9-(2,6-pyridinediyl)bis-9H-carbazoleMicromachines 13 02150 i015Host
TPBi2,2′,2″-(1,3,5-benzinetriyl)-tris(1-phenyl-1-H-benzimidazole)Micromachines 13 02150 i016ETL
CCP9-Phenyl-9H-3,9′-bicarbazoleMicromachines 13 02150 i017HTL
Alq3Tris(8-hydroxy-quinolinato)aluminiumMicromachines 13 02150 i018ETL

Appendix B

Table A2. Key performances of blue OLEDs described in this review and recent results.
Table A2. Key performances of blue OLEDs described in this review and recent results.
TypeEmitterEQEmax a
[%]
EQE1000 b
[%]
Lifetime
[h]
λmax c
[nm]
FWHM
[nm]
CIE Coordinates
[x, y]
Ref.
MR-TADFν-DABNA34.426.0 46918(0.12, 0.11)[35]
BN131.29.3-45728(0.14, 0.08)[45]
BN233.215.5-46723(0.13, 0.11)[45]
BN337.626.227.2 d45823(0.14, 0.08)[45]
tCBNDADPO30.816.2-47228(0.14, 0.22)[44]
t-DABNA25.2--46527(0.126, 0.102)[49]
t-DABNA-dtB30.128.8502 e47122(0.117, 0.112)[49]
TSCT-TADFP-Ac95-TRZ0512.111.5-479-(0.176, 0.269)[59]
dCz-Xo-TRZ27.823.9-47719(0.17, 0.29)[39]
AC-BO19.3--45671(0.147, 0.122)[60]
QAC-BO15.8--44863(0.145, 0.076)[60]
Cz-BO5.5--41243(0.163, 0.034)[60]
Metal-TADFAu(I)-120.9--450-(0.17, 0.17)[71]
Au(III)-115.259.98-47364(0.16, 0.25)[72]
Au(III)-26.762.51-46567(0.16, 0.23)[72]
Cu523.612.911.6 f47460(0.14, 0.22)[62]
a Maximum EQE; b EQE at 1000 cd m−2; c Emission maximum; d LT50 at 100 cd−2; e LT95 at 1000 cd m−2; f LT90 at 1000 cd m−2.

References

  1. Tang, C.W.; Vanslyke, S.A. Organic electroluminescent diodes. Appl. Phys. Lett. 1987, 51, 913. [Google Scholar] [CrossRef]
  2. Friend, R.H.; Gymer, R.W.; Holmes, A.B.; Burroughes, J.H.; Marks, R.N.; Taliani, C.; Bradley, D.D.C.; Dos Santos, D.A.; Brédas, J.L.; Lögdlund, M.; et al. Electroluminescence in conjugated polymers. Nature 1999, 397, 121–128. [Google Scholar] [CrossRef]
  3. Park, I.S.; Komiyama, H.; Yasuda, T. Pyrimidine-based twisted donor–acceptor delayed fluorescence molecules: A new universal platform for highly efficient blue electroluminescence. Chem. Sci. 2016, 8, 953–960. [Google Scholar] [CrossRef] [Green Version]
  4. Lee, S.Y.; Adachi, C.; Yasuda, T. High-Efficiency Blue Organic Light-Emitting Diodes Based on Thermally Activated Delayed Fluorescence from Phenoxaphosphine and Phenoxathiin Derivatives. Adv. Mater. 2016, 28, 4626–4631. [Google Scholar] [CrossRef]
  5. Cha, J.-R.; Lee, C.W.; Lee, J.Y.; Gong, M.-S. Design of ortho-linkage carbazole-triazine structure for high-efficiency blue thermally activated delayed fluorescent emitters. Dye. Pigment. 2016, 134, 562–568. [Google Scholar] [CrossRef]
  6. Sun, K.; Sun, Y.; Jiang, W.; Huang, S.; Tian, W.; Sun, Y. Highly efficient and color tunable thermally activated delayed fluorescent emitters and their applications for the solution-processed OLEDs. Dye. Pigment. 2017, 139, 326–333. [Google Scholar] [CrossRef]
  7. Braveenth, R.; Chai, K.Y. Triazine-Acceptor-Based Green Thermally Activated Delayed Fluorescence Materials for Organic Light-Emitting Diodes. Materials 2019, 12, 2646. [Google Scholar] [CrossRef] [Green Version]
  8. Chiu, T.-L.; Chuang, Y.-T. Spectral observations of hole injection with transition metal oxides for an efficient organic light-emitting diode. J. Phys. D Appl. Phys. 2015, 48, 075101. [Google Scholar] [CrossRef]
  9. 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] [PubMed] [Green Version]
  10. Reineke, S.; Thomschke, M.; Lussem, B.; Leo, K. White organic light-emitting diodes: Status and perspective. Rev. Mod. Phys. 2013, 85, 1245–1293. [Google Scholar] [CrossRef]
  11. Braveenth, R.; Bae, I.-J.; Wang, Y.; Kim, S.H.; Kim, M.; Chai, K.Y. Acridine-Triphenylamine Based Hole-Transporting and Hole-Injecting Material for Highly Efficient Phosphorescent-Based Organic Light Emitting Diodes. Appl. Sci. 2018, 8, 1168. [Google Scholar] [CrossRef] [Green Version]
  12. Huh, D.H.; Kim, G.W.; Kulshreshtha, C.; Kwon, J.H. High hole mobility hole transport material for organic light-emitting devices. Synth. Met. 2013, 180, 79–84. [Google Scholar] [CrossRef]
  13. Braveenth, R.; Bae, H.W.; Ko, I.J.; Qiong, W.; Nguyen, Q.P.B.; Jayashantha, P.G.S.; Kwon, J.H.; Chai, K.Y. Thermally stable efficient hole transporting materials based on carbazole and triphenylamine core for red phosphorescent OLEDs. Org. Electron. 2017, 51, 463–470. [Google Scholar] [CrossRef]
  14. Zheng, Z.; Dong, Q.; Gou, L.; Su, J.-H.; Huang, J. Novel hole transport materials based on N,N′-disubstituted-dihydrophenazine derivatives for electroluminescent diodes. J. Mater. Chem. C 2014, 2, 9858–9865. [Google Scholar] [CrossRef]
  15. Wei, Q.; Fei, N.; Islam, A.; Lei, T.; Hong, L.; Peng, R.; Fan, X.; Chen, L.; Gao, P.; Ge, Z. Small-Molecule Emitters with High Quantum Efficiency: Mechanisms, Structures, and Applications in OLED Devices. Adv. Opt. Mater. 2018, 6, 1800512. [Google Scholar] [CrossRef]
  16. Baldo, M.A.; O’Brien, D.F.; Thompson, M.E.; Forrest, S.R. Excitonic singlet-triplet ratio in a semiconducting organic thin film. Phys. Rev. B Condens. Matter Mater. Phys. 1999, 60, 14422–14428. [Google Scholar] [CrossRef]
  17. Forrest, S.; Bradley, D.; Thompson, M. Measuring the Efficiency of Organic Light-Emitting Devices. Adv. Mater. 2003, 15, 1043–1048. [Google Scholar] [CrossRef]
  18. Reineke, S.; Lindner, F.; Schwartz, G.; Seidler, N.; Walzer, K.; Lüssem, B.; Leo, K. White organic light-emitting diodes with fluorescent tube efficiency. Nature 2009, 459, 234–238. [Google Scholar] [CrossRef] [PubMed]
  19. 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]
  20. Baldo, M.A.; Adachi, C.; Forrest, S.R. Transient analysis of organic electrophosphorescence. II. Transient analysis of triplet-triplet annihilation. Phys. Rev. B 2000, 62, 10967. [Google Scholar] [CrossRef]
  21. Ganzorig, C.; Fujihira, M. A possible mechanism for enhanced electrofluorescence emission through triplet–triplet annihilation in organic electroluminescent devices. Appl. Phys. Lett. 2002, 81, 3137–3139. [Google Scholar] [CrossRef]
  22. Leitl, M.J.; Zink, D.M.; Schinabeck, A.; Baumann, T.; Volz, D.; Yersin, H. Copper(I) Complexes for Thermally Activated Delayed Fluorescence: From Photophysical to Device Properties. Top. Curr. Chem. 2016, 374, 25. [Google Scholar] [CrossRef] [PubMed]
  23. 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]
  24. 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]
  25. Lin, C.-C.; Huang, M.-J.; Chiu, M.-J.; Huang, M.-P.; Chang, C.-C.; Liao, C.-Y.; Chiang, K.-M.; Shiau, Y.-J.; Chou, T.-Y.; Chu, L.-K.; et al. Molecular Design of Highly Efficient Thermally Activated Delayed Fluorescence Hosts for Blue Phosphorescent and Fluorescent Organic Light-Emitting Diodes. Chem. Mater. 2017, 29, 1527–1537. [Google Scholar] [CrossRef]
  26. Huang, T.; Jiang, W.; Duan, L. Recent progress in solution processable TADF materials for organic light-emitting diodes. J. Mater. Chem. C 2018, 6, 5577–5596. [Google Scholar] [CrossRef]
  27. Komatsu, R.; Sasabe, H.; Nakao, K.; Hayasaka, Y.; Ohsawa, T.; Kido, J. Unlocking the Potential of Pyrimidine Conjugate Emitters to Realize High-Performance Organic Light-Emitting Devices. Adv. Opt. Mater. 2017, 5, 1600675. [Google Scholar] [CrossRef]
  28. Liang, X.; Tu, Z.; Zheng, Y. Thermally Activated Delayed Fluorescence Materials: Towards Realization of High Efficiency through Strategic Small Molecular Design. Chem. Eur. J. 2019, 25, 5623–5642. [Google Scholar] [CrossRef]
  29. Kitamoto, Y.; Namikawa, T.; Suzuki, T.; Miyata, Y.; Kita, H.; Sato, T.; Oi, S. Design and synthesis of efficient blue thermally activated delayed fluorescence molecules bearing triarylborane and 10,10-dimethyl-5,10-dihydrophenazasiline moieties. Tetrahedron Lett. 2016, 57, 4914–4917. [Google Scholar] [CrossRef]
  30. 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]
  31. Hatakeyama, T.; Shiren, K.; Nakajima, K.; Nomura, S.; Nakatsuka, S.; Kinoshita, K.; Ni, J.; Ono, Y.; Ikuta, T. Ultrapure Blue Thermally Activated Delayed Fluorescence Molecules: Efficient HOMO-LUMO Separation by the Multiple Resonance Effect. Adv. Mater. 2016, 28, 2777–2781. [Google Scholar] [CrossRef]
  32. Matsui, K.; Oda, S.; Yoshiura, K.; Nakajima, K.; Yasuda, N.; Hatakeyama, T. One-Shot Multiple Borylation toward BN-Doped Nanographenes. J. Am. Chem. Soc. 2018, 140, 1195–1198. [Google Scholar] [CrossRef]
  33. Oda, S.; Sugitani, T.; Tanaka, H.; Tabata, K.; Kawasumi, R.; Hatakeyama, T. Development of Pure Green Thermally Activated Delayed Fluorescence Material by Cyano Substitution. Adv. Mater. 2022, 34, 2201778. [Google Scholar] [CrossRef]
  34. Lee, H.; Karthik, D.; Lampande, R.; Ryu, J.H.; Kwon, J.H. Recent Advancement in Boron-Based Efficient and Pure Blue Thermally Activated Delayed Fluorescence Materials for Organic Light-Emitting Diodes. Front. Chem. 2020, 8, 373. [Google Scholar] [CrossRef]
  35. Kondo, Y.; Yoshiura, K.; Kitera, S.; Nishi, H.; Oda, S.; Gotoh, H.; Sasada, Y.; Yanai, M.; Hatakeyama, T. Narrowband deep-blue organic light-emitting diode featuring an organoboron-based emitter. Nat. Photonics 2019, 13, 678–682. [Google Scholar] [CrossRef]
  36. Chan, C.-Y.; Tanaka, M.; Lee, Y.-T.; Wong, Y.-W.; Nakanotani, H.; Hatakeyama, T.; Adachi, C. Stable pure-blue hyperfluorescence organic light-emitting diodes with high-efficiency and narrow emission. Nat. Photonics 2021, 15, 203–207. [Google Scholar] [CrossRef]
  37. Jeon, S.O.; Lee, K.H.; Kim, J.S.; Ihn, S.-G.; Chung, Y.S.; Kim, J.W.; Lee, H.; Kim, S.; Choi, H.; Lee, J.Y. High-efficiency, long-lifetime deep-blue organic light-emitting diodes. Nat. Photonics 2021, 15, 208–215. [Google Scholar] [CrossRef]
  38. Lo, K.; Tong, G.S.M.; Cheng, G.; Low, K.; Che, C. Dinuclear Pt II Complexes with Strong Blue Phosphorescence for Operationally Stable Organic Light-Emitting Diodes with EQE up to 23% at 1000 cd m−2. Angew. Chem. Int. Ed. 2022, 61, e202115515. [Google Scholar] [CrossRef]
  39. Huang, T.; Wang, Q.; Meng, G.; Duan, L.; Zhang, D. Accelerating Radiative Decay in Blue Through-Space Charge Transfer Emitters by Minimizing the Face-to-Face Donor-Acceptor Distances. Angew. Chem. Int. Ed. 2022, 61, e202200059. [Google Scholar] [CrossRef]
  40. Zhou, D.; Wu, S.; Cheng, G.; Che, C.-M. A gold(iii)–TADF emitter as a sensitizer for high-color-purity and efficient deep-blue solution-processed OLEDs. J. Mater. Chem. C 2022, 10, 4590–4596. [Google Scholar] [CrossRef]
  41. Zhang, Y.; Zhang, D.; Wei, J.; Hong, X.; Lu, Y.; Hu, D.; Li, G.; Liu, Z.; Chen, Y.; Duan, L. Achieving Pure Green Electroluminescence with CIEy of 0.69 and EQE of 28.2% from an Aza-Fused Multi-Resonance Emitter. Angew. Chem. Int. Ed. 2020, 59, 17499–17503. [Google Scholar] [CrossRef]
  42. Yang, M.; Shikita, S.; Min, H.; Park, I.S.; Shibata, H.; Amanokura, N.; Yasuda, T. Wide-Range Color Tuning of Narrowband Emission in Multi-resonance Organoboron Delayed Fluorescence Materials through Rational Imine/Amine Functionalization. Angew. Chem. Int. Ed. 2021, 60, 23142–23147. [Google Scholar] [CrossRef]
  43. Jiang, P.; Miao, J.; Cao, X.; Xia, H.; Pan, K.; Hua, T.; Lv, X.; Huang, Z.; Zou, Y.; Yang, C. Quenching-Resistant Multiresonance TADF Emitter Realizes 40% External Quantum Efficiency in Narrowband Electroluminescence at High Doping Level. Adv. Mater. 2022, 34, 2106954. [Google Scholar] [CrossRef]
  44. Bian, J.; Chen, S.; Qiu, L.; Tian, R.; Man, Y.; Wang, Y.; Chen, S.; Zhang, J.; Duan, C.; Han, C.; et al. Ambipolar Self-Host Functionalization Accelerates Blue Multi-Resonance Thermally Activated Delayed Fluorescence with Internal Quantum Efficiency of 100%. Adv. Mater. 2022, 34, 2110547. [Google Scholar] [CrossRef]
  45. Lv, X.; Miao, J.; Liu, M.; Peng, Q.; Zhong, C.; Hu, Y.; Cao, X.; Wu, H.; Yang, Y.; Zhou, C.; et al. Extending the π-Skeleton of Multi-Resonance TADF Materials towards High-Efficiency Narrowband Deep-Blue Emission. Angew. Chem. Int. Ed. 2022, 61, e202201588. [Google Scholar] [CrossRef]
  46. Pershin, A.; Hall, D.; Lemaur, V.; Sancho-Garcia, J.-C.; Muccioli, L.; Zysman-Colman, E.; Beljonne, D.; Olivier, Y. Highly emissive excitons with reduced exchange energy in thermally activated delayed fluorescent molecules. Nat. Commun. 2019, 10, 597. [Google Scholar] [CrossRef] [Green Version]
  47. Xia, D.; Wang, B.; Chen, B.; Wang, S.; Zhang, B.; Ding, J.; Wang, L.; Jing, X.; Wang, F. Self-Host Blue-Emitting Iridium Dendrimer with Carbazole Dendrons: Nondoped Phosphorescent Organic Light-Emitting Diodes. Angew. Chem. Int. Ed. 2014, 53, 1048–1052. [Google Scholar] [CrossRef]
  48. Xu, Y.; Cheng, Z.; Li, Z.; Liang, B.; Wang, J.; Wei, J.; Zhang, Z.; Wang, Y. Molecular-Structure and Device-Configuration Optimizations toward Highly Efficient Green Electroluminescence with Narrowband Emission and High Color Purity. Adv. Opt. Mater. 2020, 8, 1902142. [Google Scholar] [CrossRef]
  49. Park, J.; Kim, K.J.; Lim, J.; Kim, T.; Lee, J.Y. High Efficiency of over 25% and Long Device Lifetime of over 500 h at 1000 nit in Blue Fluorescent Organic Light-Emitting Diodes. Adv. Mater. 2022, 34, 2108581. [Google Scholar] [CrossRef]
  50. Han, S.H.; Jeong, J.H.; Yoo, J.W.; Lee, J.Y. Ideal blue thermally activated delayed fluorescence emission assisted by a thermally activated delayed fluorescence assistant dopant through a fast reverse intersystem crossing mediated cascade energy transfer process. J. Mater. Chem. C 2019, 7, 3082–3089. [Google Scholar] [CrossRef]
  51. Xue, Q.; Xie, G. Thermally Activated Delayed Fluorescence beyond Through-Bond Charge Transfer for High-Performance OLEDs. Adv. Opt. Mater. 2021, 9, 2002204. [Google Scholar] [CrossRef]
  52. Chen, X.-K.; Bakr, B.W.; Auffray, M.; Tsuchiya, Y.; Sherrill, C.D.; Adachi, C.; Bredas, J.-L. Intramolecular Noncovalent Interactions Facilitate Thermally Activated Delayed Fluorescence (TADF). J. Phys. Chem. Lett. 2019, 10, 3260–3268. [Google Scholar] [CrossRef]
  53. Wu, C.; Liu, W.; Li, K.; Cheng, G.; Xiong, J.; Teng, T.; Che, C.; Yang, C. Face-to-Face Orientation of Quasiplanar Donor and Acceptor Enables Highly Efficient Intramolecular Exciplex Fluorescence. Angew. Chem. Int. Ed. 2021, 60, 3994–3998. [Google Scholar] [CrossRef]
  54. Tsujimoto, H.; Ha, D.-G.; Markopoulos, G.; Chae, H.S.; Baldo, M.A.; Swager, T.M. Thermally Activated Delayed Fluorescence and Aggregation Induced Emission with Through-Space Charge Transfer. J. Am. Chem. Soc. 2017, 139, 4894–4900. [Google Scholar] [CrossRef]
  55. 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]
  56. Peng, C.; Yang, S.; Li, H.; Xie, G.; Cui, L.; Zou, S.; Poriel, C.; Jiang, Z.; Liao, L. Highly Efficient Thermally Activated Delayed Fluorescence via an Unconjugated Donor-Acceptor System Realizing EQE of Over 30%. Adv. Mater. 2020, 32, e2003885. [Google Scholar] [CrossRef]
  57. Tang, X.; Cui, L.-S.; Li, H.-C.; Gillett, A.J.; Auras, F.; Qu, Y.-K.; Zhong, C.; Jones, S.T.E.; Jiang, Z.-Q.; Friend, R.H.; et al. Highly efficient luminescence from space-confined charge-transfer emitters. Nat. Mater. 2020, 19, 1332–1338. [Google Scholar] [CrossRef]
  58. Wang, X.; Yang, S.; Tian, Q.; Zhong, C.; Qu, Y.; Yu, Y.; Jiang, Z.; Liao, L. Multi-Layer π-Stacked Molecules as Efficient Thermally Activated Delayed Fluorescence Emitters. Angew. Chem. Int. Ed. 2020, 60, 5213–5219. [Google Scholar] [CrossRef]
  59. Shao, S.; Hu, J.; Wang, X.; Wang, L.; Jing, X.; Wang, F. Blue Thermally Activated Delayed Fluorescence Polymers with Nonconjugated Backbone and Through-Space Charge Transfer Effect. J. Am. Chem. Soc. 2017, 139, 17739–17742. [Google Scholar] [CrossRef]
  60. Zhao, Z.; Zeng, C.; Peng, X.; Liu, Y.; Zhao, H.; Hua, L.; Su, S.-J.; Yan, S.; Ren, Z. Tuning Intramolecular Stacking of Rigid Heteroaromatic Compounds for High-Efficiency Deep-blue Through-Space Charge-Transfer Emission. Angew. Chem. Int. Ed. 2022, 61, e202210864. [Google Scholar] [CrossRef]
  61. Yang, Z.; Mao, Z.; Xie, Z.; Zhang, Y.; Liu, S.; Zhao, J.; Xu, J.; Chi, Z.; Aldred, M.P. Recent advances in organic thermally activated delayed fluorescence materials. Chem. Soc. Rev. 2017, 46, 915–1016. [Google Scholar] [CrossRef] [PubMed]
  62. Tang, R.; Xu, S.; Lam, T.; Cheng, G.; Du, L.; Wan, Q.; Yang, J.; Hung, F.; Low, K.; Phillips, D.L.; et al. Highly Robust Cu I-TADF Emitters for Vacuum-Deposited OLEDs with Luminance up to 222 200 cd m−2 and Device Lifetimes (LT90) up to 1300 h at an Initial Luminance of 1000 cd m−2. Angew. Chem. Int. Ed. 2022, 61, e202203982. [Google Scholar] [CrossRef]
  63. Czerwieniec, R.; Yu, J.; Yersin, H. Blue-Light Emission of Cu(I) Complexes and Singlet Harvesting. Inorg. Chem. 2011, 50, 8293–8301. [Google Scholar] [CrossRef] [PubMed]
  64. Czerwieniec, R.; Leitl, M.J.; Homeier, H.H.; Yersin, H. Cu(I) complexes—Thermally activated delayed fluorescence. Photophysical approach and material design. Coord. Chem. Rev. 2016, 325, 2–28. [Google Scholar] [CrossRef]
  65. Yersin, H.; Czerwieniec, R.; Shafikov, M.Z.; Suleymanova, A.F. TADF Material Design: Photophysical Background and Case Studies Focusing on CuI and AgI Complexes. ChemPhysChem 2017, 18, 3508–3535. [Google Scholar] [CrossRef]
  66. To, W.-P.; Cheng, G.; Tong, G.S.M.; Zhou, D.; Che, C.-M. Recent Advances in Metal-TADF Emitters and Their Application in Organic Light-Emitting Diodes. Front. Chem. 2020, 8, 653. [Google Scholar] [CrossRef]
  67. Di, D.; Romanov, A.S.; Yang, L.; Richter, J.M.; Rivett, J.P.H.; Jones, S.; Thomas, T.H.; Jalebi, M.A.; Friend, R.H.; Linnolahti, M.; et al. High-performance light-emitting diodes based on carbene-metal-amides. Science 2017, 356, 159–163. [Google Scholar] [CrossRef] [Green Version]
  68. Zhou, D.; To, W.; Tong, G.S.M.; Cheng, G.; Du, L.; Phillips, D.L.; Che, C. Tetradentate Gold(III) Complexes as Thermally Activated Delayed Fluorescence (TADF) Emitters: Microwave-Assisted Synthesis and High-Performance OLEDs with Long Operational Lifetime. Angew. Chem. Int. Ed. 2020, 59, 6375–6382. [Google Scholar] [CrossRef]
  69. Malmberg, R.; Venkatesan, K. Recent Advances in the Development of Blue and Deep-Blue Emitting Gold(I) and Gold(III) Molecular Systems. Eur. J. Inorg. Chem. 2021, 2021, 4890–4902. [Google Scholar] [CrossRef]
  70. Shi, S.; Jung, M.C.; Coburn, C.; Tadle, A.; Sylvinson, M.R.D.; Djurovich, P.I.; Forrest, S.R.; Thompson, M.E. Highly Efficient Photo- and Electroluminescence from Two-Coordinate Cu(I) Complexes Featuring Nonconventional N-Heterocyclic Carbenes. J. Am. Chem. Soc. 2019, 141, 3576–3588. [Google Scholar] [CrossRef]
  71. Conaghan, P.J.; Matthews, C.S.B.; Chotard, F.; Jones, S.T.E.; Greenham, N.C.; Bochmann, M.; Credgington, D.; Romanov, A.S. Highly efficient blue organic light-emitting diodes based on carbene-metal-amides. Nat. Commun. 2020, 11, 1758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Zhou, D.; Cheng, G.; Tong, G.S.M.; Che, C. High Efficiency Sky-Blue Gold(III)-TADF Emitters. Chem. Eur. J. 2020, 26, 15718–15726. [Google Scholar] [CrossRef]
  73. To, W.-P.; Zhou, D.; Tong, G.S.M.; Cheng, G.; Yang, C.; Che, C.-M. Highly Luminescent Pincer Gold(III) Aryl Emitters: Thermally Activated Delayed Fluorescence and Solution-Processed OLEDs. Angew. Chem. Int. Ed. 2017, 56, 14036–14041. [Google Scholar] [CrossRef] [PubMed]
  74. Ying, A.; Huang, Y.-H.; Lu, C.-H.; Chen, Z.; Lee, W.-K.; Zeng, X.; Chen, T.; Cao, X.; Wu, C.-C.; Gong, S.; et al. High-Efficiency Red Electroluminescence Based on a Carbene-Cu(I)-Acridine Complex. ACS Appl. Mater. Interfaces 2021, 13, 13478–13486. [Google Scholar] [CrossRef]
  75. Liu, Z.; Qiu, J.; Wei, F.; Wang, J.; Liu, X.; Helander, M.G.; Rodney, S.; Wang, Z.; Bian, Z.; Lu, Z.; et al. Simple and High Efficiency Phosphorescence Organic Light-Emitting Diodes with Codeposited Copper(I) Emitter. Chem. Mater. 2014, 26, 2368–2373. [Google Scholar] [CrossRef]
  76. Hamze, R.; Peltier, J.L.; Sylvinson, D.; Jung, M.; Cardenas, J.; Haiges, R.; Soleilhavoup, M.; Jazzar, R.; Djurovich, P.I.; Bertrand, G.; et al. Eliminating nonradiative decay in Cu(I) emitters: >99% quantum efficiency and microsecond lifetime. Science 2019, 363, 601–606. [Google Scholar] [CrossRef]
  77. Xie, M.; Han, C.; Liang, Q.; Zhang, J.; Xie, G.; Xu, H. Highly efficient sky blue electroluminescence from ligand-activated copper iodide clusters: Overcoming the limitations of cluster light-emitting diodes. Sci. Adv. 2019, 5, eaav9857. [Google Scholar] [CrossRef] [Green Version]
  78. Murawski, C.; Leo, K.; Gather, M.C. Efficiency Roll-Off in Organic Light-Emitting Diodes. Adv. Mater. 2013, 25, 6801–6827. [Google Scholar] [CrossRef] [Green Version]
  79. Schmidbauer, S.; Hohenleutner, A.; König, B. Chemical Degradation in Organic Light-Emitting Devices: Mechanisms and Implications for the Design of New Materials. Adv. Mater. 2013, 25, 2114–2129. [Google Scholar] [CrossRef]
  80. Yersin, H.; Mataranga-Popa, L.; Czerwieniec, R.; Dovbii, Y. Design of a New Mechanism beyond Thermally Activated Delayed Fluorescence toward Fourth Generation Organic Light Emitting Diodes. Chem. Mater. 2019, 31, 6110–6116. [Google Scholar] [CrossRef]
  81. Yersin, H.; Czerwieniec, R.; Mataranga-Popa, L.; Mewes, J.; Cheng, G.; Che, C.; Saigo, M.; Kimura, S.; Miyata, K.; Onda, K. Eliminating the Reverse ISC Bottleneck of TADF Through Excited State Engineering and Environment-Tuning toward State Resonance Leading to Mono-Exponential Sub-µs Decay. High OLED External Quantum Efficiency Confirms Efficient Exciton Harvesting. Adv. Funct. Mater. 2022, 32, 2201772. [Google Scholar] [CrossRef]
  82. Aizawa, N.; Pu, Y.-J.; Harabuchi, Y.; Nihonyanagi, A.; Ibuka, R.; Inuzuka, H.; Dhara, B.; Koyama, Y.; Nakayama, K.-I.; Maeda, S.; et al. Delayed fluorescence from inverted singlet and triplet excited states. Nature 2022, 609, 502–506. [Google Scholar] [CrossRef]
  83. Audebert, P.; Kroke, E.; Posern, C.; Lee, S.-H. State of the Art in the Preparation and Properties of Molecular Monomeric s-Heptazines: Syntheses, Characteristics, and Functional Applications. Chem. Rev. 2021, 121, 2515–2544. [Google Scholar] [CrossRef] [PubMed]
  84. de Silva, P. Inverted Singlet-Triplet Gaps and Their Relevance to Thermally Activated Delayed Fluorescence. J. Phys. Chem. Lett. 2019, 10, 5674–5679. [Google Scholar] [CrossRef] [PubMed]
  85. Sanz-Rodrigo, J.; Ricci, G.; Olivier, Y.; Sancho-García, J.C. Negative Singlet-Triplet Excitation Energy Gap in Triangle-Shaped Molecular Emitters for Efficient Triplet Harvesting. J. Phys. Chem. A 2021, 125, 513–522. [Google Scholar] [CrossRef] [PubMed]
  86. Pollice, R.; Friederich, P.; Lavigne, C.; Gomes, G.D.P.; Aspuru-Guzik, A. Organic molecules with inverted gaps between first excited singlet and triplet states and appreciable fluorescence rates. Matter 2021, 4, 1654–1682. [Google Scholar] [CrossRef]
Figure 1. Molecular structures of BN1–BN3. (a) Device architecture for MR-TADF OLEDs with energy level alignment of the relevant material, (b) EL spectra, with CIE coordinates shown in the inset, (c) J–V–L characteristics, and (d) EQE–luminance curves of MR-TADF OLEDs. Reproduced with permission from Xialei Lv, Jingsheng Miao, Meihui Liu et al., “Extending the π-Skeleton of Multi-Resonance TADF Materials towards High-Efficiency Narrowband Deep-Blue Emission”; published by John Wiley and Sons, 2022. [45].
Figure 1. Molecular structures of BN1–BN3. (a) Device architecture for MR-TADF OLEDs with energy level alignment of the relevant material, (b) EL spectra, with CIE coordinates shown in the inset, (c) J–V–L characteristics, and (d) EQE–luminance curves of MR-TADF OLEDs. Reproduced with permission from Xialei Lv, Jingsheng Miao, Meihui Liu et al., “Extending the π-Skeleton of Multi-Resonance TADF Materials towards High-Efficiency Narrowband Deep-Blue Emission”; published by John Wiley and Sons, 2022. [45].
Micromachines 13 02150 g001
Figure 2. Host–guest integration strategy for constructing tCBNDADPO to combine high color purity and efficiency. (a) Energy level diagram of devices and chemical structures of host and carrier transporting materials; (b) CIE coordinates of PL and EL emissions from DBFDPO:30% tCBNDADPO and device photo at 5 V; (c) concentration dependence of EL spectra and comparison to PL spectrum of 30% doped film; (d) variation of current density (J, hollow symbols)–voltage–luminance (solid symbols) characteristics at different x%; (e) dependence of efficiency versus luminance curves on x%. Reproduced with permission from Jinkun Bian, Su Chen, Lili Qiu et al., “Ambipolar Self-Host Functionalization Accelerates Blue Multi-Resonance Thermally Activated Delayed Fluorescence with Internal Quantum Efficiency of 100%”; published by John Wiley and Sons, 2022 [44].
Figure 2. Host–guest integration strategy for constructing tCBNDADPO to combine high color purity and efficiency. (a) Energy level diagram of devices and chemical structures of host and carrier transporting materials; (b) CIE coordinates of PL and EL emissions from DBFDPO:30% tCBNDADPO and device photo at 5 V; (c) concentration dependence of EL spectra and comparison to PL spectrum of 30% doped film; (d) variation of current density (J, hollow symbols)–voltage–luminance (solid symbols) characteristics at different x%; (e) dependence of efficiency versus luminance curves on x%. Reproduced with permission from Jinkun Bian, Su Chen, Lili Qiu et al., “Ambipolar Self-Host Functionalization Accelerates Blue Multi-Resonance Thermally Activated Delayed Fluorescence with Internal Quantum Efficiency of 100%”; published by John Wiley and Sons, 2022 [44].
Micromachines 13 02150 g002
Figure 3. Molecular structures and DFT calculations of t-DABNA and t-DABNA-dtB. Reproduced with permission from Jinho Park, Ki Ju Kim, Junseop Lim et al., “High Efficiency of over 25% and Long Device Lifetime of over 500 h at 1000 nit in Blue Fluorescent Organic Light-Emitting Diodes”; published by John Wiley and Sons, 2022 [49].
Figure 3. Molecular structures and DFT calculations of t-DABNA and t-DABNA-dtB. Reproduced with permission from Jinho Park, Ki Ju Kim, Junseop Lim et al., “High Efficiency of over 25% and Long Device Lifetime of over 500 h at 1000 nit in Blue Fluorescent Organic Light-Emitting Diodes”; published by John Wiley and Sons, 2022 [49].
Micromachines 13 02150 g003
Figure 4. (a) J–V–L and (b) EQE-L characteristics, (c) EL spectra and CIE coordinates (inset), and (d) lifetime data of single stack and tandem devices. (e) Extrapolated plot of the tandem device. Reproduced with permission from Jinho Park, Ki Ju Kim, Junseop Lim et al., “High Efficiency of over 25% and Long Device Lifetime of over 500 h at 1000 nit in Blue Fluorescent Organic Light-Emitting Diodes”; published by John Wiley and Sons, 2022 [49].
Figure 4. (a) J–V–L and (b) EQE-L characteristics, (c) EL spectra and CIE coordinates (inset), and (d) lifetime data of single stack and tandem devices. (e) Extrapolated plot of the tandem device. Reproduced with permission from Jinho Park, Ki Ju Kim, Junseop Lim et al., “High Efficiency of over 25% and Long Device Lifetime of over 500 h at 1000 nit in Blue Fluorescent Organic Light-Emitting Diodes”; published by John Wiley and Sons, 2022 [49].
Micromachines 13 02150 g004
Figure 5. Molecular structures of (a) TSCT-based polymers and (b) the corresponding control polymers. Reprinted with permission from Shao, S.; Hu, J.; Wang, X.; Wang, L.; Jing, X.; Wang, F. “Blue Thermally Activated Delayed Fluorescence Polymers with Nonconjugated Backbone and Through-Space Charge Transfer Effect” [J]. J. Am. Chem. Soc. 2017, 139, 17739−17742. Copyright 2017 American Chemical Society [59].
Figure 5. Molecular structures of (a) TSCT-based polymers and (b) the corresponding control polymers. Reprinted with permission from Shao, S.; Hu, J.; Wang, X.; Wang, L.; Jing, X.; Wang, F. “Blue Thermally Activated Delayed Fluorescence Polymers with Nonconjugated Backbone and Through-Space Charge Transfer Effect” [J]. J. Am. Chem. Soc. 2017, 139, 17739−17742. Copyright 2017 American Chemical Society [59].
Micromachines 13 02150 g005
Figure 6. TSCT-TADF molecules with different types of donor–acceptor alignments, and molecular structures of mCz-Xo-TRZ and dCz-Xo-TRZ with their HOMO/LUMO distribution. Reproduced with permission from Tianyu Huang, Qi Wang, Guoyun Meng et al., “Accelerating Radiative Decay in Blue Through-Space Charge Transfer Emitters by Minimizing the Face-to-Face Donor–Acceptor Distances”; published by John Wiley and Sons, 2022 [39].
Figure 6. TSCT-TADF molecules with different types of donor–acceptor alignments, and molecular structures of mCz-Xo-TRZ and dCz-Xo-TRZ with their HOMO/LUMO distribution. Reproduced with permission from Tianyu Huang, Qi Wang, Guoyun Meng et al., “Accelerating Radiative Decay in Blue Through-Space Charge Transfer Emitters by Minimizing the Face-to-Face Donor–Acceptor Distances”; published by John Wiley and Sons, 2022 [39].
Micromachines 13 02150 g006
Figure 7. (a) The relationship between structures and HOMO levels of phenyl acridine and its skeleton-modified derivatives based on the DFT simulations. (b) Design strategy for controlling intramolecular stacking of rigid heteroaromatic compounds and the chemical structures of the designed compounds. (c) Device configuration and the related energy levels. (d) EL spectra with the photographs of devices (inset). (e) CIEy and FWHM of the EL spectra with CIE plot of EL spectra (inset). (f) EQE–current density curves with EQE–luminance curves (inset). (g) Current density–voltage–luminance curves. (h) EQEmaxs and EQEs at 100 cd m−2 (EQE100) of this work and the reported TSCT emitters. Reproduced with permission from Zhennan Zhao, Cheng Zeng, Xiaomei Peng et al., “Tuning Intramolecular Stacking of Rigid Heteroaromatic Compounds for High-Efficiency Deep-Blue Through-Space Charge-Transfer Emission”; published by John Wiley and Sons, 2022 [60].
Figure 7. (a) The relationship between structures and HOMO levels of phenyl acridine and its skeleton-modified derivatives based on the DFT simulations. (b) Design strategy for controlling intramolecular stacking of rigid heteroaromatic compounds and the chemical structures of the designed compounds. (c) Device configuration and the related energy levels. (d) EL spectra with the photographs of devices (inset). (e) CIEy and FWHM of the EL spectra with CIE plot of EL spectra (inset). (f) EQE–current density curves with EQE–luminance curves (inset). (g) Current density–voltage–luminance curves. (h) EQEmaxs and EQEs at 100 cd m−2 (EQE100) of this work and the reported TSCT emitters. Reproduced with permission from Zhennan Zhao, Cheng Zeng, Xiaomei Peng et al., “Tuning Intramolecular Stacking of Rigid Heteroaromatic Compounds for High-Efficiency Deep-Blue Through-Space Charge-Transfer Emission”; published by John Wiley and Sons, 2022 [60].
Micromachines 13 02150 g007
Figure 8. Molecular structure of Au(I)-1. (a) Schematic OLED architecture used; (b) normalized EL spectra, (c) EQE–current density, (d) current–voltage, and (e) luminance–voltage characteristics of OLEDs with Au(I)-1 in host-free and host-guest structures. Reproduced with permission from Patrick J. Conaghan et al., “Highly efficient blue organic light-emitting diodes based on carbene-metal-amides”; published by Springer Nature, 2020 [71].
Figure 8. Molecular structure of Au(I)-1. (a) Schematic OLED architecture used; (b) normalized EL spectra, (c) EQE–current density, (d) current–voltage, and (e) luminance–voltage characteristics of OLEDs with Au(I)-1 in host-free and host-guest structures. Reproduced with permission from Patrick J. Conaghan et al., “Highly efficient blue organic light-emitting diodes based on carbene-metal-amides”; published by Springer Nature, 2020 [71].
Micromachines 13 02150 g008
Figure 9. (a) Top: previously reported arylgold(III)-TADF complex a5 and its HOMO and LUMO surfaces. Bottom: donor–acceptor type C^N^C gold(III) complexes Au(III)-1–Au(III)-3 in this work; (b) normalized EL spectra; and (c) EQE–luminance characteristics of OLEDs based on Au(III)-1, Au(III)-2, and Au(III)-3 at their optimized doping concentration. Reproduced with permission from Dongling Zhou, Gang Cheng, Glenna So Ming Tong et al., “High Efficiency Sky-Blue Gold(III)-TADF Emitters”; published by John Wiley and Sons, 2020 [72].
Figure 9. (a) Top: previously reported arylgold(III)-TADF complex a5 and its HOMO and LUMO surfaces. Bottom: donor–acceptor type C^N^C gold(III) complexes Au(III)-1–Au(III)-3 in this work; (b) normalized EL spectra; and (c) EQE–luminance characteristics of OLEDs based on Au(III)-1, Au(III)-2, and Au(III)-3 at their optimized doping concentration. Reproduced with permission from Dongling Zhou, Gang Cheng, Glenna So Ming Tong et al., “High Efficiency Sky-Blue Gold(III)-TADF Emitters”; published by John Wiley and Sons, 2020 [72].
Micromachines 13 02150 g009
Figure 10. (a) Normalized EL spectra, (b) EQE–luminance, (c) current density–voltage, and (d) luminance–voltage characteristics of OLEDs with different concentrations of Au(III)-1 and v-DABNA. The image of the sensitized OLED is shown in the inset of (a). Reproduced from Ref. [40] with permission from the Royal Society of Chemistry.
Figure 10. (a) Normalized EL spectra, (b) EQE–luminance, (c) current density–voltage, and (d) luminance–voltage characteristics of OLEDs with different concentrations of Au(III)-1 and v-DABNA. The image of the sensitized OLED is shown in the inset of (a). Reproduced from Ref. [40] with permission from the Royal Society of Chemistry.
Micromachines 13 02150 g010
Figure 11. (a) Molecular structure of Cu5; (b) normalized EL spectra; (c) EQE–luminance of OLEDs based on Cu5 with various doping concentrations ranging from 2 to 6 wt%; (d) EL spectra; and (e) relative luminance versus operation time at an L0 of 1000 cd m−2 of the hyper-OLEDs with Cu5 and v-DABNA. Reproduced with permission from Rui Tang, Shuo Xu, Tsz-Lung Lam et al., “Highly Robust CuI-TADF Emitters for Vacuum-Deposited OLEDs with Luminance up to 222 200 cd m−2 and Device Lifetimes (LT90) up to 1300 hours at an Initial Luminance of 1000 cd m−2”; published by John Wiley and Sons, 2022 [62].
Figure 11. (a) Molecular structure of Cu5; (b) normalized EL spectra; (c) EQE–luminance of OLEDs based on Cu5 with various doping concentrations ranging from 2 to 6 wt%; (d) EL spectra; and (e) relative luminance versus operation time at an L0 of 1000 cd m−2 of the hyper-OLEDs with Cu5 and v-DABNA. Reproduced with permission from Rui Tang, Shuo Xu, Tsz-Lung Lam et al., “Highly Robust CuI-TADF Emitters for Vacuum-Deposited OLEDs with Luminance up to 222 200 cd m−2 and Device Lifetimes (LT90) up to 1300 hours at an Initial Luminance of 1000 cd m−2”; published by John Wiley and Sons, 2022 [62].
Micromachines 13 02150 g011
Figure 12. (a) Molecular structure of DSH; (b) simplified energy level diagram for the DSH molecule; (c) normalized EL spectra and (d) EQE–luminance characteristics of OLEDs with different concentrations of DSH. Reproduced with permission from Hartmut Yersin, Rafał Czerwieniec, Larisa Mataranga-Popa et al., “Eliminating the Reverse ISC Bottleneck of TADF Through Excited State Engineering and Environment-Tuning Toward State Resonance Leading to Mono-Exponential Sub-µs Decay. High OLED External Quantum Efficiency Confirms Efficient Exciton Harvesting”; published by John Wiley and Sons, 2022 [81].
Figure 12. (a) Molecular structure of DSH; (b) simplified energy level diagram for the DSH molecule; (c) normalized EL spectra and (d) EQE–luminance characteristics of OLEDs with different concentrations of DSH. Reproduced with permission from Hartmut Yersin, Rafał Czerwieniec, Larisa Mataranga-Popa et al., “Eliminating the Reverse ISC Bottleneck of TADF Through Excited State Engineering and Environment-Tuning Toward State Resonance Leading to Mono-Exponential Sub-µs Decay. High OLED External Quantum Efficiency Confirms Efficient Exciton Harvesting”; published by John Wiley and Sons, 2022 [81].
Micromachines 13 02150 g012
Figure 13. (a) Schematic diagram of singlet and triplet excited states split in energy by the exchange interaction (middle) and then inverted by including the double-excitation effect (right); (b) molecular structures of the heptazine analogues examined in the computational screening. Reproduced with permission from Naoya Aizawa et al., “Delayed fluorescence from inverted singlet and triplet excited states”; published by Springer Nature, 2022 [82].
Figure 13. (a) Schematic diagram of singlet and triplet excited states split in energy by the exchange interaction (middle) and then inverted by including the double-excitation effect (right); (b) molecular structures of the heptazine analogues examined in the computational screening. Reproduced with permission from Naoya Aizawa et al., “Delayed fluorescence from inverted singlet and triplet excited states”; published by Springer Nature, 2022 [82].
Micromachines 13 02150 g013
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. https://doi.org/10.3390/mi13122150

AMA Style

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(12):2150. https://doi.org/10.3390/mi13122150

Chicago/Turabian Style

Gao, Yiting, Siping Wu, Guogang Shan, and Gang Cheng. 2022. "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 13, no. 12: 2150. https://doi.org/10.3390/mi13122150

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop