Next Article in Journal
Interaction of Cerium Oxide Nanoparticles and Ionic Cerium with Duckweed (Lemna minor L.): Uptake, Distribution, and Phytotoxicity
Previous Article in Journal
Improving the Luminescence Performance of Monolayer MoS2 by Doping Multiple Metal Elements with CVT Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Status and Challenges of Blue OLEDs: A Review

by
Iram Siddiqui
1,
Sudhir Kumar
2,
Yi-Fang Tsai
1,
Prakalp Gautam
1,
Shahnawaz
1,
Kiran Kesavan
1,
Jin-Ting Lin
1,
Luke Khai
1,
Kuo-Hsien Chou
1,
Abhijeet Choudhury
1,
Saulius Grigalevicius
3,* and
Jwo-Huei Jou
1,*
1
Department of Materials Science and Engineering, National Tsing Hua University, Hsinchu 30013, Taiwan
2
Institute for Chemical and Bioengineering, ETH Zürich, 8093 Zürich, Switzerland
3
Department of Polymer Chemistry and Technology, Kaunas University of Technology, LT-50254 Kaunas, Lithuania
*
Authors to whom correspondence should be addressed.
Nanomaterials 2023, 13(18), 2521; https://doi.org/10.3390/nano13182521
Submission received: 17 August 2023 / Revised: 4 September 2023 / Accepted: 5 September 2023 / Published: 8 September 2023
(This article belongs to the Section Nanoelectronics, Nanosensors and Devices)

Abstract

:
Organic light-emitting diodes (OLEDs) have outperformed conventional display technologies in smartphones, smartwatches, tablets, and televisions while gradually growing to cover a sizable fraction of the solid-state lighting industry. Blue emission is a crucial chromatic component for realizing high-quality red, green, blue, and yellow (RGBY) and RGB white display technologies and solid-state lighting sources. For consumer products with desirable lifetimes and efficiency, deep blue emissions with much higher power efficiency and operation time are necessary prerequisites. This article reviews over 700 papers covering various factors, namely, the crucial role of blue emission for full-color displays and solid-state lighting, the performance status of blue OLEDs, and the systematic development of fluorescent, phosphorescent, and thermally activated delayed fluorescence blue emitters. In addition, various challenges concerning deep blue efficiency, lifetime, and approaches to realizing deeper blue emission and higher efficacy for blue OLED devices are also described.

1. Introduction

1.1. World Revenues and Forecasts for Lighting and Display

Organic light-emitting diodes (OLEDs) have attracted great attention due to their applications in high-contrast, innately thin, easily adaptable, full-color, flat-panel displays and solid-state lighting [1,2,3,4,5,6,7,8]. Nowadays, OLEDs are almost unanimously leading the portable display market and strongly contesting the market for large-area displays with reasonably sound efficiency and operational stability. Particularly, all leading smartphone and smartwatch companies, including Samsung, Apple, LG, Sony, Nokia, Huawei, Xiaomi, OPPO, OnePlus, and Vivo, are marketing their flagship devices with high-quality OLED displays. OLED revenues are currently driven by high-quality display applications. In addition, the lighting market is also gradually emerging. Due to their numerous inherently disruptive advantages, including a plane light form, a very high color rendering index, tunable color temperature, diffused emission, flexibility, transparency, being fully dimmable, human-friendliness, a wide viewing angle, being made with mercury-free, sustainable raw materials, being ultra-thin and light-weight, an ultimate high contrast, fast response time, and energy-saving, etc., flat panel displays and solid-state lighting panels will, sooner or later, take up a major portion of revenue and eventually disrupt conventional technologies [5,9,10,11,12,13,14].
Figure 1 compares the global revenue of OLED displays with that of LCD displays reported by DSCC’s Quarterly Display Capex and Equipment Market Share Report [15]. Overall, the revenue of LCDs experienced a catastrophic fall, as it dropped from 12.4 billion in 2016 to 5.6 billion in 2020. It is expected to drop further to less than 1 billion in 2024. In contrast, the revenue of OLED displays peaked at 13.9 billion in 2017. Since then, it has slightly oscillated around 10 billion and is projected continue to do so going into 2024. OLED is a disruptive technology in the display market, especially since its revenue surpassed that of LCD displays. Undoubtedly, OLED is and will remain the mainstream display technology for a few more decades.
A recent report from Statista depicted global annual revenue for LED lighting and estimated it to be 70 billion in 2019. An increment of 41% is expected until 2023, with a forecasted amount of 99 billion annually. However, on closer analysis, there was an approximate annual increase in value of less than 10 billion from 2019–2023 (Figure 2) [16]. This reduction in market size can be attributed to the rise of OLEDs, which are expected to overcome the LED market.
On the other hand, ElectroniCast, a US-based company, says that the estimated global revenue for the OLED lightings is set to grow speedily from 0.14 billion in markets in 2015 to nearly 7 billion by 2023, representing a compound annual growth rate (CAGR) of 7% in the next ten years (Figure 3). According to the above statics, OLED lighting revenue is expected to grow rapidly from 2015–2025. The forecast also displays a slight decline in the annual revenue around 2025, with an estimated value of 5.5 billion [17]; however, on studying the market size, OLEDs for lighting applications show a significant rise and are expected to hold the entire global market.

1.2. OLED—The Ultimate Display Technology

Twenty-five years ago, the first OLED product, a monochromatic display in car stereos, was launched by Pioneer [18]. Since then, researchers have made tremendous efforts in developing high-quality OLED displays. Nowadays, OLEDs are extensively applied in portable display devices, such as laptops, smartphones, mp3/mp4 players, digital cameras, smart watches, electric razors, and PDAs [19]. After years of struggle, large-size OLED displays are reaching the verge of a commercial breakthrough. In 2007, Sony developed the world’s first 11-inch OLED TV (XEL-1); in 2019, Sony developed the largest screen available on the market, i.e., the 77-inch A9G [20]. Other large companies, such as LG, Samsung, Panasonic, Sharp, AU Optronics, and BOE, have also launched ultra-high definition (UHD) full-color OLED TVs with various display sizes. A few years back, AU Optronics was set to demonstrate a 17.3-inch display with a 4K (225 PPI) resolution fabricated using inkjet technology. Additionally, AUO will be showcasing a 5.6-inch foldable active-matrix organic light-emitting diode (AMOLED) [21]. Meanwhile, BOE also fabricated an inkjet printed 55-inch OLED display TV with an 8K resolution and a claimed maximum brightness of 400 nits with a color gamut of 95% [22]. The global enterprise trends of OLED displays are rapidly growing, and it is clear that OLED displays will dominate the flat panel display industry in the coming years. In 2016, LG began producing their flat panel display units after long use of curved displays. This panel design has been adopted by some other major companies for applications in TVs, laptops, and desktops [23].

1.3. OLED—The Best Lighting Technology

There is a strong demand for energy-efficient and optimal next-generation lighting because approximately 20% of total global electricity is consumed for general lighting [24]. Researchers have extensively accepted that the OLEDs must become a mainstream technology for next-generation illumination sources within a few years because of their numerous distinctive features, e.g., lightweight, uniform, cold-light, human-eye friendly, and soft surface emission, feasible to form on conformable substrates, free from ultraviolet (UV) and near-infrared (NIR) emission, glare-free, environmentally friendly, high CRI, and extremely energy efficient [10,13,14,25,26]. Conventional light sources cannot provide these advantages, such as incandescent bulbs, fluorescent tubes, and light-emitting diodes (LEDs). Therefore, due to high demand for a human-friendly light source, intensive research is ongoing for better OLED lighting.
Kido et al. discovered the first white OLED in 1993 [27]. Since then, there has been rapid progress in both academic and industrial OLED research and development. In 2009, Leo et al. reported a white OLED structure showing a fluorescent tube efficacy of 90 lm/W with a CRI of 80 at 1000 cd/m2 [28]. Panasonic introduced a 4 mm2 OLED with 128 lm/W and 1 cm2 white OLED panels with an efficacy of 114 lm/W at 1000 cd/m2 [29]. Panasonic also developed a 25 cm2 white OLED panel with 110 lm/W efficacy and a lifetime of over 100,000 h (L50) at 1000 cd/m2 by employing all phosphorescent materials [30]. In 2014, Konica Minolta demonstrated an OLED with a record high efficacy of 139 lm/W at 1000 cd/m2 using all phosphorescent materials [31].
Commercially, Osram introduced the first OLED-based functional table lamp in 2008 [32,33]. In 2009, Philips started selling its OLED lighting panel under the trade name of Lumiblade [34,35]. Lumiotech introduced a high CRI of 90 for natural white lighting panels with an efficacy of 45 lm/W and a lifetime of 40,000 h (L50) at 3000 cd/m2 [36]. LG Chem. developed white OLED panels with an efficacy of 60 lm/W and a lifetime of 40,000 h (L70) at 3000 cd/m2 [34] (Table 1). Other companies such as UDC, Idemitsu, Mitsubishi Chemical, Verbatim, Fraunhofer COMEDD, and Konica Minolta have followed the same path to enhance product viability in the consumer market [37].
Furthermore, high-efficiency, tunable color-temperature, high-CRI, and low-MSI OLED panels are being fabricated, and few are exclusively available in the market. The time is not far when OLED panels will light every household, workplace, library, and even hospital, and human health will no longer be at risk.

2. The Crucial Roles of Blue Emitters

To realize high-quality OLED displays and solid-state lighting, extensively energy-efficient and long-lifetime red, green, and blue (RGB) primary emitters are indispensable. Many high-potential red and green emitters have been developed, and they are extensively accepted for commercial OLED products [41,42]. High-quality blue emitters are still a major challenge, and the development of blue-light emitting materials exhibiting high efficiency and long lifetimes is an issue of current interest.

2.1. Importance of Blue Emission

Red and green lights have been used for over a century. The need for blue light has risen to fabricate energy-efficient white light to revolutionize the lighting industry. Red, green, and blue light can provide a continuous spectrum, emitting white light. Blue emission is the color of light between violet and green, which can be classified in four different chromaticity variables representing different CIEy value ranges: ultra-deep blue (CIEy < 0.05), deep blue (CIEy < 0.10), blue (CIEy < 0.25), and sky-blue (CIEy < 0.35). The development of a deep blue emitter is an essential requirement to realize high-quality displays and solid-state lighting. In addition, promising deep-blue emitters must also match the National Television System Committee (NTSC) standard with Commission Internationale de L’Eclairage (CIE) coordinates of (0.14, 0.08) for full-color display applications [41].
In the 1990s, Shuji Nakamura and his co-workers, Akasaki and Anamo, invented the blue light-emitting diode (LED) using high-quality crystals of GaN. The main purpose of this invention was to produce energy- and resource-efficient lighting that consumes comparatively less power for illumination and is cheap enough to light every household. They were awarded a Nobel prize for this esteemed invention that consumes much less energy than other available lighting resources. These LEDs hold a long operating lifetime of over 100,000 h, which is 110 times higher than incandescent bulbs and white filament lamps, respectively [43,44,45].

2.2. To Enable a White Light

Typically, white electroluminescence in OLED devices has been produced through three major strategies: (i) mixing of primary red, green, and blue (RGB) emissions, (ii) development of a white emitter that consists of multiple spectral peaks covering the entire visible range, and (iii) utilization of complementary emitters, such as orange and blue or yellow and blue emitters. Consequently, blue emitters are crucial primary emission components for producing white emission chromaticity.

2.3. To Enable a High Color Gamut

The color gamut refers to the range of colors that any display can potentially display. There are two types of color gamut, additive and subtractive. An additive color gamut refers to color generated by mixing two or more color lights to generate a desired color. Subtractive color is generated by mixing more than one dye that absorbs certain wavelengths of light and reflects others to produce a color; it is used in color printing [46,47,48]. Considering a display’s high color gamut range, the organic emitters used in an OLED display should exhibit color purity in RGB emission [49,50]. Researchers have broadly accepted that extremely deep-blue emitters can play a considerable role in developing an incredibly high color gamut. The color gamut of NTSC can be exceptionally enhanced by replacing the sky-blue emitter with an extremely deep blue counterpart in displays [51,52,53].

2.4. To Enable a High Color Rendering Index

The color rendering index (CRI) is a significant indicator representing the quality of illumination. High CRI is a noteworthy indicator of light quality for applications in art galleries and exhibition centers because it can reveal the true color of paintings and artworks faithfully under artificial lighting. Additionally, it has been extensively confirmed that deep-blue emission plays a crucial role in achieving a high CRI. Leo’s group reported a hybrid white OLED that used phosphorescent red and green emitters with fluorescent deep-blue counterparts, which exhibited a CRI of 86 with an efficacy of 22 lm/W at 1000 cd/m2 [54]. Yang et al. developed a CRI of 91 using a triple emissive layer device architecture [55]. In 2011, Jou et al. fabricated a double emissive layer white OLED with a record high CRI of 96 and 5.2 lm/W efficacy [56]. In the industrial sector, LG Chem. and Lumiotech achieved over 90 CRI for OLED lighting panels with a power efficacy of 60 lm/W and 28 lm/W, respectively, at 3000 cd/m2 [57,58].

3. Papers

For nearly a decade, the research communities and industry have worked on high-performance long-lifetime deep blue-emitting OLED devices. OLEDs are divided into three groups based on their luminescent organic materials: fluorescent (Generation 1, Gen-1), phosphorescent (Generation 2, Gen-2), and thermally activated delayed fluorescence (TADF) (Generation 3, Gen-3). Blue fluorescent OLEDs have a good operating lifespan and emit a deep-blue light, but they are less effective at emitting light per unit area than phosphorescent OLEDs. Phosphorescent blue emitters are commonly used to achieve high performance. Nevertheless, owing to their short operating life and low color coordination, the extensive use of the phosphorescent emitters in OLEDs is limited. The trends and the developments of previously recorded blue OLED journals have been listed in this section.

Trends and Development

The trend and development of the number of journal papers on fluorescent-based blue OLEDs are illustrated in the bar graph in Figure 4. Adachi and his team reported the first blue OLED device in 1990 in Applied Physics Letters [59]. In 1992, around six fluorescent material-based blue OLED papers were reported. From 1993 to 1996, there were only two papers reported per year. However, the average number of papers per year was recorded as eight from 1997–2000. In 2001 and 2002, the number of publications decreased to one paper per year, but gradually increased in the following years. The average number of papers per year was determined to be around sixteen from 2006–2012. Furthermore, after 2012, we found a drastic enhancement in the publications, and more than 30 publications were counted per year. The highest number of journal papers per year was counted as 68; these were published over 2021.
Figure 5 displays the number of publications on phosphorescent-based blue OLEDs (PHOLED) per year. In 2001, Adachi et al. exhibited a phosphorescent blue emitter, FIrpic, and published an article in Applied Physics Letters [60]. In the succeeding year, Kawamura et al. published an article in the same journal on blue PHOLED devices fabricated using PVK as a host and Ir (III) complex as a dopant [61]. Moreover, we noticed that from 2002, the number of journal papers gradually increased, with around ten papers published each year till 2007. Furthermore, a rise in publication was observed till 2015, however, there still were fewer than fifteen papers per year. In contrast, there was a drastic increment from 2015–2019, with an average of 36 papers per year. The highest count reached eighty in 2018. On average, 66 articles were published on blue PHOLED from 2016 to 2021.
Figure 6 illustrates yearly publication on thermally activated delayed fluorescent (TADF)-based blue OLEDs. The bar graph shows that the number of publications rapidly increased since the first TADF-based blue OLED paper was reported and peaked to 71 in 8 years. In 2012, Lee et al. reported a blue TADF emitter CC2TA comprising a bicarbazole donor and phenyltriazine acceptor units in the Applied Physics Letter [62]. In 2013, Adachi and his team published three papers [63,64,65]. In 2015, a report from Mallesham et al. demonstrated pyrene excimer-based blue emission [66]. Subsequently, in the same year, Liu et al. also published an article based on designing and synthesizing blue TADF emitters for OLED devices [67]. In addition, Adachi et al. reported a sky-blue OLED device in the same year and published an article. They studied a sky-blue OLED based on a TADF molecule, 1,2-bis(carbazol-9-yl)-4,5-dicyanobenzene (2CzPN) [68]. A total of 16 articles were published in 2015. For the following years, the number of reports continuously rose, with the publication peaking in 2020 and slightly decreasing in 2021.
Figure 7 depicts the rise of blue OLED papers using fluorescent (Gen-1), phosphorescent (Gen-2) and TADF (Gen-3) emitters. The yearly paper amount for Gen-2 emitter-composed devices surpassed that of Gen-1 counterparts in 2008 (after the first paper was published in 2001), while the yearly paper amount of Gen-3-based devices surpassed that of Gen-2 in 2021 (after the first paper was published in 2012).

4. Patent

Figure 8 depicts the number of granted patents for blue OLEDs from 1995–2020. Blue OLEDs were seemingly quite challenging to devise, since there was just one patent in 1998, after the first patent in 1995. The first patent was filed by Hosokawa et al. from Idemitsu Kosan Co. Ltd. (Tokyo, Japan) [69], demonstrating the feasibility of fabricating a blue OLED on the basis of a polycarbonate containing a styryl-amine or diarylvinylene arylene skeleton.
In 1998, the first small molecule-based blue OLED patent was granted to Antoniadis et al. from Agilent Technologies Inc., wherein oxadiazole-, thiadiazole-, or triazole-based blue emitters were used [70]. There were no blue OLED patents issued in 1999 or 2000.
Figure 8. Blue OLEDs were seemingly quite challenging to devise since there was only one patent filed in 1998 after the first patent appeared in 1995. There were no blue OLED patents issued in 1999 or 2000. After that, blue OLED patents were continuously granted, peaking to 23 in the year 2017. The overall increasing trend over the past 20 and some years indicates that blue OLEDs are critically crucial. Keyword: Blue organic light-emitting diodes. Source: patentscope.wipo.int [71].
Figure 8. Blue OLEDs were seemingly quite challenging to devise since there was only one patent filed in 1998 after the first patent appeared in 1995. There were no blue OLED patents issued in 1999 or 2000. After that, blue OLED patents were continuously granted, peaking to 23 in the year 2017. The overall increasing trend over the past 20 and some years indicates that blue OLEDs are critically crucial. Keyword: Blue organic light-emitting diodes. Source: patentscope.wipo.int [71].
Nanomaterials 13 02521 g008
In 2001, Adachi et al. from Princeton University received a patent based on transition metals, wherein zinc-based derivatives exhibited both blue fluorescence and phosphorescence, the beginning of the phosphorescent mechanism [71].
After that, blue OLED patents were continuously granted, peaking to 23 in 2017. The overall increasing trend over the past 20 and some years indicates that blue OLEDs are critically crucial.

5. Performance Status

Nearly two decades ago, both academia and industries started making efforts to develop high-efficiency and long-lifetime OLED devices with deep-blue emission. Based on luminescent organic materials, OLEDs are generally classified into three main categories: fluorescent, phosphorescent, and thermally activated delayed fluorescence (TADF).
Blue fluorescent OLEDs possess a fair operational lifetime and deep-blue emission, but they are typically less efficient at emitting light per unit area than their phosphorescent counterparts.
To realize high efficiency, phosphorescent blue emitters are extensively used. However, the widespread use of phosphorescent emitters is limited due to the absence of deep-blue emission. Further, to resolve the limitation, TADF materials were developed and many more are under development. In this section, we describe the efficiency and lifetime records for previously reported blue OLEDs.

5.1. Fluorescent Blue OLEDs

5.1.1. EQE vs. CIEy

Figure 9 shows the EQE performance vs. CIEy of fluorescent blue OLEDs with dry and wet processes. The horizontal black dashed lines represent the theoretical 5% upper limit. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Some deep-blue device efficiencies were reported with a greater than 5% EQE at 1000 cd/m2 with a CIEy less than 0.1 by using the dry process. Among these, Yamazaki et al. reported a highest EQE of 11% at 1000 cd/m2 with CIE coordinates of (0.14, 0.09) using a BD-06 deep-blue emitter. The same report published a 12% EQE blue emission with coordinates of (0.14, 0.12) using a BD-05 emitter [72]. Liu et al. reported, at 100 cd/m2, a deep-blue emitter BBPA (at 5 wt%) with an 8.7% EQE at (0.15, 0.05) [73].
For wet-processed deep-blue fluorescent OLEDs, Wu et al. reported a 6.8% EQE at 100 cd/m2 with CIE coordinates of (0.16, 0.08) by using a host-free emitter T3 [74], while at 1000 cd/m2, Ma et al. reported a 6.2% EQE at (0.15, 0.08) by using a host-free emitter (G0) [75].
For emission with a greater than 0.1 CIEy, the maximum reported EQE is 10% with coordinates of (0.15, 0.28), which displays a bluish-green emission using DSA-Ph emitter doped in an MADN (host) [76]. At 100 cd/m2, the EQE achieved is 5.1% at (0.14, 0.19) by employing a host-free Blu2 blue emitter [77].

5.1.2. Power Efficacy vs. CIEy

Figure 10 shows the power efficacy (PE) performance vs. CIEy of fluorescent blue OLEDs with dry and wet processes. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively. None of the deep-blue devices show a greater than 10 lm/W power efficacy at any reported luminance, regardless of whether dry or wet processes were used.
Among the dry-processed deep-blue devices, Liu et al. published a PE of 6.1 lm/W at 100 cd/m2 with CIE coordinates of (0.14, 0.1) by using an emitter NI-2-PhTPA [78]. Li et al. reported a PE of 17 lm/W for a greenish-blue emission with coordinates (0.15, 0.26) by utilizing a DSAph emitter doped in an MBA (host) [79].
For wet-processed deep-blue devices, Ma et al. reported a maximum PE (PEmax) of 3 lm/W with CIExy of (0.15, 0.08) using a host-free emitter (G0); the same device showed a 2.4 lm/W PE at 1000 cd/m2 [75]. Li et al. reported a PE of 13 lm/W at 100 cd/m2 for a bluish-green emission with CIE coordinates of (0.15, 0.28) by using a DSA-Ph emitter [76]. Moreover, Zhen et al. reported a PE of 8.5 lm/W at 1000 cd/m2 with coordinates (0.14, 0.29) by utilizing a host-free Blu2 emitter [77].

5.1.3. Current Efficacy vs. CIEy

Figure 11 illustrates the current efficacy (CE) vs. CIEy for dry- and wet-processed fluorescent blue OLEDs. From the display application perspective, commercially viable products are those with a CIEy lying on or above the dashed line with a CE equal to 70*CIEy [80]. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Among the dry-processed devices, Kin and Hong et al. reported a maximum CE (CEmax) of 5.1 cd/A for deep-blue emission with CIExy (0.15, 0.08) by using a T1b emitter [81]. Meanwhile, Liu et al. reported a 3.6 cd/A CE at 100 cd/m2 with coordinates (0.14, 0.08) by using an NI-2-TPA emitter [78].
Takita et al. reported one commercially viable deep-blue emitter showing a 9.2 cd/A CE at 1000 cd/m2 at (0.14, 0.09) by using (at 1 wt%) a BD-06 emitter. The other commercially viable emitter reported by the same group showed a CE of 11.6 cd/A at the same luminance for blue emission with coordinates (0.14, 0.12) by using (at 1 wt%) a BD-05 emitter [72].
For the wet-processed devices, Li et al. reported a highest CE of 14.7 cd/A for a greenish-blue emission with coordinates (0.15, 0.28) by employing a DSAph emitter [76].
Zou et al. published two commercially viable deep-blue emitters with a CEmax of 5.4 cd/A at (0.16, 0.07) by using a T3 host-free emitter. Moreover, the same device showed a 5.1 cd/A CE at 100 cd/m2 with the same CIEy [74]. However, no commercially viable device has been obtained with a 1000 cd/m2 yet.

5.1.4. Lifetime and Current Efficacy vs. CIEy

Figure 12 shows the lifetime and current efficacy vs. CIEy for fluorescent blue OLEDs. The shaded area represents a CE ≥ 70*CIEy [30,37,41,42,43]. As shown, one deep-blue fluorescent device met the commercially viable criteria. Takita et al. reported a device with a lifetime of 125 h and a CEmax of 9.2 cd/A corresponding to a CIEy of 0.09 by using a BD-06 emitter [72]. Moreover, Jeon et al. reported a device closest to the criteria showing a lifetime of 168 h and a CEmax of 9.1 cd/A corresponding to a CIEy of 0.15 by using a BD-6MDPA emitter [82].
Wei et al. reported a device close to the criteria with a lifetime of 110 h and a 4.9 cd/A CEmax corresponding to a CIEy 0.11 by using a 5 wt% BN1 emitter [83].
Table 2 shows the efficiency and lifetime performance of dry-processed fluorescent blue OLEDs. All the lifetime data have also been converted to the same 1000 nits for easy comparison. Takita et al. reported a lifetime of 125 h at a 10% lifetime decay (LT90) and a maximum EQE (EQEmax) of 11.8% at 1000 nits for a deep-blue emission with a CIEy of 0.09 by using a BD-06 emitter [72]. Wei et al. reported a lifetime of 110 h at a half lifetime decay (LT50) at 1176 cd/m2 and an EQEmax of 8.6% for a blue emission with a CIEy of 0.11. A lifetime of 121 h estimated as the initial brightness is converted to 1000 cd/m2 [83].
S.W. Wen et al. reported a highest lifetime of 7000 h (at LT50) at 100 cd/m2 and a 9.7 CEmax and 11.5% EQEmax for a bluish-green emission with a CIEy of 0.32 by using a 3 wt% DSA-Ph emitter in an MADN host. At 1000 cd/m2, the same device is estimated to have a 1662 h (LT50) of lifetime [84].
Table S1 shows a comprehensive list of dry-processed blue fluorescent OLED devices, including the publication year, device structure, employed host, key dopant, doping concentration, HOMO, LUMO, EQE, PE, CE, CIE (X), CIE (Y), and reference. The years span from 2006 to 2020. A complete set of efficiencies at 100 and/or 1000 cd/m2 and/or the maximum efficiency are reported for deep-blue and blue emission.
Table S2 shows a comprehensive list of wet-processed blue fluorescent OLED devices, including the publication year, device structure, employed host, key dopant, doping concentration, HOMO, LUMO, EQE, PE, CE, CIE (X), CIE (Y), and reference. The years span from 2006 to 2020. A complete set of efficiencies at 100 and/or 1000 cd/m2 and/or the maximum efficiency are reported for deep-blue and blue emission.

5.2. Phosphorescent Blue OLEDs

5.2.1. EQE vs. CIEy

Figure 13 shows the EQE performance vs. CIEy of phosphorescent blue OLEDs with dry and wet processes. The horizontal black dashed lines represent the 20% theoretical upper limit. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Some deep-blue devices were reported with a greater than 20% EQE at a maximum luminance dimmer than 100 nits with a CIEy less than 0.1 by using dry-process. Park et al. reported an EQEmax of 25% for a deep-blue emission with coordinates of (0.14, 0.08) using a TSP01 emitter doped in a mer-Ir1 host. The same report published a 17 and 13% EQE at 100 and 1000 cd/m2, respectively [86].
Sarma et al. reported a highest EQEmax of 31% for a blue emission with CIE coordinates of (0.15, 0.3) using an 8 wt% MCPCN emitter. The same report published a 31 and 28% EQE at 100 and 1000 cd/m2, respectively [87].
For wet processes, none of the deep-blue devices shows an EQE greater than 20%. Li et al. published a highest EQEmax of 25% for a bluish-green emission at CIE coordinates of (0.15, 0.35) using a 5 wt% FIrpic doped in an m-CzPyPz host [88].
At 100 nits, no device shows a deep-blue emission or an EQE greater than 20%. However, Kim et al. reported a device close to the theoretical limit with an EQE of 18% at 100 cd/m2 for a bluish-green emission at CIE coordinates of (0.18, 0.34) using a 3 wt% fac-tris(2,3-dimethylimidazo[1,2-f]phenanthridinato-k2 C, N)iridium(III)] emitter [89].
At 1000 cd/m2, none of the devices shows a deep-blue emission. Nevertheless, Seok Oh et al. published a 24% EQE for a bluish-green emission with coordinates of (0.19, 0.41) using a 10 wt% Ir(dbi)3 doped in an mCBP-CN host [90].

5.2.2. Power Efficacy vs. CIEy

Figure 14 shows the power efficacy performance vs. CIEy of fluorescent blue OLEDs with dry and wet processes. No deep-blue devices show a power efficacy greater than 10 lumens per watt at 100 or 1000 nits. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
For dry-processed devices, Park et al. reported a 17 lm/W PEmax for a deep-blue emission with CIExy (0.14, 0.08) using a TSP01 emitter doped in a mer-Ir1 host. The same device showed a 9.1 and 5.5 lm/W PE at 100 and 1000 cd/m2, respectively [86]. Ying et al. reported a highest PEmax of 60 lm/W for a bluish-green emission with CIExy (0.14, 0.32) using a 10 wt% FIrpic doped in an mCBP:PO-T2T (host:co-host) [91].
Meanwhile, Sarma et al. reported a PE of 51 lm/W at 100 cd/m2 for a bluish-green emission with CIE coordinates of (0.15, 0.3) by using an 8 wt% mCPCN emitter doped in an MS19 host [87]. Furthermore, Son et al. reported a 42 lm/W PE at 1000 cd/m2 for a bluish-green emission with CIExy (0.16, 0.34) by using a 6 wt% FIrpic doped in a TATA host [92].
Among the wet-processed devices, none shows a deep-blue emission. However, Feng et al. reported a device close to a deep-blue region with a PEmax of 11 lm/W at CIE coordinates of (0.14, 0.11) by using a 10 wt% Ir2 doped in a t-BuCPO host. The same device showed a PE of 8.8 and 5 lm/W at 100 and 1000 cd/m2, respectively [93].
Meanwhile, Seok Oh et al. reported a highest PEmax of 57 lm/W for a bluish-green emission with coordinates of (0.19, 0.41) using an Ir(dbi)3 emitter doped in an mCBP-CN host. The same device showed a 37 lm/W PE at 1000 cd/m2 [90]. Moreover, Jou et al. reported a PE of 26 lm/W at 100 cd/m2 for a bluish-green emission with coordinates of (0.18, 0.34) using an FIrpic emitter doped in a CBP host [94].

5.2.3. Current Efficacy vs. CIEy

Figure 15 shows the current efficacy (CE) performance vs. CIEy for dry- and wet-processed phosphorescent blue OLEDs. The black dashed lines represent the minimum CE requirement for display. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Many dry-processed devices exhibit a CE much greater than 70*CIEy in the typical blue region, while only a few exhibit deep-blue emission. Among these, Park et al. reported a commercially viable device with a CEmax of 21 cd/A for a deep-blue emission at (0.14, 0.08) by using a mer-Ir1 emitter doped in a TSP01 host. The same device showed a 14 and 11 cd/A CE at 100 and 1000 cd/m2, respectively [86].
Wu et al. also reported a commercially viable device with a CEmax of 59 cd/A for a bluish-green emission with CIExy (0.13, 0.37) by using a 10 wt% DPP-2 emitter doped in a 26DCzPPy host [95]. Sarma et al. also reported a commercially viable device with a 56 cd/A CE at 100 cd/m2 for a bluish-green emission at (0.15, 0.30) by using an 8 wt% MS19 emitter doped in an mCPCN host. The same device showed a 54 cd/A CE at 1000 cd/m2 [87].
For wet-processes, no devices met the criteria to be commercially viable in the deep-blue region. However, in the typical blue region, many wet-processed devices met the criteria. Among these, Li et al. reported a CEmax of 49 cd/A for a bluish-green emission with coordinates of (0.15, 0.35) by using a 5 wt% FIrpic emitter doped in an m-CzPyPz host. The same device showed a 41 cd/A CE at 1000 cd/m2 [88]. Moreover, Wee et al. reported a 31 cd/A CE at 100 cd/m2 for a greenish-blue emission at CIExy (0.15, 0.23) by using a 10 wt% FIr6 emitter doped in an Opt-MCBP host [96].

5.2.4. Lifetime and CE vs. CIEy

Figure 16 shows the lifetime and current efficacy vs. CIEy for phosphorescent blue OLEDs. Those that fall on the grey area are plausibly commercially viable. The shaded area represents a CE ≥ 70*CIEy.
None of the deep-blue fluorescent devices met the commercially viable criteria. However, Sarma et al. reported a commercially viable device close to the deep-blue region with a lifetime of 2200 h and a CEmax of 43 cd/A corresponding to a CIEy of 0.16 by using an 8 wt% mCPCN emitter doped in an MS19 host [87].
Some phosphorescent OLEDs show relatively high current efficacy and a comparatively long lifetime in the sky-blue region. Among these, Weaver et al. reported a highest lifetime of 100,000 h and a CEmax of 32 cd/A corresponding to a CIEy of 0.38 by using a 6 wt% FIrpic doped in a CBP host [97]. Meanwhile, Konidena et al. reported a highest CEmax of 52 cd/A and a lifetime of 20 h corresponding to a CIEy of 0.32 by using a 3CNCzBN emitter [98].
Table 3 shows the efficiency and lifetime performance of dry-processed phosphorescent blue OLEDs. All the lifetime data have also been converted to the same 1000 nits for easy comparison.
No devices show a deep-blue emission. However, Jung et al. reported a device close to deep-blue emission with a highest EQEmax of 28% and a lifetime of 10,000 h (LT50) at 100 cd/m2 corresponding to a CIEy of 0.13. A lifetime of 2400 h, estimated as the initial brightness, is converted to 1000 cd/m2 [99].
Weaver et al. reported a highest lifetime 100,000 h (LT50) at 200 cd/m2, an EQEmax of 14%, and a CEmax of 32 cd/A corresponding to a CIEy of 0.38. At 1000 cd/m2, the same device is estimated to have a 37,000 h (LT50) of lifetime [97].
Table S3 shows a comprehensive list of dry-processed blue phosphorescence OLED devices, including the publication year, device structure, key dopant, doping concentration, employed host, HOMO, LUMO, EQE, PE, CE, CIE (X), CIE (Y), and reference. The years span from 2001 to 2020. A complete set of reported efficiencies (maximum at 100 and 1000 cd/m2) corresponding to the device composition are displayed for deep-blue and blue emission.
Table S4 shows a comprehensive list of dry-processed blue phosphorescence OLED devices, including the publication year, device structure, key dopant, doping concentration, employed host, HOMO, LUMO, EQE, PE, CE, CIE (X), CIE (Y), and reference. The years span from 2001 to 2020. A complete set of reported efficiencies (maximum at 100 and 1000 cd/m2) corresponding to the device composition are displayed for deep-blue and blue emission.

5.3. TADF-Based Blue OLEDs

5.3.1. EQE vs. CIEy

Figure 17 shows the EQE performance vs. CIEy of TADF blue OLEDs with dry and wet processes. The shaded area represents the upper and lower limits expected for a TADF-based device. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Some deep-blue devices were successfully fabricated with an EQE greater than 20% via a dry process. Among these, Liang et al. reported an EQEmax of 20% with CIE coordinates of (0.16, 0.03) by using an NTN-PCZ host-free emitter [104]. Meanwhile, Lim et al. reported a highest EQEmax of 28%, exceeding the 20% theoretical limit, with a CIExy (0.14, 0.09) by using a 20 wt% TDBA-SAF emitter doped in a DPEPO host. The same device showed a 24 and 18% EQE at 100 and 1000 cd/m2, respectively [105].
Moreover, some blue devices achieved an EQE much greater than 20%. Among these, Li et al. reported a highest EQEmax of 33% for a bluish-green emission at (0.17, 0.33) by using a 30 wt% TspiroS-TRZ emitter doped in a DPEPO host. The same device showed a 31 and 24% EQE at 100 and 1000 cd/m2, respectively [106]. Furthermore, Miwa et al. reported a 33% EQEmax for a greenish-blue emission with CIExy (0.16, 0.22) by using a CCX-I-6B-OC emitter doped in a PPF host [107].
For wet-processed devices, Kim et al. reported an EQEmax of 9.2% for a deep-blue emission at (0.17, 0.07) by using a TB-3Cz host-free emitter [108]. None of the devices show a deep-blue emission at 100 or 1000 cd/m2. Meanwhile, Xie et al. reported a 25% EQEmax for a greenish-blue emission with coordinates of (0.16, 0.24) by using a 2tCz2CzBn host-free emitter. The same device showed a 12 and 4.5% EQE at 100 and 1000 cd/m2, respectively [109].

5.3.2. PE vs. CIEy

Figure 18 shows the power efficacy performance vs. CIEy of TADF blue OLEDs with dry and wet processes. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
One device shows around 20 lm/W power efficacy at 100 nits with deep-blue emission via dry-processing. Lim et al. reported a highest PEmax of 20 lm/W with coordinates of (0.14, 0.09) by using a 20 wt% TDBA-SAF emitter doped in a DPEPO host. The same device showed a 14 and 7.4 lm/W PE at 100 and 1000 cd/m2, respectively [105].
Meanwhile, Ganesan et al. reported a highest PEmax of 68 lm/W for a bluish-green emission with CIExy (0.18, 0.34) by using a 22 wt% 2NPMAF emitter doped in a DPEPO host. The same device showed a 46 lm/W PE at 100 nits [110]. Moreover, Oh et al. reported a PE of 44 lm/W at 1000 cd/m2 for a bluish-green emission at (0.17, 0.33) by using a 30 wt% 23CT emitter doped in a DPEPO host [111].
For wet-processed devices, Kim et al. reported a PEmax of 4.1 lm/W for a deep-blue emission with CIExy (0.15, 0.08) by using a TB-P3Cz host-free emitter [108]. None of the devices shows a deep-blue emission at 100 and 1000 nits.
Meanwhile, Ban et al. reported a PEmax of 36 lm/W for a bluish-green emission at (0.17, 0.35) by using a MeCz-4CzCN host-free emitter [112]. Furthermore, Xie et al. reported a PE of 24 lm/W at 100 cd/m2 for a bluish-green emission with CIExy (0.18, 0.35) by using a 2PhCz2tCzBn host-free emitter. The same device showed an 8.1 lm/W PE at 1000 nits [109].

5.3.3. CE vs. CIEy

Figure 19 shows the current efficacy (CE) performance vs. CIEy for dry- and wet-processed TADF blue OLEDs. The black dashed lines represent the minimum CE requirement for display. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
One dry-processed device exhibits a near 25 cd/A CEmax with relatively deep-blue emission. Greater than 20 cd/A is also reported at 100 nits. Lim et al. reported a commercially viable device with a CEmax of 24 cd/A for a deep-blue emission at (0.14, 0.09) by using a 20 wt% TDBA-SAF emitter doped in a DPEPO host. The same device showed a 21 and 15 cd/A CE at 100 and 1000 cd/m2, respectively [105].
Meanwhile, Li et al. reported two commercially viable devices with a CEmax of 71 cd/A for a bluish-green emission at (0.17, 0.33) by using a 30 wt% TspiroS-TRZ emitter doped in a DPEPO host. The same device showed a 70 cd/A CE at 100 cd/m2. Another device showed a 56 cd/A CE at 1000 cd/m2 for a bluish-green emission at (0.19, 0.39) by using a 30 wt% TspiroF-TRZ emitter doped in a DPEPO host [106].
For wet-processed devices, none of the deep-blue devices met the criteria to be commercially viable. However, Kim et al. reported a device close to the criteria with a CEmax of 4.6 cd/A at (0.15, 0.08) by using a TB-P3Cz host-free emitter [108]. None of the devices show a deep-blue emission at 100 or 1000 nits.
Nonetheless, Xie et al. reported two commercially viable devices with a CEmax of 44 cd/A for a bluish-green emission at (0.19, 0.38) by using a 2tCz2PhCzBn host-free emitter. The same device showed an 18 cd/A CE at 1000 nits. Moreover, another device showed a CE of 38 cd/A at 100 nits for a bluish-green emission at (0.18, 0.35) by using a 2PhCz2tCzBn host-free emitter [109].

5.3.4. Lifetime and CE vs. CIEy

Figure 20 shows the lifetime and current efficacy (CEs) vs. CIEy for TADF blue OLEDs. The shaded area represents a CE ≥ 70*CIEy. None of the devices show a lifetime even close to a deep-blue region. However, a couple TADF-based OLEDs show a fair lifetime with high current efficacy in the sky-blue region. Yoon et al. reported a lifetime of 24 h and CEmax of 42 cd/A corresponding to a CIEy of 0.35 by using a 30 wt% DBA-BI emitter doped in a pCzPybCz host [113].
Table 4 shows the efficiency and lifetime performance of dry-processed TADF blue OLEDs. All the lifetime data have also been converted to the same 1000 nits for easy comparison.
None of the devices shows a lifetime in the deep-blue region. However, Su et al. reported a device close to the deep-blue region with a lifetime of 95 h (LT50) at 1000 nits and an EQEmax of 12% corresponding to a CIEy of 0.19 by using a BCz-TRZ emitter doped in an ICz-TRZ host. The same report showed a highest lifetime of 495 h (LT50) at 1000 nits and an EQEmax of 20% corresponding to a CIEy of 0.3 by using a BCz-TRZ emitter doped in a PYD2 host [114].
Table 4. Efficiency and lifetime performance of dry-processed TADF blue OLEDs. All the lifetime data have been converted to the same 1000 nits for easy comparison.
Table 4. Efficiency and lifetime performance of dry-processed TADF blue OLEDs. All the lifetime data have been converted to the same 1000 nits for easy comparison.
CIEyEQEmax
(%)
L0 (cd/m2)Lifetime (h)Lifetime (h)
@L0:1000
(cd/m2)
CEmax
(cd/A)
YearReference
0.1911.8100095 (LT50)95 (LT50)-2020[17,115,116]
0.210100045 (LT50)45 (LT50)-2020
0.2318.11000174 (LT50)174 (LT50)-2020
0.2417.3100072 (LT50)72 (LT50)-2020
0.2613.650021 (LT80)13.6 (LT80)-2017
0.32410005 (LT95)5 (LT95)40.72020[6,113]
0.324100012 (LT90)12 (LT90)40.72020
0.319.61000495 (LT50)495 (LT50)-2020[115]
0.3417.15000.28 (LT80)0.18 (LT80)-2017
0.3522.7100011 (LT95)11 (LT95)42.22020[113]
0.3522.7100024 (LT90)24 (LT90)42.22020
0.3517.35003 (LT90)1.9 (LT90)-2020[45,117]
0.3517.350019.1 (LT75)12.4 (LT75)-2020
0.3514.85002.4 (LT90)1.5 (LT90)-2020
0.3514.850014.1 (LT75)9.1 (LT75)-2020
0.3512.95002.8 (LT90)1.8 (LT90)-2020
0.3512.950019.9 (LT75)12.9 (LT75)-2020
0.3512.25001.7 (LT90)1.1 (LT90)-2020
0.3512.25009.9 (LT75)6.4 (LT75)-2020
Table S5 shows a comprehensive list of dry-processed blue TADF OLED devices, including the publication year, device structure, key dopant, doping concentration, employed host, HOMO, LUMO, EQE, PE, CE, CIE (X), CIE (Y), and reference. The years span from 2015 to 2020. A complete set of reported efficiencies (maximum at 100 and 1000 cd/m2) corresponding to the device composition are displayed for deep-blue and blue emission.
Table S6 shows a comprehensive list of wet-processed blue TADF OLED devices, including the publication year, device structure, key dopant, doping concentration, employed host, HOMO, LUMO, EQE, PE, CE, CIE (X), CIE (Y), and reference. The years span from 2016 to 2020. A complete set of reported efficiencies (maximum at 100 and 1000 cd/m2) corresponding to the device composition are displayed for deep-blue and blue emission.

6. Chemical Structures and Characteristics of Blue Emitters

As mentioned earlier, the materials can be classified into three categories; fluorescent, phosphorescent, and TADF, which can be further sub-categorized according to their derivatives. This section describes the historical development and current advancement of fluorescent, phosphorescent, and TADF blue emitters based on their structural importance, color purity, and efficiency.

6.1. Fluorescent Type

Over the past two decades, researchers have developed various kinds of blue-fluorescent emitters, such as distyrylarylene (DSA), spiro-shaped, biaryl, silane, and polycyclic aromatic hydrocarbon (e.g., pyrene, anthracene, carbazole, fluorene, etc.) derivatives to enhance the performance of blue OLEDs. Among them, polycyclic aromatic hydrocarbon (PAH)-based small molecules, oligomers, starburst, tetrahedral, dendrimer derivatives, and long-chain polymers are extensively applicable as emitters and/or host materials in OLED devices. In this section, we comprehensively review nine series of fluorescent-type blue-light emitting materials (Figure 21).
In 1992, Hamada et al. reported a multi-layer blue OLED device employing an oxadiazole derivative, 1,3-bis(N,N-dimethylaminophenyl)-1,3,4-oxadiazole (OXD-8), as an emitter for the first time (Figure 22). It showed a maximum luminance of over 1000 cd/m2 and an emission peaking at 470 to 480 nm [118,119]. In the same year, Leising et al. reported an EQE of 0.05% for a single-layer blue OLED device using poly(p-phenylene) (PPP) (Figure 14) [120,121].
Hosokawa et al. also observed a blue emission for a non-conjugated polymer-based emitter with an electron donor polycarbonate-containing styrylamine repeating unit (SA-PC) (see Figure 14) in a multilayered device. The power efficacy was 0.05 lm/W with a peak emission at 460 nm [122].

6.1.1. Distyrylarylene (DSA)

Over the last twenty years, researchers have extensively investigated the electroluminescent characteristics of DSA, DSA-ph, and their numerous derivatives as blue-light emitting materials in blue OLED devices [122,123,124,125,126]. In 1990, Hosokawa et al. measured light-emitting characteristics of DSA derivatives and reported that the exciplex formation or charge transfer complex with a hole-transporting layer drastically influenced device performance [123].
Researchers at Idemitsu Kosan Co. Ltd. reported that the formation of exciplex or charge transfer complexes can be avoided by introducing electron donor groups, such as ethyl, methoxy, diethylamine, phenylamine, and N-ethyl-carbazole, in DSA derivatives (Figure 23).
Later in 1993, Tokailin et al. reported a series of nonplanar DSA derivatives with different DSA cores (Figure 24), namely, DTVB and DTVX, and different alkyl group substituents at the end of the molecule, namely, DPVBi, DTVBi, and DTBPVBi, as blue emitters. They also investigated the electroluminescence characteristics of these emitters in a multi-layer OLED device (ITO as anode; TDP as HTL; DPVBi or DTVBi or DTBPVBi as EML; Alq as ETL; Mg:Ag as cathode). The OLED device fabricated with a DPVBi emitter exhibited, at 1000 cd/m2, a power efficacy of 0.6 lm/W and maximum luminance of 2280 cd/m2, with blue emission peaking at 475 nm. The EL spectra did not exhibit any emission in the longer wavelength region, supporting that both DPVBi and DTVBi had entirely avoided the formation of exciplex or charge transfer complexes [127].
High device efficiencies and long operation lifetimes were realized in 1995 by doping a fluorescent styrylamine-based guest, BCzVB, into a DSA derivative, DPVBi, host material. Hosokawa et al., in 1995, utilized a DDPVBi as a blue host material and styrylamine (known as BCzVB) as a dopant [126]. Further improvement of this system was later published, in which the current efficacy reached 10.2 cd/A at 1.89 mA/cm2 with 1931 CIE coordinates of (0.17, 0.33) and a half-life of 20,000 h at an initial brightness (Lo) of 100 cd/m2. When oligoamine was used as a hole-injection layer for this device, the operational lifetime was further improved to 10,000 h with Lo = 500 cd/m2 [84]. It appears that the sky-blue emission was specifically designed by Idemitsu Kosan for the application of color-changing media (CCM) technology for full-color organic EL devices [128].
For a 3 wt% DSA-ph emitter doped with a 2-methyl-9,10-di(2-napthyl)anthracene (MADN) host, the device exhibited a half-decay lifetime of 46,000 h (LT50) at an initial brightness of 100 cd/m2. The resultant OLED also showed a much higher power efficacy of 5.5 lm/W and a current efficacy of 9.7 cd/A at 20 mA/cm2 with a sky-blue emission (0.16, 0.32) [84].
In 2007, Ho et al. reported a new series of emitters based on a seven-membered N-hetero-cyclic core structure of iminodibenzyl-distyrylarylene (IDB). In the distyrylarylene structure, a replacement of the amino group by an iminodibenzyl group improved the thermal properties (Tg varying between 119 and 137 °C) and slightly blue-shifted the emission with increasing phenyl units. The emission wavelength for DSA-ph is 458 nm, while it was 449, 447, and 443 nm for IDB-ph, IDB-diph, and IDB-triph, respectively [129].
In 2018, Wanshu Li et al. reported a wet-processed simple structure and high efficiency OLED using a DSA-Ph emitter doped in an MADN host. In this work, MADN is also used as a hole-transporting material (HTL). At 3 wt% dopant concentration, a 10% EQEmax was reported at CIE coordinates of (0.15, 0.28). Moreover, a high CE of 15 cd/A and 13 lm/W PE was obtained [76].

6.1.2. Pyrenes

In the last fifteen years, numerous pyrene derivatives and hybrid-based blue-emitters were investigated in OLED devices due to their rigid planar structure [130,131], high thermal stability [70,132,133], high fluorescence quantum yield [134,135,136,137,138], long fluorescent lifetime [139,140], and effective carrier mobility [141,142].
However, the utilization of pyrene-based blue emitters was limited, owing to their strong tendency to form excimers and π-aggregates, which would cause red shifts and quenching of fluorescence, resulting in a lower solid-state PL quantum yield and quantum efficiency. Numerous efforts in the structural alteration of pyrene-based compounds have hence been made to enhance the PL quantum yield and diminish the aggregation characteristics.
In recent years, pyrene derivatives were extensively employed in efficient blue OLEDs, including polypyrenes [134,143,144], dipyrenylbenzene [145], alkynylpyrene [146,147], pyrene-functionalized calix[4]arenes [148], aryl-functionalized pyrenes [116,136,149,150,151], as well as oligothiophenes with pyrenyl side groups [152], pyrene-carbazole [132,135,146,153,154], pyrene-fluorene [155,156,157,158,159,160,161,162,163], pyrene-triphenylamine [142,164,165], and pyrene-anthracene [166] hybrids, due to their good emissive characteristics and hole mobility.
In 2011, Figueira-Duarte and Müllen comprehensively reviewed the synthesis techniques for numerous types of small molecular pyrene, oligo-pyrene, pyrene-dendrimer, and polypyrenes. They also reviewed the thermal, electrochemical, photophysical, and optical characteristics of all pyrene-based emitters and their electroluminescence characteristics in OLED devices [167].
In this part, we have only covered the pyrene derivatives that were reported in the past few years. The Sonogashira, Sonogashira-Hogihara, Suzuki, Suzuki-Miyaura, Heck, and Yamamoto-coupling reactions have been extensively used to synthesize pyrene hybrid derivatives. The sky-blue emitter, 1-((9,9-diethyl-9H-fluorene-2yl)ethynyl)pyrene (PA-1) (Figure 25), was synthesized by the Sonogashira method via a cross-coupling reaction of 1-bromopyrene with 9,9-diethyl-2ethynyl-9H-fluorene by using a Pd(PPh3)2Cl2/PPh3/CuI catalytic system. The PA-1 exhibited a purplish-blue emission with CIE coordinates of (0.15, 0.06), a power efficacy of 1.2 lm/W, a current efficacy of 1.9 cd/A, and EQE of 3.5% at 100 cd/m2 as a 3 wt% sky-blue emitter doped into the polarity matching the CBP host [162].
A series of pyrenoimidazole-based emitters that possess markedly high thermal characteristics were synthesized in three simple steps. Firstly, pyrene-4,5-dione was prepared via the oxidation of pyrene in the presence of ruthenium trichloride and sodium periodate. Subsequently, the pyrene-4,5-dione was reacted with the corresponding aldehyde in the presence of ammonium acetate to form sparingly soluble pyrenoimidazoles. Finally, the alkylation reaction was either executed under phase transfer catalysis to synthesize mono-pyrenoimidazoles or by employing the potassium carbonate/dimethylformamide (K2CO3/DMF) to synthesize bis-pyrenoimidazoles. These rigid pyrenoimidazole moiety-based hybrid compounds offered a very high thermal stability with decomposition temperatures from 462 to 512 °C.
With a CBP host doped with a 3 wt% mono-pyrenoimidazole, 9-butyl-10-(pyrene-1-yl)-9H-pyreno[4,5-d]imidazole (Figure 25), the resultant multi-layer OLED device showed, at 100 cd/m2, for example, a power efficacy of 1.2 lm/W, a current efficacy of 2.4 cd/A, and an external quantum efficiency (EQE) of 2.2% with CIE coordinates of (0.15, 0.32). The resultant device also exhibited a maximum luminance of 2740 cd/m2 [168].
It is well established that pyrene-based emitters, either molecular or polymeric, possess a planar or partial planar molecular structure, which tends to show excimer emissions and lower quantum efficiencies, resulting in an extensive bathochromic shift in emission spectra owing to the π–π stacking of pyrene chromophore units in the solid state.
In 2012, Hu and Yamato et al. synthesized a series of pyrene-based Y-shaped blue emitters using the Suzuki cross-coupling reaction of 7-tert-butyl-1,3-dibromopyrene with p-substituted phenylboronic acids. These emitters realized a low degree of π-stacking due to the steric effect of the tert-butyl group in the pyrene ring at the 7-position [134,169].
In the search for new light-emitting compounds, Promarak et al. synthesized a series of solution-processable pyrene-functionalized carbazole derivatives, N-dodecyl-3,6-di(pyren-1-yl)carbazole (CP2), N-dodecyl-1,3,6-tri(pyren-1-yl)carbazole (CP3) and N-dodecyl-1,3,6,8-tetra(pyren-1-yl)carbazole (CP4) (Figure 25), by using the Suzuki-coupling reaction [170].
These pyrene-carbazole hybrids showed a glass-transition temperature and decomposition temperature ranging from 70 to 170 °C and 375 to 407 °C, respectively. Among the four pyrene-moieties containing carbazole compounds, CP4, showed the highest thermal stability (Tg = 170 °C and Td = 407 °C).
However, these materials showed an undesirable donor-acceptor (DA) behavior because the carbazole central unit acts as a donor and the pyrene functional group acts as an acceptor. It is known in the literature that the DA molecular systems tend to extend π-conjugation and generate an intramolecular charge transfer (ICT) state, inducing an unwanted bathochromic shift in the search for a deep-blue emission.
An OLED device based on non-doped CP4 showed, by thermal evaporation deposition, for example, a maximum current efficacy of 6.7 cd/A and maximum luminance of 25,000 cd/m2 with CIE coordinates of (0.14, 0.26), and 2.5 cd/A and 9700 cd/m2 with CIE coordinates of (0.21, 0.28) via spin-coating [170].

6.1.3. Anthracenes

Adachi et al. reported a 9,10-diphenyl-anthracene (DPA) emitter with a 100% photoluminescence quantum yield (PLQY) [59]. The other major blue-doped emitter was developed by Shi et al. at Kodak in 2002 and utilized diphenylanthracene derivative 9,10-di(2-naphthyl)anthracene (ADN) (Figure 26) as a host and 2,5,8,11-tetra(t-butyl)-perylene (TBP) as a dopant to generate a blue emission of CIE (0.15, 0.23). The blue device showed a current efficacy of around 3.5 cd/A with a half-life of 4000 h at an initial light output of 700 cd/m2 [171].
In 2007, Shih et al. developed a blue-emitting material 2-tert-butyl-9,10-bis[4-(1,2,2-triphenylvinyl)phenyl]anthracene (TPVAn) (Figure 26) that contained an anthracene core and two tetraphenylethylene end-capped groups. The material possessed a high Tg of 155 °C and was morphologically stable due to the presence of the sterically congested terminal groups. A bright saturated-blue emission with CIE coordinates (0.14, 0.12) along with a 5.3% EQE crossing the theoretical limit of fluorescent materials was reported, the best performance among the non-doped blue devices [172].
In 2008, three materials 9,10-Bis(3′,5′-diphenylphenyl)anthracene [MAM], 9-(3′,5′-diphenylphenyl)-10-(3‴,5‴-diphenylbiphenyl-4″-yl)anthracene [MAT], and 9,10-bis(3″,5″-diphenylbiphenyl-4′-yl)anthracene [TAT] (Figure 26) were designed and emitted in the range of 439–445 nm. MAM and MAT have a Tg higher than 64 °C, while TAT has above 150 °C. The Tg for TAT is twice as high as the commonly used host material DPVBi (64 °C) and slightly higher than MADN (120 °C). The CIE coordinates were (0.15, 0.08) with an EQE of 7.2% [173].
In 2010, Zheng et al. reported a series of non-doped devices with three newly designed emitters, 2-tert-butyl-9,10-bis(9,9-dimethylfluorenyl) anthracene (TBMFA), 2-tert-butyl-9,10-bis[4-(2-naphthyl)phenyl] anthracene (TBDNPA), and 2-tert-butyl-9,10-bis[4-(9,9-dimethylfluorenyl)phenyl] anthracene (TBMFPA) (Figure 26), using naphthalene or 9,9-dimethylfluorene side units. Due to the anthracene unit, the resultant devices showed a deep-blue emission. These materials also showed a high Tg ≥ 133 °C, due to the presence of sterically congested terminal groups. The TBDNPA-based device outperformed the others by showing a low turn-on voltage of 3.0 V with an initial brightness of 1 cd/m2. The device displayed an EQE of 5.2% at 8.4 mA/cm2 with CIEy < 0.1 [174].
In 2011, Kim et al. synthesized four molecules displaying blue emission using anthracene moieties. The material, 3-(anthracen-9-yl)-9-ethyl-9H-carbazole (AC), 3,6-di(anthracen-9-yl)-9-ethyl-9H-carbazole (DAC), 3-(anthracen-9-yl-)-9-phenyl-9H-carbazole (P-AC), and 3,6-di(anthracen-9-yl)-9-phenyl-9H-carbazole (P-DAC) (Figure 26), was also comprised of covalent carbazole units. Among these, the undoped device with P-DAC showed a near deep-blue emission at (0.16, 0.13) with a luminance efficacy of 3.1 cd/A and EQE of 2.8%. Moreover, P-DAC as a host material doped with a BDAVBi emitter displayed a 7.7 cd/A CE and 4.7% EQE with CIE coordinates of (0.15, 0.21) at 100 mA/cm2. A negligible efficiency roll-off is observed over a broad range of current density [175].
In 2012, Zhuang et al. reported a series of non-doped devices using emitters 2-(4-(anthracen-9-yl)phenyl)-1-phenyl-1H-phenanthro[9,10-d]imidazole (ACPI), 2-(4-(10-(naphthalen-1-yl)anthracen-9-yl)phenyl)-1-p-henyl-1H-phenanthro[9,10-d]imidazole (1-NaCPI), and 2-(4-(10-(naphthalen-2-yl)anthracen-9-yl)phenyl)-1-phenyl-1H-phenanthro[9,10-d]imidazole (2-NaCPI) (Figure 26). The materials possess good film-forming capability and a high Td (290–354 °C). Among them, ACPI displayed the highest CEmax, at 1.6 cd/A at a low turn-on voltage of 3.0 V corresponding to CIExy (0.16, 0.17). A negligible efficiency roll-off is observed at high current densities [176].
In 2019, a material (4-(diphenylamino)phenyl)anthracen-9-yl)pyren-1-yl)-N,N-diphenylaniline (p-TPA-AP-TPA) (Figure 26) was synthesized with a dual-core moiety of anthracene and pyrene with a triphenylamine (TPA) side group attached at a para position to control the emission in the blue region. Moreover, the bulky molecule yielded a twisted molecular structure that prevents intermolecular interactions. The PL emission was around 473 nm in the film with a high Tg of 194 °C. The device showed a low turn-on of 3.1 V with an EQE of 8.4% and a maximum luminance of 7100 cd/m2 at CIE of (0.14, 0.15) [177].

6.1.4. Fluorenes

Fluorenes were first reported in 2006 by Lai et al. for OLED applications. They designed oligofluorenes with a certain length to obtain a high quantum yield (ՓPL = 38–75%) for a pure-blue emission. The PL emission was observed around 400 nm. An emissive layer of 130 nm was spin-casted and displayed a peak brightness of 1300 cd/m2 at a turn-on voltage of 5.3 V with a low current density of 90 mA/cm2. The CIE coordinates of (0.15, 0.09) displayed a deep-blue emission with an EQE of 2.0% and CE of 2.1 cd/A. These hindered six-arm architectures (highly branched and dendritic structures) suppress aggregation and self-quenching while the triazatruxene core facilitates charge injection, transport, and combination. Moreover, the powerful microwave-enhanced synthesis diminished impurities and chain defects that may result in quenching or keto-defects [178].
In 2007, Lai et al. reported a kinked star-shaped fluorene/triazatruxene co-oligomer (C1, C2 and C3) (Figure 27) that can effectively suppress aggregation and crystallization. The HOMO level was well matched to that of the anode’s work function that significantly improved the carrier transportation. The single-layer EL showed almost no change in color even at a higher voltage. The device displayed a low turn-on voltage of 3.3 V, EQE of 2.2%, and maximum luminance of 2400 cd/m2 [179].
In 2008, Jou et al. reported 2,7-bis{2[phenyl(m-tolyl)amino]-9,9-dimethyl-fluorene-7-yl}-9,9-dimethyl—fluorene (MDP3FL) (Figure 27) blue emitter doped in a low-polarity host, 4,4′-bis(9-carbazolyl)-biphenyl (CBP). The neat film of MDP3FL exhibits an EQE of 1.9% at 100 cd/m2 displaying a CIE of (0.15, 0.12). When doped with a 10 wt% in CBP, a 5.1% EQE was observed for a deep-blue emission at CIE (0.14, 0.08) [180].
In 2009, Wang et al. reported a solution-processable fluorescent π-conjugated dendrimer G0 (Figure 27). The material displayed a CEmax of 5.3 cd/A with a deep-blue emission at (0.15, 0.09). The EQEmax was 4.2% while 3.6% at 1000 cd/m2 [75].
Figure 27 shows the fluorene derivatives extensively utilized in efficient blue OLEDs, such as fluorene-bridged anthracene, [181] oligofluorenes, [74,77,182,183,184] p-difluorophenylene [185], bis(4-diphenylaminophenyl)carbazole end-capped fluorene [186], phosphine oxide [187,188], perfluorinated polymer [189], poly(3,6-Dimethoxy-9,9-dialkylsilafluorene)s [190], and polycyclic [191].
In 2020, Liu et al. reported a blue dye TPA-DFCP with a twisted D–π–A configuration where triphenylamine (TPA), benzonitrile (CP), and 9,9-dioctylfluorene served as D, A, and π-conjugation units, respectively. TPA is a strong electron donor group and has a high HOMO level. Moreover, a certain twist angle is present in the molecular space of TPA to avoid molecular accumulation. The non-doped device showed a high EQE of 7.7% with CIE coordinates of (0.15, 0.07) displaying a deep-blue emission. Meanwhile, the doped device showed an EQEmax of 8.3% and a radiative exciton utilization energy (ηr) of 58% [192].

6.1.5. Biaryl

In 2005, Fisher et al. reported two materials: N,N′-diethyl-3,3′-bicarbazyl (DEC) and 1,4,5,8-N-pentamethylcarbazole (PMC) (Figure 28). Both the materials can be used as a hole-blocking layer and blue emitter. The device based on a DEC emitter doped in a DPVBi host displayed an EQE of 3.3%, CE of 4.7 cd/A, and PE of 1.3 lm/W at CIE coordinates of (0.15, 0.10) [193].
In 2006, Tseng demonstrated a device based on a 7,8,10-triphenylfluoranthene (TPF) emitter doped in a dipyrenylfluorene (DPF) host. With a 6 wt% of the TPF dopant, the device displayed a 3.3 lm/W PE and a 2.5% EQE at CIE coordinates of (0.16, 0.18) [194].
In 2007, Jou et al. reported three emitting materials, namely, trans-1,2-bis(6-(N,N-di-p-tolylamino)-naphthalene-2-yl)ethene (BNE), 2-(N,N-diphenylamino)-6-[4-(N,N-diphenylamino)styryl]naphthalene (DPASN), and di(triphenyl-amine)-1,4-divinylnaphthalene (DPVP) (Figure 28). Among them, BNE displayed the highest EQE, at 6.0%, and a PE of 12.5 lm/W at a low turn-on voltage of 3 V for a pan-blue emission at (0.19, 0.31) [195].

6.1.6. Spiro-Shaped

In 2010, Jeon et al. reported a series of spiro-based emitters, namely, N,N,N′,N′-tetraphenylspiro[fluorene-7,9′-benzofluorene] (SFBF)-5,9-diamine (BD-6DPA), N,N′-di-(2-naphthyl)-N,N′-diphenyl-SFBF-5,9-diamine (BD-6NPA), N,N′-diphenyl-N,N′-di-m-tolyl-SFBF-5,9-diamine (BD-6MDPA), and N,N′-diphenyl-N,N′-bis(4-(trimethylsilyl)phenyl)-SFBF-5,9-diamine (BD-6TMSA) (Figure 29), synthesized using an animation reaction. Among them, BD-6MDPA displayed the highest CE, at 9.1 cd/A at 6.5 V, and an EQE of 8.2% with coordinates of (0.13, 0.15). The device showed a full width at half maxima of 48 nm and an emission peaking at 463 nm [82].
In 2013, Hou’s group designed and synthesized a series of fluorinated 9,9′-spirobifluorene derivatives (SFs). The photophysical properties, energy levels, and thermal properties can be tuned using different substitutional groups of electron-withdrawing moieties, such as F and CF3. Among the series, a 2,2′,7,7′-tetrakis(3-fluorophenyl)spiro-9,9′-bifluorene (Spiro-(3)-F) (Figure 29) blue emitter displayed the best result in a non-doped stage with an emission at CIE (0.16, 0.12) peaking at 408 nm. These SFs also served as a host for blue-emitting dopants, especially Spiro-(3)-F, with a low turn-on voltage of 3.4 V showing a CE of 6.7 cd/A and an EQE of 5.0% [196].

6.1.7. Silanes

In 2001, Yu et al. reported for the first time a silane group as a blue emitter, tris(2,3-methyl-8-hydroxyquinoline) aluminum (III) (Alm23q3). The presence of two electron-donating methyl groups widened the HOMO/LUMO gap and blue-shifted the EL peaks to 470 nm. The resultant device showed a power efficacy of 0.6 lm/W at 300 cd/m2 [197].
In 2002, Chan et al. reported a molecular glass material (4-(5-(4-(diphenylamino)phenyl)-2-oxadiazolyl)phenyl)triphenylsilane (Ph3Si(PhTPAOXD)) (Figure 30) as a blue emitter. The material possessed a Tg of 85 °C, making it less vulnerable to heat and hence more stable. The device showed a maximum luminance of 19,000 cd/m2 with an EQE of 2.4% at CIE coordinates of (0.16, 0.18) [198].
In 2015, Liu et al. reported an organosilane compound, bis(4-(1-phenylphenanthro[9,10-d]imidazol-2-yl)phenyl)diphenylsilane (Si(PPI)2) (Figure 30), as a bipolar host. The incorporation of one tetraphenylsilane and two PPI groups made the compound thermally stable with a high Td of 528 °C and a Tg of 178 °C. Moreover, when doped with a 10 wt% An(PPI)2 emitter, the device showed a PE of 8.0 lm/W and an EQE of 6.1% at CIExy (0.18, 0.17) [199].

6.1.8. Carbazole

In 2018, Jou et al. reported a carbazole-based deep-blue fluorescent emitter, 6-((9,9-dibutyl-7-((7-cyano-9-(2-ethylhexyl)-9H-carbazol-2-yl)ethynyl)-9H-fluoren-2-yl)ethynyl)-9-(2-ethylhexyl)-9H-carbazole-2-carbonitrile (JV55) (Figure 31), that showed a maximum EQE of 6.5% at CIE coordinates of (0.16, 0.06). The deep-blue emission realized a 101% color saturation according to the NTSC standard. The very high PLQY of 91%, effective host-guest energy transfer, efficient triplet energy utilization, and low doping concentration realized a low roll-off [200].
In 2020, Yang et al. reported a blue fluorescent material with a twisted A–π–D–π–A configuration, namely, CzPA-F-PD (Figure 31). This material exhibited dual fluorescent properties that emitted blue and red emissions under PL and EL processes when triggered by trifluoroacetic acid (TFA). The property had been reversed to its initial state on neutralizing TFA with triethylamine (TEA). It was observed that hydrogen protonation of TFA amplified the lone pair electrons around the nitrogen atoms in pyridine, enhancing the electron acceptor ability. This scheme also played a significant role during the dual fluorescent process. The material possessed a good exciton utilization efficiency of 67%, exceeding the upper limit of spin-statics for fluorescent materials [201].

6.1.9. Oxidiazole

In 2018, Wu et al. reported that triazine, oxadiazole, triazole, cyno-substituted benzene, and benzothiadiazole worked as acceptor units while carbazole, arylamine, phenothiazine, and their derivatives were utilized as donor moieties for a bipolar structure. Consequently, they utilized phenothiazine and 9,9-diphenyl-9,10-dihydroacridine as a donor unit and oxadiazole derivatives, 2,5-bis(4-bromophenyl)-1,3,4-oxadiazole and 2-(4-bromophenyl)-5-phenyl-1,3,4-oxadiazole, as acceptor units to synthesize a series of bipolar emitters, namely, 2DPAc-OXD, DPAc-OXD, 2PTZ-OXD, and PTZ-OXD (Figure 32). All the materials possessed a high decomposition temperature above 358 °C, with emissions in the PL range of 435–512 nm. For OLED applications, these materials showed a moderate turn-on voltage of around 4.2 V along with a highest EQE of 4.0% when using the material 2PTZ-OXD [202].
Later that year, Wang et al. reported a bipolar molecule, bis(4’-(9H-carbazol-9-yl)-[1,1-biphenyl]-4-yl)-1, 3, 4-oxadiazole (Oxd-bCz) (Figure 32), by incorporating electron-donor and electron-withdrawal moieties. The material emitted a purple-blue fluorescence at 431 nm when diluted in dichloromethane. From a device perspective, the device turned on at 3.6 V and displayed a maximum EQE of 5.6% at CIEy ~ 0.07 [203].

6.2. Phosphorescent Type

Developing phosphorescent emitters is essential as they have the potential to realize a high-efficiency OLED. In 1998, Forrest and Thompson et al. discovered a phosphorescent emitter to develop a nearly 100% internal quantum efficiency because of their ability to efficiently harvest both singlet and triplet excitons [204,205,206,207,208]. Since then, very high-efficiency phosphorescent green and red emitters have emerged. However, it was a great challenge to synthesize a highly energy-efficient phosphorescent blue emitter that exhibits a long lifetime with high color purity. Deep blue emitters should have a high triplet energy (~3 eV), a high photoluminance quantum yield (PLQY), and a short excited-state lifetime. To realize the above-mentioned features, considerable efforts have been made to develop a high performance deep-blue emitter.
On the basis of previous investigations, it is well established that complexes of heavy metals such as iridium (III), platinum (II), and ruthenium (II) have been utilized to synthesize promising blue emitters [209,210,211]. Among them, iridium complex-based phosphors have shown enormous potential to realize high-efficiency blue OLEDs. Moreover, to achieve a deep-blue emission, thermally and electrochemically stable fluorine-free cyclometalating ligands [210,212,213] were introduced either by using electron-withdrawing groups [214,215] or electron-deficient heterocycles [215,216]. Furthermore, to stabilize HOMO, ancillary ligands with strong ligand field strength, such as isocyanide or phosphine, were utilized [215,217].

6.2.1. Iridium (III) Complexes

In 2001, Adachi et al. reported a sky-blue phosphorescent emitter, FIrpic (Figure 33), with an EQE of 5.7% and CIE coordinates of (0.16, 0.29) [60]. In 2003, Holmes et al. synthesized a blue emitter, iridium(III) bis(2,4-difluorophenylpyridinato)tetrakis(1-pyrazolyl)borate (FIr6) (Figure 33), with an EQE of 12% at (0.16, 0.26) [218]. In the same year, Bazan’s group reported a series of anthracene-containing binaphthol emitters with CIE coordinates of (0.18, 0.21) without revealing device performance [209].
In 2005, Holmes reported fac- and mer-isomers of fluorine-free emitters tris(phenyl-methyl-benzimidazole)iridium(III) (f-Ir(pmb)3) and (m-Ir(pmb)3), respectively (Figure 33). When f-Ir(pmb)3 was doped into a p-bis(triphenylsilyly)benzene (UGH2) host, the device showed an EQE of 5.8% for a deep-blue emission with CIE coordinates of (0.17, 0.06) [210].
Moreover, other fluorine-free Ir complexes which resulted in a blue emission were also reported. They were mer-tris(Ndibenzofuranyl-N-methylimidazole) iridium(III) [Ir-(dbfmi)] (Figure 33) [212] and tris(1-cyanophenyl-3-methylimidazolin-2-ylidene-C,C2′) iridium(III) [Ir(cnpmic)] [213].
In 2009, Kang et al. introduced a blue emitter based on an iridium(III) complex of 2′,6′-difluoro-2,3′-bipyridine (fac-Ir(dfpypy)3) (Figure 33) with CIE coordinates of (0.14, 0.12) [219]. Meanwhile, the same group reported a comparatively more efficient device with an EQE of 10% at 1000 cd/m2 at (0.15, 0.24) by employing its modified complex, (dfpypy)2Ir(μ-Cl)2) [220].
In 2010, Kido et al. reported a halogen-free blue emitter, mer-tris(N-dibenzofuranyl-N′-methylimidazole)iridium(III) (Ir(dbfmi)) (Figure 33). By doping a 10 wt% Ir(dbfmi) into a 3,6-bis(diphenylphosphoryl)-9-phenylcarbazole (PO9) host, the resultant device showed a maximum EQE of 19% with CIExy (0.15, 0.19) [212]. In 2011, Hsieh et al. reported blue emitters of iridium bis(carbene) complexes, namely, Ir(mpmi)2(pypz), Ir(fpmi)2-(pypz), and Ir(fpmi)2(tfpypz) (Figure 33), that showed an EQE of 15, 14, and 7.6% and CIExy of (0.14, 0.27), (0.14, 0.18), and (0.14, 0.10), respectively [221].
In 2012, Fan et al. reported two phosphoryl/sulfonyl-substituted iridium complexes, POFIrpic and SOFIrpic (Figure 33), by introducing phosphoyl/sulfonyl moieties into the 5′-position of a phenyl ring in the structural frame of FIrpic. By doping a POFIrpic emitter into PVK:OXD-7 hosts, the solution-processable blue device showed CIE coordinates of (0.16, 0.29) and an EQE of 3.5% at 1000 cd/m2 [222]. In 2013, Fan et al. modified the FIrpic by introducing fluorine, chlorine, and bromine atoms to the 4-position of a pyridine ring. The 4-F-FIrpic-based device showed an EQE of 15% for a blue emission with CIExy (0.15, 0.28) [223]. Meanwhile, Kessler et al. synthesized a heteroleptic iridium complex based on a 4-(tert-butyl)-2′,6′-difluoro-2,3′-bipyridine ligand (FK306) (Figure 33) that showed an EQE of 13% at 1000 cd/m2 for a greenish-blue emission at (0.16, 0.25) [224]. In the same year, Yoon et al. reported a series of Ir(III) complexes using (4′-substituted-2′-pyridyl)-1,2,4-triazole ancillary ligands. By doping a 10 wt% of iridium complex of 5-(4′-methylpyridine-2′-yl)-3-trifluoromethyl-1,2,4-triazolate ancillary ligand (Ir5) (Figure 33) into a 4,4′-bis(9-carbazolyl)-2,2′-dimethylbiphenyl (CDBP) host, the resultant device showed a blue emission with CIE coordinates of (0.15, 0.18) [225].
Moreover, Lee et al. reported a series of near deep-blue iridium(III) complexes, (TF)2Ir(pic), (TF)2Ir(fptz), (HF)2Ir(pic), and (HF)2Ir(fptz) (Figure 33). They consisted of 2′,4″-difluororphenyl-3-methylpyridine with a trifluoromethyl carbonyl or heptafluoropropyl carbonyl group as the main ligand and a picolinate or a trifluoromethylated triazole as the ancillary ligand. Among them, (TF)2Ir(fptz) achieved an EQE of 7.4% at 100 cd/m2 for a blue emission with CIE coordinates of (0.14, 0.11) [226].
In 2014, Liu et al. reported a host molecule, 3,6-bis(diphenylphosphoryl)-9-(4′-(diphenylphosphoryl) phenyl)-carbazole (TPCz) (Figure 33). The solvent effect was studied by dissolving a mixture of TPCz and FIrpic (emitter) in various polar solvents. They thought that the higher polar halogen-free solvent isopropanol provided good solubility, helping resist the aggregation of FIrpic and hence render a more densely packed emissive layer. The efficiencies of 22 cd/A and 16 lm/W at (0.15, 0.32) were also reported [227].
In 2015, Sun et al. reported a series of 3,3′-bicarbazole (mCP)-functionalized tetraphenylsilane derivatives (SimCPx) (Figure 33). On changing the number of mCP subunits on the central Si atom, four molecules, namely, bis(3,5-di(9H-carbazol-9-yl)phenyl)diphenylsilane (SimCP2), tris(3,5-di(9H-carbazol-9-yl)phenyl)methylsilane (SimCP3-CH3), tris(3,5-di(9H-carbazol-9-yl)phenyl)phenylsilane (SimCP3-Ph), and tetrakis(3,5-di(9H-carbazol-9-yl)phenyl)silane (SimCP4), were synthesized as bipolar hosts. The derivatives possessed a wide bandgap and good solubility. Moreover, a high thermal and morphological stability was beneficial for the formation of amorphous and homogeneous film via wet processing. Among them, SimCP4 showed the best electron-transporting ability and hence, at a low driving voltage of 4.8 eV, achieved an EQEmax of 14%, CEmax of 28 cd/A, and PEmax of 14 lm/W corresponding to a CIExy (0.15, 0.35) by using a FIrpic emitter [228].
In 2016, Byeong et al. reported meta-linked diphenylether-based materials, 9-(3-(3-(9H-carbazol-9-yl)phenoxy)phenyl)-9H-pyrido[2,3-b]indole (CzDPEPI) and 9,90-(oxybis(3,1-phenylene))bis(9H-carbazole) (CzDPECz) (Figure 33). The presence of a pyridoindole moiety assisted in getting a better electron injection and hence a lower driving voltage. The CzDPEPI host possessed a high triplet energy and therefore was suitable for an Ir complex (Ir(dbi)3) emitter. The resultant device showed an EQEmax of 24% and PEmax of 39 lm/W at CIExy (0.19, 0.37) [229].
In 2017, Park et al. reported a series of bipolar host materials, namely, N-(3-(9H-carbazol-9-yl)phenyl)-N-(pyridin-2-yl)pyridin-2-amine (mCPPy), 5-(9H-carbazol-9-yl)-N1,N1,N3,N3-tetra(pyridin2-yl)benzene-1,3-diamine (CPDPy), and N-(3,5-di(9H-carbazol-9-yl)phenyl)-N-(pyridin2-yl)pyridin-2-amine (DCPPy) (Figure 33). The materials possessed a high bandgap (>3.5 eV) and a high triplet energy (>2.9 eV). The device composing CPDPy doped with FIrpic (emitter) showed a highest EQEmax of 22% and PEmax of 24 lm/W at CIE coordinates of (0.16, 0.31) [230].
In 2018, Kim et al. reported an emitter material, fac-tris(2,3-dimethylimidazo[1,2-f]phenanthridinato-k2C,N)iridium(III) (Figure 33), that had a facial (fac) geometry around the iridium atom. A multilayered device composed of 1,3-bis(N-carbazolyl)benzene as a host/hole-transporting layer with a 5 wt% emitter displayed a high EQEmax of 20% and PEmax of 32 lm/W with coordinates of (0.18, 0.34) [89].
In 2020, Yuan et al. reported a host material, 9,9-bis(9-phenyl-9H-carbazol-3-yl)-9H-thioxanthene,10,10-dioxide (BPhCz-ThX) (Figure 33), based on an insulating C(sp3) bridge-linked thioxanthone via a facile Friedel–Crafts-type substitution reaction. The material possessed a high thermal stability, good film-forming ability, a high triplet energy (3.05 eV), and a total yield of 63%. On using a FIrpic emitter, the device showed an EQEmax of 9.6%, CEmax of 21 cd/A, and PEmax of 8.5 lm/W at (0.19, 0.36) [231].

6.2.2. Platinum Complexes

Platinum (Pt) complexes are the second most appreciated organic dopants to realize a highly efficient deep-blue emitter. Pt-based phosphors appear to have the highest potential for EL applications due to their higher triplet quantum yield, relatively short triplet state lifetime, and tunable emission color [211,213,232]. Based on their structural features, these phosphors are classified into three main categories: (i) tetradentate ligands, (ii) tridentate ligands, and (iii) bidentate ligands.
We have described a few platinum complexes based on the above three categories. In 2008, Yang et al. reported a blue device by utilizing a platinum(II) 1,3-difluoro-4,6-di(2-pyridinyl)benzene chloride (Pt-4) (Figure 34) emitter, and the resultant device exhibited CIE coordinates of (0.15, 0.26) with a maximum EQE of 16% [211].
In 2013, Turner et al. synthesized emitters based on cyclometalated tetradentate platinum complexes, a blue PtOO1, and a blue-green PtOO2 emitter (Figure 34). The device using PtOO2 exhibited an EQE of 8.7% with CIE coordinates of (0.21, 0.38) at 1000 cd/m2 [232]. In the same year, Hang et al. reported Pt complexes with tetradentate ligands. These complexes have a conventional cyclometalated fragment bridged with oxygen to an ancillary chelate, such as phenoxyl pyridine (POPy) or carbazolyl pyridine (CbPy). Doping 6 wt% emitters, Pt[pmi-O-POPy] (PtOO7), Pt[pmi-O-CbPy] (PtON7) and Pt[ppz-O-CbPy] (PtON1), into a 2,6-bis(N-carbazolyl)pyridine (26mCPy) host resulted in CIE coordinates of (0.15, 0.10), (0.15, 0.14), and (0.15, 0.13) and an EQE of 0.5, 15, and 21%, respectively, at 1000 cd/m2 [213].
In 2015, Liao et al. reported a series of Pt(II) metal complexes [Pt(Ln)], n = 1–5 based on tetradentate bis(pyridyl azolate) chelates. Among them, L5 (Figure 34) showed a high EQEmax of 15%, CEmax of 36 cd/A, and PEmax of 38 lm/W at CIE coordinates of (0.19, 0.34) [233].
In 2019, Klimes et al. reported a step-wise graded doping of platinum(II) 9-(pyridin-2-yl)-2-(9-(pyridin-2-yl)-9H-carbazol-2-yloxy)-9H-carbazole (PtNON) (Figure 34). The emissive layer was composed of a 20% PtNON:mCBP (10 nm)/10% PtNON:mCBP (4 nm)/2% TBPe:mCBP (2 nm)/10% PtNON:mCBP (4 nm)/6% PtNON:mCBP (10 nm). Consequently, the device showed an EQEmax of 18% and a lifetime of 709 h [LT70] at 1000 cd/m2 with CIExy (0.16, 0.28) [102].

6.2.3. Other Metal Complexes

Due to limited availability and high cost, some research efforts have been made to substitute heavy metal complexes with comparatively low-cost metals such as zinc (Zn), copper (Cu), etc. In 2004, Wang et al. synthesized a Zn(II) metal complex-based phosphorescent blue emitter using 4,4′-diphenyl-6,6′-dimethyl-2,2′-bipyrimidine ligands, [(Zn(pmbp)2(ClO4)2 and Zn(pmbp)Cl2]. However, no efficiencies or color coordinates were reported [234].
In 2008, Xu et al. reported a deep-blue emitting Zn(II) complex, Zn(Lc)2 (Lc− = 2-(1-(6-(9H-carbazol-9-yl)hexyl)-1H-benzo[d]imidazol-2-yl)phenolate) (Figure 35), based on a carbazole-functionalized N^O ligand. The material possessed a Td of 427 °C and an emission peak at 422 nm [235]. In 2013, Lee et al. synthesized two Zn complexes, bis(2-(oxazol-2-yl)phenolate)zinc (ZnOx2) and bis(2-(1-methyl-1h-imidazole-2-yl)phenolate)zinc (ZnIm2) host materials. The blue device based on a 3 wt% FIrpic (emitter) doped in ZnIm2 reached an EQE of 20% at 100 cd/m2 and 18% at 1000 nits [236].
Suitable host materials were presented to aid the emission of phosphorescent-based blue emitters as reported by Chaskar et al [1] and Yook et al [237].

6.3. Thermally Activated Delayed Fluorescence (TADF) Type

High price and rare availability of heavy metals have limited the low-cost and long-term mass production of phosphorescent blue emitters. To resolve these problems, the search for a highly efficient deep-blue fluorescent emitter is still a subject of current interest, as they have a greatest possibility of a longer operational lifetime and a deep-blue emission at low cost with highly sustainable purely organic luminescent molecules [238].
In 2009, Adachi et al. designed and synthesized the first heavy metal-free TADF emitter [239]. Since then, numerous efforts have been made to realize efficient TADF-based blue emitters, especially with a smaller singlet-triplet energy difference ( Δ E S T ) [239,240,241,242,243,244,245]. Presently, aromatic- and metal (such as Li, Al, Mg, Cu, and Zn)-based blue emitters have been found useful in achieving a small Δ E S T .

6.3.1. Aromatic-based

Over the past few years, researchers have extensively designed and synthesized aromatic-based efficient TADF emitters by using abundant elements, such as carbon, hydrogen, oxygen, nitrogen, sulphur, etc. So far, a few molecular design strategies have been reported to achieve smaller Δ E S T (<0.1 eV) values with a high PLQY (~100%). Most importantly, the delocalization of frontier molecular orbitals (highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) energy levels) in a donor–acceptor (D–A) type TADF emitter with a well-separated HOMO and LUMO levels can effectively realize a smaller Δ E S T and a nearly 100% PLQY [246]. Moreover, increasing a steric hindrance between D–A units can also lead to the spatial separation of the frontier molecular orbitals.
In aromatic-based TADF emitters, the spatial overlap between the frontier molecular orbitals can be successfully controlled by choosing a suitable D–A pair. Moreover, it has been found that the donor–acceptor–donor (D–A–D) type emitters are more effective than those of the D–A type in achieving a higher PLQY and a strong spatial separation of HOMO LUMO levels [65,247]. Since 2009, several D–A- and D–A–D-type bipolar emitters have been reported. Among them, a few emitters demonstrated a nearly 100% internal quantum efficiency (IQE) and a 20% EQE in blue devices [246,248,249].
On the basis of D–A pairs, aromatic-based TADF blue emitters can be classified into seven sub-categories (Figure 36), which comprise diphenylamine, acridine, carbazole, and phenoxazines as electron-donor units and diphenylsulfone, thioxanthone, benzophenone, triazine, oxadiazole, triazole, and phthalonitrile as electron-acceptor units (Figure 37).

Diphenylamine and Diphenylsulfone (DPA-DPS) Derivatives

Diphenylamine moiety-containing aromatic compounds are well-established hole-transporting materials in OLED devices. In contrast, diphenylsulfone moiety-comprising materials are well studied as electron transporting materials. These moieties have also been used to design bipolar hosts [250,251,252,253]. In accordance with the molecular design strategy, in 2012, Adachi et al. synthesized a series of bipolar TADF emitters, bis[4-(diphenylamino)phenyl] sulfone (DPA-DPS) and bis{4-[bis(4-tert-butylphenyl)amine]phenyl} sulfone (DTPA-DPS) (Figure 38), using diphenylamine donor and diphenylsulfone acceptor moieties [254]. The DPA-DPS and DTPA-DPS emitters possessed a ΔEST of 0.54 and 0.45 eV and a PLQY of 57 and 65%, respectively. In a doped device with a bis[2-(diphenylphosphino)phenyl]ether oxide (DPEPO) host, the DPA-DPS-based device exhibited an EQEmax of 2.9%, while the DTPA-DPS-based device showed an almost double EQEmax of 5.9%. The higher EQE of the DTPA-DPS emitter is attributed to a low ΔEST, which possessed an effective triplet to singlet-exciton upconversion.

Acridine and Diphenylsulfone (ACR-DPS) Derivatives

Acridine derivatives were utilized as hole-transporting and bipolar host materials in OLEDs [255,256,257,258,259,260,261,262]. Combinations of acridine-based electron-donor and diphenylsulfone-based electron-acceptor units have been widely studied for TADF type emitters. In 2014, Adachi’s group designed and synthesized an acridine donor and diphenylsulfone acceptor moiety-based 9,9-(dimethyl-9,10-dihydroacridin)diphenyl sulfone (DMAC-DPS) (Figure 39) emitter. The emitter possessed a ΔEST of 0.08 eV with an 80% PLQY and a 1.7 to 3 μs TADF decay time. By doping into a DPEPO host, the DMAC-DPS-based device exhibited an EQEmax of 20% for a greenish-blue emission at (0.16, 0.20). The resultant device performance was almost independent of the emitter doping concentration [248]. Moreover, the same group reported an EQEmax of 20% for a sky-blue emission with (0.16, 0.29) by using a non-doped DMAC-DPS emitter [249].
In 2014, Nakanotani et al. achieved an EQEmax of 13% for a bluish-green emission with coordinates (0.17, 0.30) by using a 1 wt% fluorescent emitter, 2,5,8,11-tetra-tert-butylperylene, along with a 15 wt% of TADF emitter, 10-phenyl-10H, 10′H-spiro[acridine-9, 9′-anthracen]-10′-one (ACRSA), in a bis-(2-(diphenylphosphino)phenyl)ether oxide (DPEPO) host [263].
In 2017, Nakao reported three emitters, namely, Ac-46DPPM, Ac-26DPPM, and CzAc-26DPPM (Figure 39), with 2,4,6-triphenylpyrimidine as an acceptor and tri- or penta-substituted acridine as a donor unit. Among them, CzAc-26DPPM displayed an EQE of 15% and CE of 35 cd/A at 100 cd/m2 for a bluish-green emission with coordinates (0.21, 0.38) [264].

Carbazole and Diphenylsulfone (Cz-DPS) Derivatives

Carbazole-based polycylic aromatic hydrocarbon compounds are extensively studied in OLED devices due to their numerous superlative characteristics, such as a strong hole-transporting ability, high thermal stability, large optical band gap, high triplet energy, rational PLQY, and easy functionalization [265,266]. On the other hand, diphenyl sulfone derivatives are robust electron-transporting materials for OLED devices [254].
In 2012, Zhang et al. reported an EQEmax of 10% for a deep-blue emission with CIE coordinates of (0.15, 0.07) by doping a TADF blue emitter, bis[4-(3,6-di-tert-butylcarbazole)phenyl] sulfone (DTC-DPS) (Figure 40), into a DPEPO host. The resultant device realized an EQE over three times higher than that of a diphenylamine-based DPA-DPS and nearly two times higher than that of a tert-butyl diphenylamine-derivative DTPA-DPS. The DTC-DPS exhibiting a high EQE may be attributed to a relatively small ΔEST value of 0.32 eV and a high PLQY of 69% because of a considerably strong electron donating characteristic of carbazole units [254].
In order to further decrease the ΔEST value, in 2014, Wu et al. designed a blue emitter, bis [4-(3,6-dimethoxycarbazole)phenyl]sulfone (DMOC-DPS) (Figure 40), by using a methoxy substituted carbazole with an electron-acceptor sulfone. The emitter realized a ΔEST of 0.21 eV with a 93 μs TADF decay time and a 56% PLQY. In a doped blue device with a DPEPO host, the DMOC-DPS exhibited an EQEmax of 15% with (0.16, 0.16) coordinates [267]. The reason why the DMOC-DPS resulted in an EQE much higher than that of the DTC-DPS may be attributed to a comparatively smaller ΔEST.
Furthermore, in 2016, Lee et al. reported a TADF emitter, TMC-DPS (Figure 40), with diphenyl sulfone (DPS) as an electron acceptor and 3,6-dimethoxycarbazole (DMOC) and 1,3,6,8-Tetramethyl-9H-carbazole (TMC) as electron donor units. The material possessed a low ΔEST of 0.09 eV, much lower than that of DMOC-DPS, due to a large dihedral angle that creates a small spatial overlap between the frontier molecular orbitals [268].

Carbazole and Triazine (Cz-TRZ) Derivatives

Nitrogen-containing heterocyclic compounds, such as triazine and 1,3,4-triazole, have been widely used as electron-accepting units in the molecular scaffold of electron-transporting materials because of their electron-deficient nature. To design a bipolar TADF emitter, strong hole-transporting carbazole moieties have been added onto transporting triazine moieties. Both moieties are extensively utilized to design an efficient bipolar host and blue light-emitting materials.
In 2013, Serevičius et al. reported a 6% EQEmax for a greenish-blue OLED with (0.23, 0.40) by using a 9-(4,6-diphenyl-1,3,5-triazin-2-yl)-9′-phenyl-3,3′-bicarbazole (CzT) (Figure 41) emitter [269]. In 2014, Hirata and co-workers reported a series of bipolar blue light-emitting materials (2a, 2b, and 2c) (Figure 41) by employing indolocarbazole and triazine moieties. The resultant device employing a 2a emitter doped in a DPEPO host realized a highest EQE of 21% for a sky-blue device with (0.19, 0.35) [246].
In 2015, Lee et al. reported a pair of bipolar TADF type blue emitters (DCzTRz and DDCzTRz) (Figure 41) with dicarbazolylbenzene donor and triazine acceptor moieties [270]. Both DCzTRz and DDCzTRz exhibited a ΔEST of 0.25 and 0.27 eV and a solid state PLQY of 43 and 66%, respectively. As both emitters were doped into a DPEPO host, the DCzTRz emitter-based device exhibited a maximum EQE of 18% with (0.15, 0.15), while a 19% EQE with (0.16, 0.22) was reported for the DDCzTRz-based device. Furthermore, owing to high thermal and electrical stability, the DDCzTRz-based device also demonstrated an operational lifetime (T80 = 52 h at an initial luminance of 500 cd/m2).
In 2020, Su et al. reported a 3,3′-bicarbazole/triphenyltriazine-derivative (BCz-TRZ) (Figure 41) emitter with a ΔEST of 0.37 eV by controlling the length of donor and acceptor moieties and increasing the D–A distance. The BCz-TRZ-based device exhibited a lifetime (LT50) of 495 h and an EQEmax of 20% at an emission wavelength of 486 nm [114].

Carbazole and Phthalonitrile (Cz-PN) Derivatives

Phthalonitrile-derivative compounds have attracted considerable attention because of their strong electron-withdrawing nature. Phthalonitrile-functionalized aromatic compounds are widely used as electron-transporting, charge-generation, and light-emitting materials in OLEDs. To achieve bipolarity in the molecular design of phthalonitrile-based aromatic compounds, such as bipolar hosts and TADF emitters, strong electron donor carbazole units are appended onto the strong electron donor phthalonitrile unit [271,272].
In 2012, Uoyama et al. designed a series of carbazole and phthalonitrile materials. Among them, 4,5-di(9H-carbazol-9-yl) phthalonitrile (CzTPN) (Figure 42) realized a sky-blue emission. The CzTPN exhibited a ΔEST of 0.29 eV with a 47% PLQY. In a doped device with a CBP host, the CzTPN emitter exhibited a maximum EQE of 8% for a sky-blue emission with an emission wavelength at 473 nm [271]. Furthermore, in 2013, Masui et al. achieved a maximum EQE of 13% with (0.17, 0.30) by doping the CzTPN emitter into an mCP host [64].
In 2015, Lee et al. reported a maximum EQE of 16% for a blue device with (0.17, 0.19) by doping a 15 wt% 4,6-di(9H-carbazol-9-yl)isophthalonitrile (DCzIPN) (Figure 42) in a mCP host [273]. In the same year, Lee et al. designed and synthesized two blue emitters, 4,5-bis(benzofuro[3,2-c]carbazol-5-yl)phthalonitrile (BTCz-2CN) and 4,5-bis(benzo[4,5]thieno [3,2-c]carbazol-5-yl)phthalonitrile (BFCz-2CN) (Figure 42), by employing benzofuro- and benzothieno-functionalized carbazole units with a phthalonitrile unit. BTCz-2CN and BFCz-2CN exhibited a ΔEST value of 0.13 and 0.17 eV with a 95 and 94% PLQY, respectively. The BTCz-2CN showed a decay lifetime of 2.6 μs, which is larger than that of the BFCz-2CN counterpart (1.98 μs). The short decay time of the BTCz-2CN and BFCz-2CN may be attributed to their small ΔEST. Moreover, BTCz-2CN and BFCz-2CN also showed a high Tg of 204 and 219 °C and Td of 397 and 434 °C, respectively. In doped devices with an mCP host, the BTCz-2CN- and BFCz-2CN-based devices resulted in a maximum EQE of 12.1 and 11.8% with a similar EL emission peaking at 486 nm [274].
In 2020, Yi et al. reported three molecules, namely, 4CzIPN-CF3, 3CzIPN-H-CF3, and 3CzIPN-CF3 (Figure 42). They were synthesized by lowering the electron-withdrawing ability of the acceptor group or by reducing the number of carbazole donors of the base molecule, 1,2,3,5-tetrakis(carbazol-9-yl)-4,6-dicyanobenzene (4CzIPN). Among them, the 4CzIPN-CF3- and 3CzIPN-H-CF3-based devices with a SimCP:oCF3-T2T co-host system displayed a blue-shifted emission with coordinates (0.21, 0.41) and (0.16, 0.19), respectively, as compared with a 4CzIPN emission (0.22, 0.48). The devices showed an EQEmax of 23 and 17%, and 21 and 15% at 1000 nits, respectively [275].

Acridine and Phenoxaborin (Cz-PXB) Derivatives

In 2015, Numata et al. synthesized a series of D–A-type emitters (1-4) (Figure 43) with acridine derivatives as donor units and a 10H-phenoxaborin-type acceptor [276]. The resultant maximum EQE ranged from 20 to 13%, while chromaticity coordinates for a sky-blue emission ranged from (0.14, 0.23) to deep-blue (0.15, 0.08) by doping a 20 wt% of 1–4 emitters into an 8-bis(diphenylphosphine oxide) dibenzofuran (PPF) host. Moreover, the resulting sky-blue device also demonstrated a maximum EQE of 22% when the doping concentration of emitter 1 was further increased to 50 wt% in the same PPF host. The high efficiencies may be attributed to three key factors. First, the ACR-PXB series emitters possessed a higher PLQY value ranging from 56 to 100%, a shorter decay time ranging from 1.6 to 4.02 μs, and a smaller ΔEST value ranging from 0 to 0.16 eV. Second, rational bipolar characteristics of the emitters may help to achieve an effective charge carrier injection balance in the emissive layer. Third, the high triplet energy (3.21 eV)-host material restricted any backward energy transfer and facilitated an effective host-to-guest energy transfer. As reported, the device performance was greatly influenced by the ΔEST and PLQY of the emitters. The PLQY increased from 56 to 100% for emitter 1 as ΔEST decreased from 0.16 to 0 eV [276].

Phenoxazines and Triazole (PXZ-TAZ) Derivatives

Bipolar compounds based on phenoxazine electron-donor and triazole electron-acceptor moieties have been widely investigated to design the bipolar host materials for OLEDs [refs]. In 2013, Adachi et al. reported two TADF blue emitters, 10-(4-(4,5-diphenyl-4H-1,2,4-triazol-3-yl)phenyl)-10H-phenoxazine (PXZ-TAZ) and 10,10′-((4-phenyl-4H-1,2,4-triazole-3,5-diyl)bis(4,1-phenylene))-2-bis(10H-phenoxazine) (2PXZ-TAZ) (Figure 44), by using 1,2,4-triazole core-based emitters with a 10H-phenoxazine group. The PXZ-TAZ and 2PXZ-TAZ emitters exhibited a PLQY of 13 and 15% in neat films. The PLQY was greatly enhanced from 15 to 52% as a 6 wt% 2PXZ-TAZ emitter was doped into a DPEPO host. The 2PXZ-TAZ also showed a maximum EQE of 6.4% with (0.16, 0.15) and an EL emission peaking at 456 nm [65].

Acridine and Triazine (ACR-TRZ) Derivatives

As described in prior sections, both phenoxazine and triazine moieties have been extensively utilized to design bipolar hosts and TADF emitters for OLEDs. In 2015, Tsai et al. designed and synthesized a DMAC-TRZ (Figure 45) emitter based on 9,9-dimethyl-9,10-dihydroacridine as a donor unit and 2,4,6-triphenyl-1,3,5-triazine as an acceptor unit. The DMAC-TRZ emitter showed a ΔEST of 0.05 eV and a delay lifetime of 3.6 μs with an 83% PLQY. Furthermore, the ΔEST and the delay lifetime were decreased to 0.046 eV and 1.9 μs, respectively, while the PLQY was increased to 90% as the DMAC-TRZ emitter was doped into a 9-(3-(9H-carbazol-9-yl)phenyl)-9H-carbazole-3-carbonitrile (mCPCN) host. At 100 nits, the resultant non-doped DMAC-TRZ-based device exhibited a maximum EQE of 20% with a bluish-green emission, while a doped emissive layer (at 8 wt% DMAC-TRZ and mCPCN)-based device showed a maximum EQE of 27% with a slightly blue shifted emission. These devices also exhibited relatively high power efficacy, at 46 and 66 lm/W, and current efficacy of 61 and 67 cd/A at 100 cd/m2, respectively, for non-doped and doped emissive layer-based devices [277].
In 2020, Li et al. reported an emitter, DspiroAc-TRZ (Figure 45), with a tight and regular intermolecular packing composed of X- or L-shaped J-aggregates. PLQYs of 79 and 84% were obtained for the DspiroAc-TRZ crystal and nondoped film, respectively, due to the synergistic effect of the intermolecular packing mode and monomolecular structure. Thus, a very high EQEmax of 26% and CEmax of 56 cd/A were realized for a sky-blue emission with (0.17, 0.35) by using a host-free emissive layer [278].

6.3.2. Metal-Based TADF Emitters

Metal-based TADF blue emitters had also been reported to realize a very small ΔEST, which can effectively up-convert the entire fraction of triplet excitons (75%) into singlet excitons via ultra-fast reverse intersystem crossing (RISC) [279,280,281,282]. This prudent combination of ultra-fast RISC and a very small ΔEST would result in a 100% delayed fluorescence yield, [282,283,284,285] and significantly diminish the efficiency roll-off, especially at high luminance or current density [286].
The metal-based TADF emitters also exhibited a higher thermal stability than that of pure aromatic-based counterparts [286,287,288,289]. By virtue of these advantages, a considerable number of blue TADF emitters based on non-precious and abundant metals, such as lithium (Li), magnesium (Mg), copper (Cu), zinc (Zn), and tin (Sn), were reported (Figure 46) [290]. However, only a few of them were investigated in OLEDs [239,291,292,293,294,295,296,297].

Cu Complexes

Copper (Cu) is one of the highly abundant coinage metals which has been investigated in the design of high-electroluminescence (EL) active complexes for low-cost OLEDs. In order to realize a sustainable way out, Ma et al. designed and synthesized the first Cu complex-based emitter in 1999 [298,299]. Later, many Cu complex-based phosphorescent emitters were reported because of their ability to realize a high rate of intersystem crossing (ISC) (S1→T1) and triplet radiative decay (T1→S0) [282,300].
Due to a small ΔEST, a few Cu complexes showed effective RISC (T1→S1) and singlet radiative decay (S1→S0) rates [294,301]. These complexes exhibited a small ΔEST and a fast RISC because of low-lying metal-ligand charge transfer (MLCT) excited states, which played a crucial role in realizing an effective triplet up-conversion [291,292,293,294,295,296,297].
Over the past few years, several Cu-based TADF emitters have been utilized in OLEDs [302,303,304,305,306,307,308,309,310]. Among them, only a few emitters demonstrated a blue emission. In 2011, Yersin et al. designed and synthesized a series of Cu (I) complexes (Cu-4 to Cu-6) (Figure 47) [298]. The ΔEST and PLQY of these compounds ranged from 0.09 to 0.16 eV and 45 to 90%, respectively, in solid state films. The chromaticity coordinates of the Cu-4 to Cu-6 emitters ranged from (0.14, 0.11) to (0.16, 0.22). In the same year, Leitl et al. reported a design strategy to realize a strong TADF emission in another Cu complex [311]. They observed that a small ΔEST value can be obtained by selecting heteroleptic ligands with a low torsion angle. A Cu complex (Cu-7) with two N-heterocyclic ligands, 1,3-bis(2,6-diisopropylphenyl)imidazol-2-ylidene and di(2-pyridyl)dimethylborate), showed a blue emission peaking at 475 nm. The Cu-7 exhibited a solid state PLQY of 70% and a ΔEST of 0.09 eV, resulting in a TADF decay fraction of 62% with a weak phosphorescence decay fraction of 38% [311].
In 2013, Chen et al. reported a series of Cu complex (Cu-1 to Cu-3) (Figure 47)-based emitters [312]. These complexes showed a constant ΔEST of 0.17 ± 0.01 eV with a PLQY ranging from 56 to 87%. Via wet-processing, the resultant maximum EQE ranged from 8.5 to 1.5%, with chromaticity coordinates ranging from greenish-blue (0.24, 0.36) to sky-blue (0.20, 0.30) as a 20 wt% of the emitters was doped into a hole-transporting host 26mCPy. The Cu-3 complex-based device also exhibited a maximum current efficiency of 24 cd/A.
In 2014, Yersin et al. reported a series of Cu complex-based emitters. Their structures were similar to the phosphorescent emitters reported earlier [311,313].
In 2015, Kang et al. reported a series of binuclear Cu halide complexes (Cu-8–Cu-10) (Figure 47) with different halogen molecules (Cl, Br, and I) [291]. These complexes exhibited a ΔEST value from 0.05 to 0.04 eV, a PLQY from 42 to 95%, and a decay time ranging from 4.9 to 5.9 μs with an emission wavelength ranging from blue (482 nm) to greenish blue (490 nm).

Zn Complexes

Over the past few years, several Zn complexes have been employed as EL-active emitters, hosts, and electron transporting materials in OLEDs [235,314,315,316,317,318,319,320,321,322]. However, Zn complexes exhibited a poorer device performance than that of Cu complex-based counterparts while displaying a higher stability against atmospheric moisture and air. Therefore, numerous efforts have been devoted to design Zn complex-based TADF emitters [296,323,324].
In order to devise a highly stable metal complex-based emitter, in 2015, Adachi’s group reported two different Zn complexes (Zn-1 and Zn-2) (Figure 48) with blue emission [286]. The Zn-1 and Zn-2 complexes demonstrated a PLQY of 78 and 58% and a ΔEST of 0.06 and 0.18 eV, respectively. In addition, the Zn-1 emitter also realized a low TADF decay time of 38 μs. In an mCBP host-based doped device, the Zn-1 complex showed a maximum EQE of 20% with a greenish-blue emission.
Moreover, Cheng et al. designed and synthesized a series of tetranuclear zinc(II) complexes of substituted 7-azaindoles. The decay time and PLQY of these complexes ranged from 0.015 to 0.008 μs and 96 to 71%, respectively, as a 5 wt% emitter was doped into a PMMA host. The PLQY values of these compounds were measured in solid state film. The resultant maximum EQE ranged from 5.6 to 3.3%, with chromaticity coordinates ranging from (0.16, 0.19) to (0.18, 0.26) as an 8 wt% of the emitters was doped into a mixture of PVK and OXD-7 hosts [325].

Other Metal Complexes

In 2015, Adachi et al. studied other metal (Li, Mg, Al, and Sn) complexes for OLED device applications [286]. The lithium (Li) complex-based TADF emitter (Li-1) exhibited a ΔEST of 0.08 eV with a total PLQY of 70%. In a doped device with an mCBP host, the Li-1 (Figure 49) realized a maximum EQE of 13% with a greenish-blue emission. Moreover, a magnesium-based (Mg-1) (Figure 49) complex achieved a ΔEST of 0.08 eV and a PLQY of 71%, including a 58% delayed quantum yield. In a device doped with an mCBP host, the Mg-1 complex-based device exhibited a maximum EQE of 17% for a greenish-blue emission.

7. Challenges and Approaches

As seen, high efficiency and long lifetime had been reported for devices of light-blue emission, i.e., CIEy > 0.1, using phosphorescent and TADF-based emitters. Even still, a host-free device displayed a much higher efficiency (EQEmax: 25%) for a greenish-blue emission with (0.16, 0.24) coordinates [109]. Moreover, a very high EQEmax of 33% was reported for two different devices, one with a greenish-blue emission at CIExy (0.16, 0.22) [107] and the other with a bluish-green emission at (0.17, 0.33) [106]. The longest lifetime of 100,000 h (LT50) at 200 nits was reported for a bluish-green emission with a CIEy of 0.38 [97]. However, obtainina high efficiency and a long lifetime in the deep-blue region (CIEy < 0.1) remains challenging.

7.1. Challenges for Deep-Blue Emission

Researchers have focused on designing and synthesizing suitable emitters for both blue and deep-blue OLEDs. Employment of various ligands/subunits in a molecule helps to achieve a comparatively high efficiency and long lifetime for blue emission. However, only a few displayed a moderate efficiency or lifetime in the deep-blue region.

7.1.1. High-Efficiency Deep-Blue Emission

As discussed earlier, phosphorescent and TADF-based devices displayed EQE > 20% for a deep-blue emission. For example, an EQE of 25% with a CIEy of 0.08, but at a luminance below 100 nits, was reported by using a phosphorescent emitter, TSP01 [86]. Moreover, an EQE of 20% with a CIEy of 0.03, but also at a luminance below 100 nits, was reported by using a host-free TADF emitter, NTN-PCZ [104]. However, Lim et al. reported an EQEmax of 28% with a CIEy of 0.09 crossing the upper theoretical limit for TADF emitters by using a TDBA-SAF TADF emitter. Most importantly, an EQE of 24% was reported at 100 nits and 18% at 1000 nits [105]. Realizing a high EQE with a CIEy much less than 0.1 at applicable high luminance remains challenging.

7.1.2. Long-Lifetime Deep-Blue Emission

A deep-blue fluorescent device with a CIEy of 0.09 was created by Takita et al. with a lifetime of 125 h (LT90) at 1000 nits [72]. Before 2021, no phosphorescent or TADF-based devices had displayed any lifetime data in the deep-blue region. In 2021, there were two articles reporting deep-blue devices with a CIEy of 0.09. Nam et al. reported a deep-blue device with a CIEy of 0.09 and a lifetime of 254 h (LT50) at 1000 nits by using a phosphorescent sensitizer with a fluorescent emitter [326]. Jeon et al. reported another deep-blue device with a CIEy of 0.09 and a lifetime of 151 ± 3 h (LT50) at 1000 nits by using a TADF emitter [327]. However, achieving a long lifetime is still required.

7.1.3. High Efficiency and Long Lifetime

As shown above, the deep-blue fluorescent device reported by Takita et al. also displayed an EQEmax of 12%, beyond the theoretical limit for fluorescent emitters [72]. Since then, no devices were reported with a high efficiency and long lifetime for a deep-blue emission until 2020. In 2021, Jeon et al. reported a deep-blue device with a CIEy of 0.09, displaying a higher EQEmax of 34% and a lifetime of 151 ± 3 h (LT50) at 1000 nits by using a TADF emitter [327]. However, realizing a deep-blue device with a long lifetime along with high efficiency is still essential.

7.2. Approaches for Generating Deep-Blue Emission

In the past decades, a few approaches were reported for realizing a deep-blue emission in OLEDs. They include but are not limited to the following: (i) reducing dopant concentration in the emissive layer, (ii) using a polarity-matching host, (iii) rational emitter designs, (iv) thermally activated delayed fluorescence, (v) hybrid local charge transfer, and (vi) triplet–triplet annihilation.

7.2.1. Reducing Dopant Concentration in the Emissive Layer

Reports have confirmed that an undesirable, red-shifted emission, which is induced due to self-aggregation of emitter molecules, can be greatly suppressed by dispersing into a host material [51]. Unlike rational molecular design, the self-aggregation effect in the emitter molecules can be effectively diminished by decreasing their concentration in the host [168]. Jou et al. reported a blue emission peaking at 448 nm when a pure cyano fluorene acetylene-conjugate C3FLA-2 was employed as an emissive layer. The resultant blue emission peak exhibited a strong hypsochromic-shift from 448 to 428 nm when a 7 wt% C3FLA-2 emitter was doped into a CBP host [51]. The employed host can be utilized to disperse the dopant to prevent the bathochromic shift caused by emitter aggregation. The dispersion is more effective when the dopant is more diluted [328], as indicated by the main emission peak, which was blue-shifted from 428 to 416 nm as the C3FLA-2 concentration was decreased from 7 to 1 wt% [51]. The same group reported triphenylamine and carbazole-based emitters. Compared with the devices based on the neat films of these emitters, the EL shifted towards blue as they were dispersed in a CBP host. Deep-blue emission with high efficiency was obtained with no observable red shift for doping concentration up to 9 wt% [168].
Chen et al. reported three blue emitters, TPIBCz, TPINCz, and TPIBNCz, with an EL emission ranging between 435 and 448 nm for undoped devices. When a 30 wt% of these emitters was doped into a CBP host, a hypsochromic shift was observed with emission ranging between 428 and 440 nm. Furthermore, on reducing the dopant concentration to 10 wt%, their main emission peaks further blue-shifted to between 424 and 436 nm [329].

7.2.2. Employing a Polarity Matching Host

Researchers found that a deep blue emission can also be obtained from sky-blue and/or blue emitters by employing a polarity matching host for the desirable emitters. Jou et al. achieved a deep-blue emission as a sky-blue emitter 1-((9,9-diethyl-9H-fluoren-2-yl)ethynyl)pyrene (PA-1) was doped into a polarity matching host CBP. The emitter PA-1 showed a relatively low polarity of 0.8 D, whilst CBP showed almost no polarity with a dipole moment of 0.0 D. Therefore, the high similarity between the low polarities of CBP and PA-1 could enhance emitter solubility. The resultant device showed a CIEy of 0.06 [162].
Chen et al. reported three blue emitters, TPIBCz, TPINCz, and TPIBNCz, with an EL emission at 435, 448, and 436 nm and a low polarity of 8.2, 13, and 9.6 D, respectively. When doped into a non-polar host CBP, the three emitters exhibited a blue-shifted emission at 428, 436, and 424 nm, respectively. As said, the host decreased the overall polarity of the medium and hence resulted in a blue shift [329].

7.2.3. Rational Emitter Designs

We have reviewed herein two rational designs for synthesizing deep-blue emitters. The two basic principles are to minimize the undesired intramolecular charge transfer (ICT) and shorten π -conjugation or resonance length, which may cause a red-shift in the emission under electrical excitation. The extent of the ICT effect can be minimized with the introduction of a weak D–A pair [330].
Incorporating carbazole moieties in a molecule is another way to achieve a deeper-blue emission due to their properties such as good hole-transporting, large bandgap, and weaker π -donating ability [331,332,333,334]. Jou’s group reported a deep-blue fluorescent emitter, JV234, and the composing device displayed a CIEy of 0.06 [335]. The same group reported another deep-blue fluorescent emitter, JV55, and the device showed an EQEmax of 6.5% with a CIEy of 0.06 [200].
In addition to the carbazole moieties, incorporating oxadiazole as electron-withdrawing moiety builds a bipolar molecule. The carbazole moiety helps reduce the ICT effect and blue shifts the emission. Wang et al. reported such a molecule, namely, Oxd-bCz, and the composing device showed a CIEy of 0.07 [203]. Meanwhile, Huang et al. reported that a benzofuro [2,3-b]pyrazine unit (BFPz) as a withdrawing moiety along with a carbazole moiety can help achieve a CIEy of 0.05 [336].
Furthermore, Pal et al. reported two homoleptic charge-neutral meridional (mer-) complexes (1 and 2). These complexes contain a CF3 group as an electron-withdrawing group on the cyclometalating aryl group. The only difference in their structures was that in complex 1, the CF3 substitution was in a more electron-withdrawing para position with respect to the C-Ir bond. Both complexes showed a deep-blue emission. Complex 2 as an emission layer showed a CIEy of 0.05 and a high PLQY of 72% [337]. Wang et al. reported that a limited π -conjugation length was required for achieving a deep-blue emission. They hence used a spiro-diphenylsulfone-based TADF emitter that had a short π -conjugation length to acheive a deep-blue emission with a CIEy of 0.04 for an inverted OLED [338].
Such a design strategy thus offers a formulated platform to generate functional luminescent materials with efficient deep blue emissions. These rational design features are necessary, especially for devices with a non-doped emissive layer.

7.2.4. Thermally Activated Delayed Fluorescence

Weissenseel et al. reported two deep-blue emitters composed of two donor and two acceptor units. The central core consisted of two acridine units linked with a spiro-bridge. The perpendicular conformation between the donor and acceptor moieties reduced the HOMO LUMO gap which resulted in a small singlet-triplet gap. The device with a 10 wt% SBABz4 doped in a DPEPO host showed a CIEy of 0.09 [339].
Chan et al. reported four blue emitters, DCzBN1-4, based on 3,6-substituted carbazole donor and benzonitrile acceptor units. The ΔEST was tuned by modifying the substituent on the donor moieties. The emitters exhibited an emission ranging between 404 and 470 nm. Among them, the device based on DCzBN3 that had a di-tert-butylcarbazole donor unit displayed an EQEmax of 10% and a CIEy of 0.06 [340].
Wang et al. reported a spiro-diphenylsulphone-based TADF emitter that emitted in a deep-blue region. The non-doped device showed a low efficiency roll-off and achieved an EQEmax of 5.0 and 2.1% at 1000 nits with a CIEy of 0.04 [338].
Liang et al. reported four trifluoromethyl-based compounds, namely, TN3T-PCZ, TN4T-PCZ, DTF-PCZ, and NTN-PCZ. Among them, the device based on the TN4T-PCZ emitter displayed an EQEmax of 20% and a CIEy of 0.03 [104].
Lim et al. reported a deep-blue emitter, TDBA-SAF, based on a spiro-acridine fluorene donor and an oxygen-bridged boron acceptor unit. The composed device displayed an EQEmax of 28% and a CIEy of 0.09 [105].
Luo et al. reported a UV-emissive TADF compound, CZ-MPS, by interlocking a diphenylsulfone acceptor unit with a dimethylmethylene bridge. The composed device exhibited an EL emission at 389 nm and an EQEmax of 9.3% [341].
Gudeika et al. reported a deep-blue TADF compound, di-tert-butyl-dimethylacridinyl disubstituted oxygafluorene, dispersed in an inert polymer under a nitrogen atmosphere. The composed device displayed an EQEmax of 9.9% with a CIEy of 0.08 [342].
Kim et al. reported three color tunable emitters, 9′-(2,12-di-tert-butyl-5,9-dioxa-13b-boranaphtho[3,2,1-de]anthracen-7-yl)-9′H-9,3′:6′,9″-tercarbazole (TB-3Cz), 9′-(2,12-di-tert-butyl-5,9-dioxa-13b-boranaphtho[3,2,1-de]anthracen-7-yl)-9,9″-diphenyl-9H,9′H,9″H-3,3′:6′,3″-tercarbazole (TB-P3Cz), and 9-(2,12-di-tert-butyl-5,9-dioxa-13b-boranaphtho[3,2,1-de]anthracen-7-yl)-N3,N3,N6,N6-tetraphenyl-9H-carbazole-3,6-diamine (TB-DACz). The devices based on TP-3Cz and TB-P3Cz emitters exhibited a deep-blue emission with a CIEy of 0.07 and 0.08, respectively. An EQEmax of 9.9% for TP-3Cz and 6.0% for TB-P3Cz were also observed [108].

7.2.5. Hybrid Local Charge Transfer

Yang et al. reported a deep-blue HLCT luminogen Cz-TPB, based on triphenylbenzene. The material possessed an exciton-utilization efficiency (EUE) of 86%. The non-doped device based on this emitter displayed an EQEmax of 4.3% and a CIEy of 0.09. As reported by the authors, the additional triplet excitons were effectively utilized due to the small ΔES1-T2 of the emitter [343].
Fu et al. reported an HLCT emitter based on a pyrene structure. The composed device displayed an EUE of nearly 100%, an EQE of 11%, and a CIEy of 0.06 [344].
Yu et al. reported three emitters, TPA-1PI, TPA-2PI, and TPA-3PI, based on a phenanthro[9,10-d]imidazole (PI) chromophore. All the three emitters displayed HLCT characteristics. Among them, the device based on TPA-2PI displayed a CIEy of 0.08 [345].
Usta et al. reported a deep-blue HLCT emitter, 1,4-bis((2-cyanophenyl)ethynyl)-2,5-bis(2-ethylhexyloxy)benzene (2EHO–CNPE), based on an oligo(p-phenyleneethynylene) small molecule. The material possessed a wide bandgap of 2.9 eV. The composed device displayed an EQEmax of 7.1% and a CIEy of 0.09 [346].
Chen et al. reported deep-blue HLCT emitters, TPINCz and TPIBNCz, based on naphthyl group(s) as a weak n-type π space in a donor–π–acceptor (D-π-A) system. The device based on TPINCz displayed an EQEmax of 6.6% with a CIEy of 0.05 [329].
Most recently, Zhang et al. reported an HLCT compound, 2TA-PPI, by grafting triphenylamine moieties at -C5 and -C10 positions of a 1,2-diphenyl-1H-phenanthro[9,10-d]imidazole (PPI) unit. The composed device displayed an EQEmax of 11% and a CIEy of 0.09 [347].

7.2.6. Triplet–Triplet Annihilation

As postulated/reported by Kondakov et al., TTA might occur when two excited electrons in the triplet state interact with each other, coupling with the occurrence of an up-conversion mechanism [348]. With that, one low-level and one high-level triplet-exciton are produced. The low-level triplet exciton returns to the ground state (S0) of the emitter while the other is excited to the higher-level singlet state, S2. After fast intermolecular relaxation (S2 − S1), the S1 state is re-populated, which is beneficial for the high efficiency of fluorescent emitters (Figure 50) [349].
Hu et al. reported three compounds, 1,4-bis(10-phenylanthracene-9-yl)benzene (BD1), 1-(10-phenylanthracen-9-yl)-4-(10-(4-cyanophenyl)anthracen-9-yl)-benzene (BD2), and 1-(10-(4-methoxyphenyl)anthracen-9-yl)-4-(10-(4-cyanophenyl)anthracen-9-yl)benzene (BD3). Among them, the doped device based on a BD3 emitter and CBP host displayed an EQEmax of 12% and a CIEy of 0.06. The enhanced EQE is attributed to the occurrence of the TTA mechanism [350].
Chen et al. reported an approach to developing TTA materials for high-efficiency fluorescence via a styrylpyrene core and an electron-donating group. The device with a PCzSP emitter doped in CBP host displayed a 9.1% EQEmax and a CIEy of 0.09 [351].
Zeng et al. reported two quaterphenyl derivatives with a PLQY of nearly 80%. The nondoped device based on a 4PS emitter displayed an EQEmax of 5.9% and CIEy of 0.06. The doped device displayed an EQEmax of 6.6% corresponding to a CIEy of 0.05. As said by the authors, the TTA enabled the slow roll-off in the device [352].
Peng et al. reported dianthracenylphenylene-based emitters. The non-doped devices based on these emitters displayed an EQEmax of 4.6 to 5.9% corresponding to a CIEy from 0.07 to 0.08. They have attributed the good device performance to TTA occurrence [353].
Recently, Wu et al. reported two anthracene-based molecules, 4-(10-(3-(9H-carbazol-9-yl)phenyl)-2,6-di-tert-butylanthracen-9-yl)benzonitrile (mCz-TAn-CN) and 4-(2,6-di-tert-butyl-10-(3,5-di(9H-carbazol-9-yl)phenyl)anthracen-9-yl)benzonitrile (m2Cz-TAn-CN). Among them, the doped device based on the m2Cz-TAn-CN emitter showed an EQEmax of 7.3% and CIEy of 0.09. They also attributed the efficiency improvement to TTA occurrence [354].

8. Cautions

Although high-efficiency and long-lifetime deep blue emitters have some advantages, such as reducing the pixel size/number of pixels in a display or fabricating a high-quality light for lighting, they might have some drawbacks if used inappropriately. Intense blue light tends to disrupt ecosystems, pollute night skies, suppress melatonin secretion, discolor artifacts, and cause numerous eye diseases as well as cancers.
Professor Jou, also the corresponding author of this review article, from National Tsing Hua University has recommended several precautions regarding how to avoid blue-light hazards and safeguard the health of eyes and physiologies in the popular science speeches he has delivered over 100 times in the past 10 years and the book entitled “Embracing Darkness” [355]. He has mentioned five important points to be kept in mind and followed seriously. These are:
(i)
Turn off or dim the light after dusk or at least 2 h before going to bed.
(ii)
Change indoor illumination to a blue light-free candlelight style lighting for nighttime.
(iii)
Avoid the use of 3C products before going to bed. Audio activities are preferred over video ones.
(iv)
Embrace darkness. The bedroom should be completely dark during sleep.
(v)
Consult physicians for melatonin supplements for sleep improvement, especially for older persons.
Precautions are always better than cures. Now is the right time to develop a good habit of “staying in darkness at night”.

9. Conclusions

In summary, this review depicts the global revenue and forecast for lighting and display technologies. Since 2018, OLED display revenues have surpassed those of LCD displays, proving which truly is a disruptive display technology. However, OLED lighting is not yet successful as compared against LED lighting, although its revenue is rising gradually. Until 2023, an increment of 41%, i.e., a forecasted amount of 99 billion, is expected annually.
Overall, the number of papers published and patents filed for blue OLEDs have increased significantly over the past years, especially those based on Gen-3 materials, confirming the need for blue OLEDs. It is also surprising to see that publications based on Gen-1 materials rose significantly in 2021, indicating the importance of deep-blue emitters. Some progress was seen in the development of deep-blue OLEDs with high efficiency and/or long lifetime, increasing the possibility of mass production of displays with a higher resolution that are also more energy-saving and reliable.
Having high efficiency along with a long lifetime still remains challenging, although there are several approaches available for realizing deep-blue emission.
Furthermore, while the marked progress on deep-blue OLEDs is delightful, it is strongly suggested that we keep one eye on the plausible negative impacts that might result from the same, and embracing darkness is always essential after dusk to safeguard human health and ecosystems.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13182521/s1.

Funding

This work was partly supported by the project funded by the Research Council of Lithuania (Grant No. S-MIP-22-84) and Ministry of Science and Technology (MOST), Taiwan (Grant No. 109-2923-M-007-003-MY3).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chaskar, A.; Chen, H.-F.; Wong, K.-T.; Chaskar, A.; Chen, H.-F.K.; Wong, K.-T. Bipolar Host Materials: A Chemical Approach for Highly Efficient Electrophosphorescent Devices. Adv. Mater. 2011, 23, 3876–3895. [Google Scholar] [CrossRef] [PubMed]
  2. Han, C.; Zhao, Y.; Xu, H.; Chen, J.; Deng, Z.; Ma, D.; Li, Q.; Yan, P. A Simple Phosphine-Oxide Host with a Multi-Insulating Structure: High Triplet Energy Level for Efficient Blue Electrophosphorescence. Chem.-A Eur. J. 2011, 17, 5800–5803. [Google Scholar] [CrossRef] [PubMed]
  3. Chien, C.H.; Chen, C.K.; Hsu, F.M.; Shu, C.F.; Chou, P.T.; Lai, C.H. Multifunctional Deep-Blue Emitter Comprising an Anthracene Core and Terminal Triphenylphosphine Oxide Groups. Adv. Funct. Mater. 2009, 19, 560–566. [Google Scholar] [CrossRef]
  4. Tao, Y.; Wang, Q.; Yang, C.; Zhong, C.; Qin, J.; Ma, D. Multifunctional Triphenylamine/Oxadiazole Hybrid as Host and Exciton-Blocking Material: High Efficiency Green Phosphorescent OLEDs Using Easily Available and Common Materials. Adv. Funct. Mater. 2010, 20, 2923–2929. [Google Scholar] [CrossRef]
  5. Swayamprabha, S.S.; Dubey, D.K.; Shahnawaz; Yadav, R.A.K.; Nagar, M.R.; Sharma, A.; Tung, F.-C.; Jou, J.-H. Approaches for Long Lifetime Organic Light Emitting Diodes. Adv. Sci. 2021, 8, 2002254. [Google Scholar] [CrossRef] [PubMed]
  6. Girase, J.D.; Tagare, J.; Shahnawaz; Nagar, M.R.; Siddiqui, I.; Jou, J.H.; Patel, S.; Vaidyanathan, S. Deep-Blue Emitters (CIEy~0.07) Based on Phenanthroimidazole: Remarkable Substitution Effects at the N1 Position of Imidazole on the Excited States and Electroluminescence Properties. Dye. Pigment. 2021, 196, 109791. [Google Scholar] [CrossRef]
  7. Devesing Girase, J.; Rani Nayak, S.; Tagare, J.; Shahnawaz; Ram Nagar, M.; Jou, J.H.; Vaidyanathan, S. Solution-Processed Deep-Blue (Y~0.06) Fluorophores Based on Triphenylamine-Imidazole (Donor-Acceptor) for OLEDs: Computational and Experimental Exploration. J. Inf. Disp. 2022, 23, 53–67. [Google Scholar] [CrossRef]
  8. Anupriya; Thomas, K.R.J.; Nagar, M.R.; Shahnawaz; Jou, J.H. Phenanthroimidazole Substituted Imidazo[1,2-a]Pyridine Derivatives for Deep-Blue Electroluminescence with CIEy~0.08. J. Photochem. Photobiol. A Chem. 2022, 423, 113600. [Google Scholar] [CrossRef]
  9. Jou, J.-H. Introduction to OLED; Gau Lih Book 10, Ltd.: Chiayi, Taiwan, 2015; ISBN 978-986-378-031-1. [Google Scholar]
  10. Shahnawaz; Siddiqui, I.; Nagar, M.R.; Choudhury, A.; Lin, J.T.; Blazevicius, D.; Krucaite, G.; Grigalevicius, S.; Jou, J.H. Highly Efficient Candlelight Organic Light-Emitting Diode with a Very Low Color Temperature. Molecules 2021, 26, 7558. [Google Scholar] [CrossRef]
  11. Korshunov, V.M.; Chmovzh, T.N.; Knyazeva, E.A.; Taydakov, I.V.; Mikhalchenko, L.V.; Varaksina, E.A.; Saifutyarov, R.S.; Avetissov, I.C.; Rakitin, O.A. A Novel Candle Light-Style OLED with a Record Low Colour Temperature. Chem. Commun. 2019, 55, 13354–13357. [Google Scholar] [CrossRef]
  12. Jou, J.-H.; Kumar, S.; An, C.-C.; Singh, M.; Yu, H.-H.; Hsieh, C.-Y.; Lin, Y.-X.; Sung, C.-F.; Wang, C.-W. Enabling a Blue-Hazard Free General Lighting Based on Candle Light-Style OLED. Opt. Express 2015, 23, A576. [Google Scholar] [CrossRef]
  13. Jou, J.H.; Singh, M.; Su, Y.T.; Liu, S.H.; He, Z.K. Blue-Hazard-Free Candlelight OLED. J. Vis. Exp. 2017, 121, 54644. [Google Scholar] [CrossRef]
  14. Jou, J.H.; Hsieh, C.Y.; Tseng, J.R.; Peng, S.H.; Jou, Y.C.; Hong, J.H.; Shen, S.M.; Tang, M.C.; Chen, P.C.; Lin, C.H. Candle Light-Style Organic Light-Emitting Diodes. Adv. Funct. Mater. 2013, 23, 2750–2757. [Google Scholar] [CrossRef]
  15. Quarterly Display Capex and Equipment Market Share Report—Display Supply Chain Consultants. Available online: https://www.displaysupplychain.com/report/quarterly-display-capex-and-equipment-market-share-report (accessed on 1 May 2022).
  16. Worldwide-LED Lighting Market Size 2021|Statista. Available online: https://www.statista.com/statistics/753939/global-led-luminaire-market-size/ (accessed on 2 May 2022).
  17. ElectroniCast Sees a Fast Growing OLED Lighting Market Starting in 2015|OLED-Info. Available online: https://www.oled-info.com/electronicast-sees-fast-growing-oled-lighting-market-starting-2015 (accessed on 2 May 2022).
  18. Steven Van Slyke Invented OLED Technology, Increasing Efficiency. Available online: https://www.invent.org/inductees/steven-van-slyke (accessed on 2 May 2022).
  19. OLED Displays and Their Applications|Learning Corner for Beginners. Available online: https://www.electronicsforu.com/resources/oled-displays-applications (accessed on 2 May 2022).
  20. Sony OLED|OLED-Info. Available online: https://www.oled-info.com/sony-oled (accessed on 2 May 2022).
  21. AUO. Available online: https://www.auo.com/en-global/New_Archive/detail/News_Archive_Technology_190823 (accessed on 2 May 2022).
  22. BOE Demonstrates a 55″ 8K Ink-Jet Printed OLED TV Prototype|OLED-Info. Available online: https://www.oled-info.com/boe-demonstrates-55-8k-ink-jet-printed-oled-tv-prototype (accessed on 2 May 2022).
  23. LG OLED TV 2016 Display Technology Shoot-Out. Available online: https://www.displaymate.com/OLED_TV2016_ShootOut_1.htm (accessed on 2 May 2022).
  24. Kosai, S.; Badin, A.B.; Qiu, Y.; Matsubae, K.; Suh, S.; Yamasue, E. Evaluation of Resource Use in the Household Lighting Sector in Malaysia Considering Land Disturbances through Mining Activities. Resour. Conserv. Recycl. 2021, 166, 105343. [Google Scholar] [CrossRef]
  25. Jou, J.; Su, Y.; Liu, S.; He, Z.; Sahoo, S.; Yu, H.-H.; Chen, S.-Z.; Wang, C.-W.; Lee, J.-R. Wet-Process Feasible Candlelight OLED. J. Mater. Chem. C 2016, 4, 6070–6077. [Google Scholar] [CrossRef]
  26. Jou, J.; Hsieh, C.; Chen, P.; Kumar, S.; Hong, J.H. Candlelight Style Organic Light-Emitting Diode: A Plausibly Human-Friendly Safe Night Light. J. Photonics Energy 2014, 4, 043598. [Google Scholar] [CrossRef]
  27. Sasabe, H.; Kido, J. Development of High Performance OLEDs for General Lighting. J. Mater. Chem. C Mater. 2013, 1, 1699–1707. [Google Scholar] [CrossRef]
  28. 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]
  29. Panasonic Developed a 114 Lm/W OLED Panel-Claims World’s Most Efficient Panel|OLED-Info. Available online: https://www.oled-info.com/panasonic-developed-114-lmw-oled-panel-claims-worlds-most-efficient-panel (accessed on 2 May 2022).
  30. Panasonic Develops World’s Highest Efficiency White OLED for Lighting|Headquarters News|Panasonic Newsroom Global. Available online: https://news.panasonic.com/global/press/data/2013/05/en130524-6/en130524-6.html (accessed on 2 May 2022).
  31. Kato, K.; Iwasaki, T.; Tsujimura, T. Over 130 Lm/W All-Phosphorescent White OLEDs for next-Generation Lighting. J. Photopolym. Sci. Technol. 2015, 28, 335–340. [Google Scholar] [CrossRef]
  32. Philips-GL350-Oled-Product-Sheet-Philips Lighting. Available online: https://www.yumpu.com/en/document/view/3069680/philips-gl350-oled-product-sheet-philips-lighting (accessed on 15 May 2022).
  33. OSRAM Presents the First OLED Luminaire—OSRAM Group Website. Available online: https://www.osram-group.com/en/media/press-releases/pr-2010/05-11-2010 (accessed on 15 May 2022).
  34. LG Chem’s 320x320 Mm OLED Lighting Panel Now in Production, Company Develops New Integration Solutions|OLED-Info. Available online: https://www.oled-info.com/lg-chems-320x320-mm-oled-lighting-panel-now-production-company-develops-new-integration-solutions (accessed on 17 May 2022).
  35. OLEDWorks Closes Philips Lumiblade Deal and Announces New OLED Panels|LEDs Magazine. Available online: https://www.ledsmagazine.com/leds-ssl-design/oleds/article/16696751/oledworks-closes-philips-lumiblade-deal-and-announces-new-oled-panels (accessed on 15 May 2022).
  36. Lumiotec’s New P09 OLED Lighting Panels Feature 45 Lm/W, 40,000 Hours Lifetime|OLED-Info. Available online: https://www.oled-info.com/lumiotecs-new-p09-oled-lighting-panels-feature-45-lmw-40000-hours-lifetime (accessed on 17 May 2022).
  37. OLED Lighting Introduction and Market Status|OLED-Info. Available online: https://www.oled-info.com/oled-lighting (accessed on 17 May 2022).
  38. Products|OLED Lighting|KONICA MINOLTA. Available online: https://www.konicaminolta.com/oled/products/ (accessed on 17 May 2022).
  39. Mitsubishi and Pioneer to Start Mass Producing Color-Tunable OLEDs Made Using a Wet-Coating Process|OLED-Info. Available online: https://www.oled-info.com/mitsubishi-and-pioneer-start-mass-producing-color-tunable-oleds-made-using-wet-coating-process (accessed on 17 May 2022).
  40. Kaneka Demonstrate Their New 50,000 Hours OLED Lighting Panels|OLED-Info. Available online: https://www.oled-info.com/kaneka-demonstrate-their-new-50000-hours-oled-lighting-panels (accessed on 17 May 2022).
  41. Introduction to OLED Lighting (Acuity Brands)-YouTube. Available online: https://www.youtube.com/watch?v=3nVupTrnw4s (accessed on 27 May 2022).
  42. CYNORA Announces Availability of Industry’s First Device Test Kits for TADF Deep Green Emitters for Next-Generation OLED Displays-CynoraCynora. Available online: https://cynora.com/cynora-announces-availability-of-industrys-first-device-test-kits-for-tadf-deep-green-emitters-for-next-generation-oled-displays-2/ (accessed on 27 May 2022).
  43. Nakamura, S. Shuji Nakamura-Nobel Lecture: Background Story of the Invention of Efficient Blue InGaN Light Emitting Diodes. Rev. Mod. Phys. 2015, 87, 1139–1151. [Google Scholar] [CrossRef]
  44. Blue LEDs-Filling the World with New Light. Available online: https://www.nobelprize.org/uploads/2018/06/popular-physicsprize2014-1.pdf (accessed on 15 August 2023).
  45. Nobel Prize ® and the Nobel Prize ® Medal Design Mark Are Registrated Trademarks of the Nobel Foundation. 2014. Available online: https://www.theguardian.com/science/2014/oct/07/nobel-prize-physics-inventors-blue-light-emitting-diodes-isamu-akasaki-hiroshi-amano-shuji-nakam#:~:text=Isamu%20Akasaki%20and%20Hiroshi%20Amano,and%20environment%2Dfriendly%20light%20source (accessed on 15 August 2023).
  46. Introduction to Color Imaging Science-Hsien-Che Lee-Google Books. Available online: https://books.google.com.tw/books/about/Introduction_to_Color_Imaging_Science.html?id=rN3kOMOPMmEC&redir_esc=y (accessed on 27 May 2022).
  47. Subtractive Color–HiSoUR–Hi So You Are. Available online: https://www.hisour.com/subtractive-color-24755/ (accessed on 27 May 2022).
  48. Additive and Subtractive Color Mixing. Available online: https://isle.hanover.edu/Ch06Color/Ch06ColorMixer.html (accessed on 27 May 2022).
  49. Taming the Wide Gamut Using SRGB Emulation|PC Monitors. Available online: https://pcmonitors.info/articles/taming-the-wide-gamut-using-srgb-emulation/ (accessed on 27 May 2022).
  50. Understanding the Color Gamut of an LCD Monitor. Available online: https://www.eizo.com/library/basics/lcd_monitor_color_gamut/ (accessed on 27 May 2022).
  51. Jou, J.H.; Kumar, S.; Fang, P.H.; Venkateswararao, A.; Thomas, K.R.J.; Shyue, J.J.; Wang, Y.C.; Li, T.H.; Yu, H.H. Highly Efficient Ultra-Deep Blue Organic Light-Emitting Diodes with a Wet- and Dry-Process Feasible Cyanofluorene Acetylene Based Emitter. J. Mater. Chem. C Mater. 2015, 3, 2182–2194. [Google Scholar] [CrossRef]
  52. Kim, H.K.; Cho, S.H.; Oh, J.R.; Lee, Y.H.; Lee, J.H.; Lee, J.G.; Kim, S.K.; Park, Y.I.; Park, J.W.; Do, Y.R. Deep Blue, Efficient, Moderate Microcavity Organic Light-Emitting Diodes. Org. Electron. 2010, 11, 137–145. [Google Scholar] [CrossRef]
  53. Li, W.; Yao, L.; Liu, H.; Wang, Z.; Zhang, S.; Xiao, R.; Zhang, H.; Lu, P.; Yang, B.; Ma, Y. Highly Efficient Deep-Blue OLED with an Extraordinarily Narrow FHWM of 35 nm and a y Coordinate <0.05 Based on a Fully Twisting Donor-Acceptor Molecule. J. Mater. Chem. C Mater. 2014, 2, 4733–4736. [Google Scholar] [CrossRef]
  54. Reineke, S.; Schwartz, G.; Walzer, K.; Leo, K. Reduced Efficiency Roll-off in Phosphorescent Organic Light Emitting Diodes by Suppression of Triplet-Triplet Annihilation. Appl. Phys. Lett. 2007, 91, 123508. [Google Scholar] [CrossRef]
  55. Yang, H.; Shi, Y.; Zhao, Y.; Meng, Y.; Hu, W.; Hou, J.; Liu, S. High Colour Rendering Index White Organic Light-Emitting Devices with Three Emitting Layers. Displays 2008, 29, 327–332. [Google Scholar] [CrossRef]
  56. Jou, J.H.; Chou, Y.C.; Shen, S.M.; Wu, M.H.; Wu, P.S.; Lin, C.R.; Wu, R.Z.; Chen, S.H.; Wei, M.K.; Wang, C.W. High-Efficiency, Very-High Color Rendering White Organic Light-Emitting Diode with a High Triplet Interlayer. J. Mater. Chem. 2011, 21, 18523–18526. [Google Scholar] [CrossRef]
  57. Lumiotec Announces New High CRI OLED Lighting Panels, Targets Museums|OLED-Info. Available online: https://www.oled-info.com/lumiotec-announces-new-high-cri-oled-lighting-panels-targets-museums (accessed on 27 May 2022).
  58. LG Chem Produces High-CRI OLED Panels That Achieve 20,000-Hour Lifetimes|LEDs Magazine. Available online: https://www.ledsmagazine.com/company-newsfeed/article/16680965/lg-chem-produces-highcri-oled-panels-that-achieve-20000hour-lifetimes (accessed on 27 May 2022).
  59. Adachi, C.; Tsutsui, T.; Saito, S. Blue Light-emitting Organic Electroluminescent Devices. Appl. Phys. Lett. 1998, 56, 799. [Google Scholar] [CrossRef]
  60. Adachi, C.; Kwong, R.C.; Djurovich, P.; Adamovich, V.; Baldo, M.A.; Thompson, M.E.; Forrest, S.R. Endothermic Energy Transfer: A Mechanism for Generating Very Efficient High-Energy Phosphorescent Emission in Organic Materials. Appl. Phys. Lett. 2001, 79, 2082. [Google Scholar] [CrossRef]
  61. Kawamura, Y.; Yanagida, S.; Forrest, S.R. Energy Transfer in Polymer Electrophosphorescent Light Emitting Devices with Single and Multiple Doped Luminescent Layers. J. Appl. Phys. 2002, 92, 87. [Google Scholar] [CrossRef]
  62. Youn Lee, S.; Yasuda, T.; Nomura, H.; Adachi, C. High-Efficiency Organic Light-Emitting Diodes Utilizing Thermally Activated Delayed Fluorescence from Triazine-Based Donor–Acceptor Hybrid Molecules. Appl. Phys. Lett. 2012, 101, 093306. [Google Scholar] [CrossRef]
  63. Nakanotani, H.; Masui, K.; Nishide, J.; Shibata, T.; Adachi, C. Promising Operational Stability of High-Efficiency Organic Light-Emitting Diodes Based on Thermally Activated Delayed Fluorescence. Sci. Rep. 2013, 3, 2127. [Google Scholar] [CrossRef] [PubMed]
  64. Masui, K.; Nakanotani, H.; Adachi, C. Analysis of Exciton Annihilation in High-Efficiency Sky-Blue Organic Light-Emitting Diodes with Thermally Activated Delayed Fluorescence. Org. Electron. 2013, 14, 2721–2726. [Google Scholar] [CrossRef]
  65. Lee, J.; Shizu, K.; Tanaka, H.; Nomura, H.; Yasuda, T.; Adachi, C. Oxadiazole- and Triazole-Based Highly-Efficient Thermally Activated Delayed Fluorescence Emitters for Organic Light-Emitting Diodes. J. Mater. Chem. C Mater. 2013, 1, 4599–4604. [Google Scholar] [CrossRef]
  66. Mallesham, G.; Swetha, C.; Niveditha, S.; Mohanty, M.E.; Babu, N.J.; Kumar, A.; Bhanuprakash, K.; Rao, V.J. Phosphine Oxide Functionalized Pyrenes as Efficient Blue Light Emitting Multifunctional Materials for Organic Light Emitting Diodes. J. Mater. Chem. C Mater. 2015, 3, 1208–1224. [Google Scholar] [CrossRef]
  67. Liu, W.; Zheng, C.J.; Wang, K.; Chen, Z.; Chen, D.Y.; Li, F.; Ou, X.M.; Dong, Y.P.; Zhang, X.H. Novel Carbazol-Pyridine-Carbonitrile Derivative as Excellent Blue Thermally Activated Delayed Fluorescence Emitter for Highly Efficient Organic Light-Emitting Devices. ACS Appl. Mater. Interfaces 2015, 7, 18930–18936. [Google Scholar] [CrossRef] [PubMed]
  68. Hiraga, Y.; Nishide, J.I.; Nakanotani, H.; Aonuma, M.; Adachi, C. High-Efficiency Sky-Blue Organic Light-Emitting Diodes Utilizing Thermally-Activated Delayed Fluorescence. IEICE Trans. Electron. 2015, E98.C, 971–976. [Google Scholar] [CrossRef]
  69. Organic Electroluminescence Device. Available online: https://www.freepatentsonline.com/5389444.pdf (accessed on 28 April 2022).
  70. Specification, E.P. EP000875947B1. 2002. Available online: https://data.epo.org/publication-server/document?iDocId=2079949&iFormat=0 (accessed on 15 August 2023).
  71. WO2001039234 Organic Light Emitting Diode Having a Blue Phosphorescent Molecule as an Emitter. Available online: https://patentscope.wipo.int/search/en/detail.jsf?docId=WO2001039234 (accessed on 28 April 2022).
  72. Takita, Y. Highly Efficient Deep-Blue Fluorescent Dopant for Achieving Low-Power OLED Display Satisfying BT. 2020 Chromaticity. J. Soc. Inf. Disp. 2018, 26, 55–63. [Google Scholar] [CrossRef]
  73. Liu, H.; Kang, L.; Li, J.; Liu, F.; He, X.; Ren, S.; Tang, X.; Lv, C.; Lu, P. Highly Efficient Deep-Blue Organic Light-Emitting Diodes Based on Pyreno[4,5-d]Imidazole-Anthracene Structural Isomers. J. Mater. Chem. C 2019, 7, 10273–10280. [Google Scholar] [CrossRef]
  74. Zou, Y.; Zou, J.; Ye, T.; Li, H.; Yang, C.; Wu, H.; Ma, D.; Qin, J.; Cao, Y. Unexpected Propeller-Like Hexakis(Fluoren-2-Yl)Benzene Cores for Six-Arm Star-Shaped Oligofluorenes: Highly Efficient Deep-Blue Fluorescent Emitters and Good Hole-Transporting Materials. Adv. Funct. Mater. 2013, 23, 1781–1788. [Google Scholar] [CrossRef]
  75. Wang, L.; Jiang, Y.; Luo, J.; Zhou, Y.; Zhou, J.; Wang, J.; Pei, J.; Cao, Y. Highly Efficient and Color-Stable Deep-Blue Organic Light-Emitting Diodes Based on a Solution-Processible Dendrimer. Adv. Mater. 2009, 21, 4854–4858. [Google Scholar] [CrossRef]
  76. Li, W.; Zhang, X.; Zhang, Y.; Xu, K.; Xu, J.; Wang, H.; Li, H.; Guo, J.; Mo, J.; Yang, P. Achieving Ultra-High Efficiency by Tuning Hole Transport and Carrier Balance in Fluorescent Blue Organic Light-Emitting Diode with Extremely Simple Structure. Synth. Met. 2018, 245, 111–115. [Google Scholar] [CrossRef]
  77. Zhen, C.G.; Dai, Y.F.; Zeng, W.J.; Ma, Z.; Chen, Z.K.; Kieffer, J. Achieving Highly Efficient Fluorescent Blue Organic Light-Emitting Diodes through Optimizing Molecular Structures and Device Configuration. Adv. Funct. Mater. 2011, 21, 699–707. [Google Scholar] [CrossRef]
  78. Liu, M.; Li, X.-L.; Cheng Chen, D.; Xie, Z.; Cai, X.; Xie, G.; Liu, K.; Tang, J.; Su, S.-J.; Cao, Y.; et al. Study of Configuration Differentia and Highly Efficient, Deep-Blue, Organic Light-Emitting Diodes Based on Novel Naphtho[1,2-d]Imidazole Derivatives. Adv. Funct. Mater. 2015, 25, 5190–5198. [Google Scholar] [CrossRef]
  79. Li, Z.; Gan, G.; Ling, Z.; Guo, K.; Si, C.; Lv, X.; Wang, H.; Wei, B.; Hao, Y. Easily Available, Low-Cost 9,9′-Bianthracene Derivatives as e Ffi Cient Blue Hosts and Deep-Blue Emitters in OLEDs. Org. Electron. 2019, 66, 24–31. [Google Scholar] [CrossRef]
  80. 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 Mater. 2019, 7, 5874–5888. [Google Scholar] [CrossRef]
  81. Jeong, S.; Kim, M.; Kim, S.H.; Hong, J. Efficient Deep-Blue Emitters Based on Triphenylamine-Linked Benzimidazole Derivatives for Nondoped Fluorescent Organic Light-Emitting Diodes. Org. Electron. 2013, 14, 2497–2504. [Google Scholar] [CrossRef]
  82. Jeon, Y.; Lee, J.; Kim, J.; Lee, C.; Gong, M. Deep-Blue OLEDs Using Novel Efficient Spiro-Type Dopant Materials. Org. Electron. 2010, 11, 1844–1852. [Google Scholar] [CrossRef]
  83. Wei, B.; Liu, J.; Zhang, Y.; Zhang, J.; Peng, H.; Fan, H.; He, Y.; Gao, X. Stable, Glassy, and Versatile Binaphthalene Derivatives Capable of Efficient Hole Transport, Hosting, and Deep-Blue Light Emission. Adv. Funct. Mater. 2010, 20, 2448–2458. [Google Scholar] [CrossRef]
  84. Wen, S.; Lee, M.; Chen, C.H. Recent Development of Blue Fluorescent OLED Materials and Devices. J. Disp. Technol. 2005, 1, 90–99. [Google Scholar] [CrossRef]
  85. Patil, V.V.; Lee, K.H.; Lee, J.Y. For Blue Fluorescent Organic Light-Emitting Diodes. J. Mater. Chem. C 2020, 8, 8320–8327. [Google Scholar] [CrossRef]
  86. Park, H.-Y.; Maheshwaran, A.; Moon, C.-K.; Lee, H.; Sudhaker Reddy, S.; Gopalan Sree, V.; Yoon, J.; Won Kim, J.; Hyuk Kwon, J.; Kim, J.-J.; et al. External Quantum Efficiency Exceeding 24% with CIEy Value of 0.08 Using a Novel Carbene-Based Iridium Complex in Deep-Blue Phosphorescent Organic Light-Emitting Diodes. Adv. Mater. 2020, 32, 2002120. [Google Scholar] [CrossRef]
  87. Sarma, M.; Tsai, W.L.; Lee, W.K.; Chi, Y.; Wu, C.C.; Liu, S.H.; Chou, P.T.; Wong, K.T. Anomalously Long-Lasting Blue PhOLED Featuring Phenyl-Pyrimidine Cyclometalated Iridium Emitter. Chem 2017, 3, 461–476. [Google Scholar] [CrossRef]
  88. Li, W.; Li, J.; Liu, D.; Jin, Q. Simple Bipolar Host Materials for High-E Ffi Ciency Blue, Green, and White Phosphorescence OLEDs. ACS Appl. Mater. Interfaces 2016, 8, 22382–22391. [Google Scholar] [CrossRef]
  89. Kim, M.; Lee, I.H.; Lee, J.; Hyun, S.; Park, K.; Kang, Y. Chemistry Phenanthridine Based Ir(III) Complex for Blue Phosphorescent Organic Light-Emitting Diodes with Long-Term Operational Stability. J. Ind. Eng. Chem. 2018, 58, 386–390. [Google Scholar] [CrossRef]
  90. Oh, C.S.; Lee, Y.; Noh, H.; Han, S. Molecular Design of Host Materials for High Power Efficiency in Blue Phosphorescent Organic Light-Emitting Diodes Doped with an Imidazole Ligand Based Triplet Emitter. J. Mater. Chem. C 2016, 4, 3792–3797. [Google Scholar] [CrossRef]
  91. Ying, S.; Pang, P.; Zhang, S.; Sun, Q.; Dai, Y.; Qiao, X.; Yang, D.; Chen, J.; Ma, D. Superior Efficiency and Low-Efficiency Roll-Off White Organic Light-Emitting Diodes Based on Multiple Exciplexes as Hosts Matched to Phosphor Emitters. ACS Appl. Mater. Interfaces 2019, 11, 31078–31086. [Google Scholar] [CrossRef]
  92. Son, Y.H.; Kim, Y.J.; Park, M.J.; Oh, H.Y.; Park, J.S.; Yang, J.H.; Suh, M.C.; Kwon, J.H. Small Single–Triplet Energy Gap Bipolar Host Materials for Phosphorescent Blue and White Organic Light Emitting Diodes. J. Mater. Chem. C Mater. 2013, 1, 5008–5014. [Google Scholar] [CrossRef]
  93. Feng, Y.; Zhuang, X.; Zhu, D.; Liu, Y.; Bryce, M.R. Heteroleptic Phosphorescent Iridium(III). J. Mater. Chem. C 2016, 4, 10246–10252. [Google Scholar] [CrossRef]
  94. Diode, L.; Jou, J.; Wang, W.; Hsu, M.; Liu, C.; Chen, C. Small Nano-Dot Incorporated High-Efficiency Phosphorescent Blue Organic Small Nano-Dot Incorporated High-Efficiency Phosphorescent Blue Organic Light-Emitting Diode. PIERS 2008, 4, 351–355. [Google Scholar] [CrossRef]
  95. Wu, Z.G.; Jing, Y.M.; Lu, G.Z.; Zhou, J.; Zheng, Y.X.; Zhou, L.; Wang, Y.; Pan, Y. Novel Design of Iridium Phosphors with Pyridinylphosphinate Ligands for High-Efficiency Blue Organic Light-Emitting Diodes. Sci. Rep. 2016, 6, 38478. [Google Scholar] [CrossRef]
  96. Wee, K.R.; Kim, A.L.; Jeong, S.Y.; Kwon, S.; Kang, S.O. The Relationship between the Device Performance and Hole Mobility of Host Materials in Mixed Host System for Deep Blue Phosphorescent Organic Light Emitting Devices. Org. Electron. 2011, 12, 1973–1979. [Google Scholar] [CrossRef]
  97. Weaver, M.S.; Tung, Y.-J.; D’Andrade, B.; Esler, J.; Brown, J.J.; Lin, C.; Mackenzie, P.B.; Walters, R.W.; Tsai, J.-Y.; Brown, C.S.; et al. 11.1: Invited Paper: Advances in Blue Phosphorescent Organic Light-Emitting Devices. SID Symp. Dig. Tech. Pap. 2006, 37, 127. [Google Scholar] [CrossRef]
  98. Kumar Konidena, R.; Jae Chung, W.; Yeob Lee, J.; Konidena, R.K.; Chung, W.J.; Lee, J.Y. Molecular Engineering of Cyano-Substituted Carbazole-Based Host Materials for Simultaneous Achievement of High Efficiency and Long Lifetime in Blue Phosphorescent Organic Light-Emitting Diodes. Adv. Electron. Mater. 2020, 6, 2000132. [Google Scholar] [CrossRef]
  99. Jung, M.; Lee, K.H.; Lee, J.Y.; Kim, T. A Bipolar Host Based High Triplet Energy Electroplex for an over 10,000 h Lifetime in Pure Blue Phosphorescent Organic Light-Emitting Diodes. Mater. Horiz. 2020, 7, 559–565. [Google Scholar] [CrossRef]
  100. Yang, C.Y.; Jang, H.J.; Lee, K.H.; Byeon, S.Y.; Lee, J.Y.; Jeong, H. 12-1: Analysis of Key Factors Affecting the Lifetime of Blue Phosphorescent OLED Using CN Modified Blue Host Materials. SID Symp. Dig. Tech. Pap. 2019, 50, 141–144. [Google Scholar] [CrossRef]
  101. Han, S.H.; Lee, K.H.; Lee, J.Y. Spatial Separation of Two Blue Triplet Emitters for Improved Lifetime in Blue Phosphorescent Organic Light-Emitting Diodes by Confining Excitons at the Interface between Two Emitting Layers. Org. Electron. 2019, 69, 227–231. [Google Scholar] [CrossRef]
  102. Klimes, K.; Zhu, Z.Q.; Li, J. Efficient Blue Phosphorescent OLEDs with Improved Stability and Color Purity through Judicious Triplet Exciton Management. Adv. Funct. Mater. 2019, 29, 1903068. [Google Scholar] [CrossRef]
  103. Yao, J.; Xiao, S.; Zhang, S.; Sun, Q.; Dai, Y.; Qiao, X.; Yang, D.; Chen, J.; Ma, D. High Efficiency, Low Efficiency Roll-off and Long Lifetime Fluorescent White Organic Light-Emitting Diodes Based on Strategic Management of Triplet Excitons via Triplet–Triplet Annihilation up-Conversion and Phosphor Sensitization. J. Mater. Chem. C Mater. 2020, 8, 8077–8084. [Google Scholar] [CrossRef]
  104. Liang, X.; Han, H.B.; Yan, Z.P.; Liu, L.; Zheng, Y.X.; Meng, H.; Huang, W. Versatile Functionalization of Trifluoromethyl Based Deep Blue Thermally Activated Delayed Fluorescence Materials for Organic Light Emitting Diodes. New J. Chem. 2018, 42, 4317–4323. [Google Scholar] [CrossRef]
  105. Lim, H.; Cheon, J.; Woo, S.-J.; Kwon, S.-K.; Kim, Y.-H.; Kim, J.-J.; Lim, H.; Woo, S.-J.; Kim, J.-J.; Cheon, H.J.; et al. Highly Efficient Deep-Blue OLEDs Using a TADF Emitter with a Narrow Emission Spectrum and High Horizontal Emitting Dipole Ratio. Adv. Mater. 2020, 32, 2004083. [Google Scholar] [CrossRef]
  106. Li, W.; Li, B.; Cai, X.; Gan, L.; Xu, Z.; Li, W.; Liu, K.; Chen, D.; Su, S.J. Tri-Spiral Donor for High Efficiency and Versatile Blue Thermally Activated Delayed Fluorescence Materials. Angew. Chem. Int. Ed. 2019, 58, 11301–11305. [Google Scholar] [CrossRef] [PubMed]
  107. Miwa, T.; Kubo, S.; Shizu, K.; Komino, T.; Adachi, C.; Kaji, H. Blue Organic Light-Emitting Diodes Realizing External Quantum Efficiency over 25% Using Thermally Activated Delayed Fluorescence Emitters. Sci. Rep. 2017, 7, 284. [Google Scholar] [CrossRef] [PubMed]
  108. Kim, H.J.; Godumala, M.; Kim, S.K.; Yoon, J.; Kim, C.Y.; Park, H.; Kwon, J.H.; Cho, M.J.; Choi, D.H. Color-Tunable Boron-Based Emitters Exhibiting Aggregation-Induced Emission and Thermally Activated Delayed Fluorescence for Efficient Solution-Processable Nondoped Deep-Blue to Sky-Blue OLEDs. Adv. Opt. Mater. 2020, 8, 1902175. [Google Scholar] [CrossRef]
  109. Xie, F.M.; An, Z.D.; Xie, M.; Li, Y.Q.; Zhang, G.H.; Zou, S.J.; Chen, L.; Chen, J.D.; Cheng, T.; Tang, J.X. Tert-Butyl Substituted Hetero-Donor TADF Compounds for Efficient Solution-Processed Non-Doped Blue OLEDs. J. Mater. Chem. C Mater. 2020, 8, 5769–5776. [Google Scholar] [CrossRef]
  110. Ganesan, P.; Chen, D.G.; Liao, J.L.; Li, W.C.; Lai, Y.N.; Luo, D.; Chang, C.H.; Ko, C.L.; Hung, W.Y.; Liu, S.W.; et al. Isomeric Spiro-[Acridine-9,9′-Fluorene]-2,6-Dipyridylpyrimidine Based TADF Emitters: Insights into Photophysical Behaviors and OLED Performances. J. Mater. Chem. C Mater. 2018, 6, 10088–10100. [Google Scholar] [CrossRef]
  111. Oh, C.S.; Pereira, D.D.S.; Han, S.H.; Park, H.J.; Higginbotham, H.F.; Monkman, A.P.; Lee, J.Y. Dihedral Angle Control of Blue Thermally Activated Delayed Fluorescent Emitters through Donor Substitution Position for Efficient Reverse Intersystem Crossing. ACS Appl. Mater. Interfaces 2018, 10, 35420–35429. [Google Scholar] [CrossRef]
  112. Ban, X.; Chen, F.; Pan, J.; Liu, Y.; Zhu, A.; Jiang, W.; Sun, Y. Exciplex Formation and Electromer Blocking for Highly Efficient Blue Thermally Activated Delayed Fluorescence OLEDs with All-Solution-Processed Organic Layers. Chem.-A Eur. J. 2020, 26, 3090–3102. [Google Scholar] [CrossRef]
  113. Yoon, J.; Kim, S.K.; Kim, H.J.; Choi, S.; Jung, S.W.; Lee, H.; Kim, J.Y.; Yoon, D.W.; Han, C.W.; Chae, W.S.; et al. Asymmetric Host Molecule Bearing Pyridine Core for Highly Efficient Blue Thermally Activated Delayed Fluorescence OLEDs. Chem.-A Eur. J. 2020, 26, 16383–16391. [Google Scholar] [CrossRef]
  114. Su, L.; Cao, F.; Cheng, C.; Tsuboi, T.; Zhu, Y.; Deng, C.; Zheng, X.; Wang, D.; Liu, Z.; Zhang, Q. High Fluorescence Rate of Thermally Activated Delayed Fluorescence Emitters for Efficient and Stable Blue OLEDs. ACS Appl. Mater. Interfaces 2020, 12, 31706–31715. [Google Scholar] [CrossRef]
  115. Ihn, S.G.; Lee, N.; Jeon, S.O.; Sim, M.; Kang, H.; Jung, Y.; Huh, D.H.; Son, Y.M.; Lee, S.Y.; Numata, M.; et al. An Alternative Host Material for Long-Lifespan Blue Organic Light-Emitting Diodes Using Thermally Activated Delayed Fluorescence. Adv. Sci. 2017, 4, 1600502. [Google Scholar] [CrossRef]
  116. Feng, X.; Hu, J.-Y.; Yi, L.; Seto, N.; Tao, Z.; Redshaw, C.; Elsegood, M.R.J.; Yamato, T. Pyrene-Based Y-shaped Solid-State Blue Emitters: Synthesis, Characterization, and Photoluminescence. Chem. Asian J. 2012, 7, 2854–2863. [Google Scholar] [CrossRef] [PubMed]
  117. Yang, C.Y.; Lee, K.H.; Lee, J.Y. Zig-Zag Type Molecular Design Strategy of N-Type Hosts for Sky-Blue Thermally-Activated Delayed Fluorescence Organic Light-Emitting Diodes. Chem.-A Eur. J. 2020, 26, 2429–2435. [Google Scholar] [CrossRef]
  118. Hamada, Y.; Adachi, C.; Tsutsui, T.; Saito, S. Organic Electroluminescent Devices with Bright Blue Emission. Optoelectron. Tokyo 1992, 7, 83–93. [Google Scholar]
  119. Hamada, Y.; Adachi, C.; Tsutsui, T.; Saito, S. Blue-Light-Emitting Organic Electroluminescent Devices with Oxadiazole Dimer Dyes as an Emitter. Jpn. J. Appl. Phys. 1992, 31, 1812–1816. [Google Scholar] [CrossRef]
  120. Grem, G.; Leditzky, G.; Ullrich, B.; Leising, G. Realization of a Blue-Light-Emitting Device Using Poly(p-Phenylene). Adv. Mater. 1992, 4, 36–37. [Google Scholar] [CrossRef]
  121. Grem, G.; Leditzky, G.; Ullrich, B.; Leising, G. Blue Electroluminescent Device Based on a Conjugated Polymer. Synth. Met. 1992, 51, 383–389. [Google Scholar] [CrossRef]
  122. Hosokawa, C.; Kawasaki, N.; Sakamoto, S.; Kusumoto, T. Bright Blue Electroluminescence from Hole Transporting Polycarbonate. Appl. Phys. Lett. 1998, 61, 2503. [Google Scholar] [CrossRef]
  123. Hosokawa, C.; Tokailin, H.; Higashi, H.; Kusumoto, T. Transient Behavior of Organic Thin Film Electroluminescence. Appl. Phys. Lett. 1998, 60, 1220. [Google Scholar] [CrossRef]
  124. Hosokawa, C.; Tokailin, H.; Higashi, H.; Kusumoto, T. Transient Electroluminescence from Hole Transporting Emitting Layer in Nanosecond Region. Appl. Phys. Lett. 1998, 63, 1322. [Google Scholar] [CrossRef]
  125. Vestweber, H.; Sander, R.; Greiner, A.; Heitz, W.; Mahrt, R.F.; Bässler, H. Electroluminescence from Polymer Blends and Molecularly Doped Polymers. Synth. Met. 1994, 64, 141–145. [Google Scholar] [CrossRef]
  126. Hosokawa, C.; Tokailin, H.; Higashi, H.; Kusumoto, T. Efficient Electroluminescence of Distyrylarylene with Hole Transporting Ability. J. Appl. Phys. 1998, 78, 5831. [Google Scholar] [CrossRef]
  127. Tokailin, T.H.; Matsuura, M.; Higashi, H.; Hosokawa, C.; Kusumoto, T.; Tokailin, H.; Matsuura, M.; Kusumoto, T. Characteristics of Blue Organic Electroluminescent Devices with Distyryl Arylene Derivatives. Electron. Imaging 1993, 1910, 38–47. [Google Scholar] [CrossRef]
  128. Hosokawa, C.; Matsuura, M.; Eida, M.; Fukuoka, K.; Tokailin, H.; Kusumoto, T. 4.1: Invited Paper: Full-Color Organic EL Display. SID Symp. Dig. Tech. Pap. 1998, 29, 7–10. [Google Scholar] [CrossRef]
  129. Ho, M.H.; Chang, C.M.; Chu, T.Y.; Chen, T.M.; Chen, C.H. Iminodibenzyl-Substituted Distyrylarylenes as Dopants for Blue and White Organic Light-Emitting Devices. Org. Electron. 2008, 9, 101–110. [Google Scholar] [CrossRef]
  130. Wind, M.; Wiesler, U.M.; Saalwächter, K.; Müllen, K.; Spiess, H.W. Shape-Persistent Polyphenylene Dendrimers—Restricted Molecular Dynamics from Advanced Solid-State Nuclear Magnetic Resonance Techniques. Adv. Mater. 2001, 13, 752–756. [Google Scholar] [CrossRef]
  131. Rosenfeldt, S.; Dingenouts, N.; Pötschke, D.; Ballauff, M.; Berresheim, A.J.; Müllen, K.; Lindner, P. Analysis of the Spatial Dimensions of Fully Aromatic Dendrimers. Angew. Chem. Int. Ed. 2004, 43, 109–112. [Google Scholar] [CrossRef] [PubMed]
  132. Novel Green Light-Emitting Carbazole Derivatives: Potential Electroluminescent Materials-Thomas-2000-Advanced Materials-Wiley Online Library. Available online: https://onlinelibrary.wiley.com/doi/pdf/10.1002/1521-4095%28200012%2912%3A24%3C1949%3A%3AAID-ADMA1949%3E3.0.CO%3B2-X (accessed on 16 May 2022).
  133. Tonzola, C.J.; Alam, M.M.; Kaminsky, W.; Jenekhe, S.A. New N-Type Organic Semiconductors: Synthesis, Single Crystal Structures, Cyclic Voltammetry, Photophysics, Electron Transport, and Electroluminescence of a Series of Diphenylanthrazolines. J. Am. Chem. Soc. 2003, 125, 13548–13558. [Google Scholar] [CrossRef] [PubMed]
  134. Figueira-Duarte, T.M.; Del Rosso, P.G.; Trattnig, R.; Sax, S.; List, E.J.W.; Mullen, K. Designed Suppression of Aggregation in Polypyrene: Toward High-Performance Blue-Light-Emitting Diodes. Adv. Mater. 2010, 22, 990–993. [Google Scholar] [CrossRef] [PubMed]
  135. Xing, Y.; Xu, X.; Zhang, P.; Tian, W.; Yu, G.; Lu, P.; Liu, Y.; Zhu, D. Carbazole–Pyrene-Based Organic Emitters for Electroluminescent Device. Chem. Phys. Lett. 2005, 408, 169–173. [Google Scholar] [CrossRef]
  136. Sotoyama, W.; Sato, H.; Kinoshita, M.; Takahashi, T.; Matsuura, A.; Kodama, J.; Sawatari, N.; Inoue, H. 45.3: Tetra-Substituted Pyrenes: New Class of Blue Emitter for Organic Light-Emitting Diodes. SID Symp. Dig. Tech. Pap. 2003, 34, 1294–1297. [Google Scholar] [CrossRef]
  137. Miura, Y.; Yamano, E.; Tanaka, A.; Yamauchi, J. Generation, Isolation, and Characterization of N-(Arylthio)-7-Tert-Butyl- and N-(Arylthio)-2,7-Di-Tert-Butyl-1-Pyrenylaminyl Radicals. J. Org. Chem. 1994, 59, 3294–3300. [Google Scholar] [CrossRef]
  138. Grill, E.; Hauge, B.; Schiefelbein, J.; Martinez-Zapater, J.; Gil, P.; Iba, K.; Bevilacqua, P.C.; Kierzek, R.; Johnson, K.A.; Turner, D.H. Dynamics of Ribozyme Binding of Substrate Revealed by Fluorescence-Detected Stopped-Flow Methods. Science 1992, 258, 1355–1358. [Google Scholar] [CrossRef]
  139. Kim, J.; Beardslee, R.; Phillips, D.T.; Offen, H.W. Fluorescence Lifetimes of Pyrene Monomer and Excimer at High Pressures. J. Chem. Phys. 1969, 51, 1638. [Google Scholar] [CrossRef]
  140. Kropp, J.L.; Dawson, W.R.; Windsor, M.W. Radiative and Radiationless Processes in Aromatic Molecules. Pyrene. J. Phys. Chem. 1969, 73, 1747–1752. [Google Scholar] [CrossRef]
  141. Zhao, Z.; Li, J.H.; Chen, X.; Wang, X.; Lu, P.; Yang, Y. Solution-Processable Stiff Dendrimers: Synthesis, Photophysics, Film Morphology, and Electroluminescence. J. Org. Chem. 2009, 74, 383–395. [Google Scholar] [CrossRef] [PubMed]
  142. Jia, W.-L.; McCormick, T.; Liu, Q.-D.; Fukutani, H.; Motala, M.; Wang, R.-Y.; Tao, Y.; Wang, S. Diarylamino Functionalized Pyrene Derivatives for Use in Blue OLEDs and Complex Formation. J. Mater. Chem. 2004, 14, 3344–3350. [Google Scholar] [CrossRef]
  143. Lampkins, A.J.; O’Neil, E.J.; Smith, B.D. Bio-Orthogonal Phosphatidylserine Conjugates for Delivery and Imaging Applications. J. Org. Chem. 2008, 73, 6053–6058. [Google Scholar] [CrossRef] [PubMed]
  144. Vollmer, A.; Lorbach, D.; List, E.J.W.; Müllen, K.; Loi, M.A.; Manca, M.; Baumgarten, M.; Koch, N.; Trattnig, R.; Sax, S.; et al. Deep Blue Polymer Light Emitting Diodes Based on Easy to Synthesize, Non-Aggregating Polypyrene. Opt. Express 2011, 19, A1281–A1293. [Google Scholar] [CrossRef]
  145. Wu, K.C.; Ku, P.J.; Lin, C.S.; Shih, H.T.; Wu, F.I.; Huang, M.J.; Lin, J.J.; Chen, I.C.; Cheng, C.H. The Photophysical Properties of Dipyrenylbenzenes and Their Application as Exceedingly Efficient Blue Emitters for Electroluminescent Devices. Adv. Funct. Mater. 2008, 18, 67–75. [Google Scholar] [CrossRef]
  146. Zhao, Z.; Xu, X.; Wang, H.; Lu, P.; Yu, G.; Liu, Y. Zigzag Molecules from Pyrene-Modified Carbazole Oligomers: Synthesis, Characterization, and Application in OLEDs. J. Org. Chem. 2008, 73, 594–602. [Google Scholar] [CrossRef] [PubMed]
  147. Zhao, Z.; Xu, X.; Jiang, Z.; Lu, P.; Yu, G.; Liu, Y. Oligo(2,7-Fluorene Ethynylene)s with Pyrene Moieties: Synthesis, Characterization, Photoluminescence, and Electroluminescence. J. Org. Chem. 2007, 72, 8345–8353. [Google Scholar] [CrossRef] [PubMed]
  148. Chan, K.; Lim, J.; Yang, X.; Dodabalapur, A.; Jabbour, G.E.; Sellinger, A. High-Efficiency Pyrene-Based Blue Light Emitting Diodes: Aggregation Suppression Using a Calixarene 3D-Scaffold. Chem. Commun. 2012, 48, 5106–5108. [Google Scholar] [CrossRef] [PubMed]
  149. Wang, Z.; Xu, C.; Wang, W.; Dong, X.; Zhao, B.; Pigments, B.J.-D. Novel Pyrene Derivatives: Synthesis, Properties and Highly Efficient Non-Doped Deep-Blue Electroluminescent Device. Dye. Pigment. 2012, 92, 732–736. [Google Scholar] [CrossRef]
  150. Sonar, P.; Soh, M.S.; Cheng, Y.H.; Henssler, J.T.; Sellinger, A. 1,3,6,8-Tetrasubstituted Pyrenes: Solution-Processable Materials for Application in Organic Electronics. Org. Lett. 2010, 12, 3292–3295. [Google Scholar] [CrossRef] [PubMed]
  151. Moorthy, J.N.; Natarajan, P.; Venkatakrishnan, P.; Huang, D.F.; Chow, T.J. Steric Inhibition of π-Stacking: 1,3,6,8-Tetraarylpyrenes as Efficient Blue Emitters in Organic Light Emitting Diodes (OLEDs). Org. Lett. 2007, 9, 5215–5218. [Google Scholar] [CrossRef] [PubMed]
  152. Otsubo, T.; Aso, Y.; Takimiya, K. Functional Oligothiophenes as Advanced Molecular Electronic Materials. J. Mater. Chem. 2002, 12, 2565–2575. [Google Scholar] [CrossRef]
  153. Kumchoo, T.; Promarak, V.; Sudyoadsuk, T.; Sukwattanasinitt, M.; Rashatasakhon, P. Dipyrenylcarbazole Derivatives for Blue Organic Light-Emitting Diodes. Chem. Asian J. 2010, 5, 2162–2167. [Google Scholar] [CrossRef]
  154. Zhao, Z.; Xu, X.; Wang, F.; Yu, G.; Lu, P.; Liu, Y.; Zhu, D. Synthesis and Characterization of Light-Emitting Materials Composed of Carbazole, Pyrene and Fluorene. Synth. Met. 2006, 156, 209–214. [Google Scholar] [CrossRef]
  155. Mikroyannidis, J.A.; Fenenko, L.; Adachi, C. Synthesis and Photophysical Characteristics of 2,7-Fluorenevinylene-Based Trimers and Their Electroluminescence. J. Phys. Chem. B 2006, 110, 20317–20326. [Google Scholar] [CrossRef]
  156. Tang, C.; Liu, F.; Xia, Y.; Xie, L.; Wei, A.; Li, S.-B.; Fan, Q.-L.; Huang, W. Efficient 9-Alkylphenyl-9-Pyrenylfluorene Substituted Pyrene Derivatives with Improved Hole Injection for Blue Light-Emitting Diodes. J. Mater. Chem. 2006, 16, 4074–4080. [Google Scholar] [CrossRef]
  157. Tang, C.; Liu, F.; Xia, Y.; Lin, J.; Xie, L.; Zhong, G.-Y.; Fan, Q.-L.; Huang, W. Fluorene-Substituted Pyrenes—Novel Pyrene Derivatives as Emitters in Nondoped Blue OLEDs. Org. Electron. 2006, 7, 155–162. [Google Scholar] [CrossRef]
  158. Peng, Z.; Tao, S.; Zhang, X.; Tang, J.; Lee, C.S.; Lee, S.T. New Fluorene Derivatives for Blue Electroluminescent Devices: Influence of Substituents on Thermal Properties, Photoluminescence, and Electroluminescence. J. Phys. Chem. C 2008, 112, 2165–2169. [Google Scholar] [CrossRef]
  159. Liu, F.; Xie, L.H.; Tang, C.; Liang, J.; Chen, Q.Q.; Peng, B.; Wei, W.; Cao, Y.; Huang, W. Facile Synthesis of Spirocyclic Aromatic Hydrocarbon Derivatives Based on O-Halobiaryl Route and Domino Reaction for Deep-Blue Organic Semiconductors. Org. Lett. 2009, 11, 3850–3853. [Google Scholar] [CrossRef] [PubMed]
  160. Tao, S.; Zhou, Y.; Lee, C.S.; Zhang, X.; Lee, S.T. High-Efficiency Nondoped Deep-Blue-Emitting Organic Electroluminescent Device. Chem. Mater. 2010, 22, 2138–2141. [Google Scholar] [CrossRef]
  161. Thangthong, A.; Prachumrak, N.; Tarsang, R.; Keawin, T.; Jungsuttiwong, S.; Sudyoadsuka, T.; Promarak, V. Blue Light-Emitting and Hole-Transporting Materials Based on 9, 9-Bis (4-Diphenylaminophenyl) Fluorenes for Efficient Electroluminescent Devices. Mater. Chem. 2012, 22, 6869–6877. [Google Scholar] [CrossRef]
  162. Jou, J.; Chen, Y.; Tseng, J.; Wu, R.Z.; Shyue, J.J.; Thomas, K.R.J.; Kapoor, N.; Chen, C.-T.; Lin, Y.-P.; Wang, P.-H.; et al. The Use of a Polarity Matching and High-Energy Exciton Generating Host in Fabricating Efficient Purplish-Blue OLEDs from a Sky-Blue Emitter. J. Mater. Chem. 2012, 22, 15500–15506. [Google Scholar] [CrossRef]
  163. Tao, S.; Peng, Z.; Zhang, X.; Wang, P.; Lee, C.-S.; Lee, S.-T. Highly Efficient Non-doped Blue Organic Light-emitting Diodes Based on Fluorene Derivatives with High Thermal Stability. Adv. Funct. Mater. 2005, 15, 1716–1721. [Google Scholar] [CrossRef]
  164. Lai, S.; Tong, Q.; Chan, M.; Ng, T.; Lo, M.-F.; Lee, S.-T.; Lee, C.-S. Distinct Electroluminescent Properties of Triphenylamine Derivatives in Blue Organic Light-Emitting Devices. J. Mater. Chem. 2011, 21, 1206–1211. [Google Scholar] [CrossRef]
  165. Thomas, K.; Velusamy, M.; Lin, J.T.; Chuen, C.H.; Tao, Y.-T. Hexaphenylphenylene Dendronised Pyrenylamines for Efficient Organic Light-Emitting Diodes. J. Mater. Chem. 2005, 15, 4453–4459. [Google Scholar] [CrossRef]
  166. Tao, S.; Zhou, Y.; Lee, C.S.; Lee, S.T.; Huang, D.; Zhang, X. Highly Efficient Nondoped Blue Organic Light-Emitting Diodes Based on Anthracene-Triphenylamine Derivatives. J. Phys. Chem. C 2008, 112, 14603–14606. [Google Scholar] [CrossRef]
  167. Figueira-Duarte, T.M.; Müllen, K. Pyrene-Based Materials for Organic Electronics. Chem. Rev. 2011, 111, 7260–7314. [Google Scholar] [CrossRef] [PubMed]
  168. Kumar, D.; Thomas, K.R.J.; Lin, C.C.; Jou, J.H. Pyrenoimidazole-Based Deep-Blue-Emitting Materials: Optical, Electrochemical, and Electroluminescent Characteristics. Chem. Asian J. 2013, 8, 2111–2124. [Google Scholar] [CrossRef] [PubMed]
  169. Figueira-Duarte, T.M.; Simon, S.C.; Wagner, M.; Druzhinin, S.I.; Zachariasse, K.A.; Müllen, K. Polypyrene Dendrimers. Angew. Chem. Int. Ed. 2008, 47, 10175–10178. [Google Scholar] [CrossRef] [PubMed]
  170. Kotchapradist, P.; Prachumrak, N.; Tarsang, R.; Jungsuttiwong, S.; Keawin, T.; Sudyoadsuk, T.; Promarak, V. Pyrene-Functionalized Carbazole Derivatives as Non-Doped Blue Emitters for Highly Efficient Blue Organic Light-Emitting Diodes. J. Mater. Chem. C 2013, 1, 4916–4924. [Google Scholar]
  171. Shi, J.; Tang, C.W. Anthracene Derivatives for Stable Blue-Emitting Organic Electroluminescence Devices. Appl. Phys. Lett. 2002, 80, 3201. [Google Scholar] [CrossRef]
  172. Shih, B.P.; Chuang, C.; Chien, C.; Diau, E.W.; Shu, C. Highly Efficient Non-Doped Blue-Light-Emitting Diodes Based on an Anthrancene Derivative End-Capped with Tetraphenylethylene Groups. Adv. Funct. Mater. 2007, 17, 3141–3146. [Google Scholar] [CrossRef]
  173. Kim, S.; Yang, B.; Ma, Y.; Lee, J.; Park, J. Exceedingly Efficient Deep-Blue Electroluminescence from New Anthracenes Obtained Using Rational Molecular Design. J. Mater. Chem. 2008, 18, 3376–3384. [Google Scholar] [CrossRef]
  174. Zheng, C.; Zhao, W.; Wang, Z.; Huang, D.; Ye, J.; Ou, X.; Zhang, X.; Lee, C.; Lee, S. Highly Efficient Non-Doped Deep-Blue Organic Light-Emitting Diodes Based on Anthracene Derivatives. J. Mater. Chem. 2010, 20, 1560–1566. [Google Scholar] [CrossRef]
  175. Chem, J.M.; Kim, S.H.; Cho, I.; Sim, M.K.; Park, S.; Park, S.Y. Highly Efficient Deep-Blue Emitting Organic Light Emitting Diode Based on the Multifunctional Fluorescent Molecule Comprising Covalently Bonded Carbazole and Anthracene Moieties. J. Mater. Chem. 2011, 21, 9139–9148. [Google Scholar] [CrossRef]
  176. Zhuang, S.; Shangguan, R.; Jin, J.; Tu, G.; Wang, L.; Chen, J.; Ma, D.; Zhu, X. Efficient Nondoped Blue Organic Light-Emitting Diodes Based on Phenanthroimidazole-Substituted Anthracene Derivatives. Org. Electron. 2012, 13, 3050–3059. [Google Scholar] [CrossRef]
  177. Park, J. Highly Efficient Dual-Core Derivatives with EQEs as High as 8.38% at High Brightness for OLED Blue Emitters. J. Mater. Chem. 2019, 7, 14709–14716. [Google Scholar] [CrossRef]
  178. Lai, W.Y.; Zhu, R.; Fan, Q.L.; Hou, L.T.; Cao, Y.; Huang, W. Monodisperse Six-Armed Triazatruxenes: Microwave-Enhanced Synthesis and Highly Efficient Pure-Deep-Blue Electroluminescence. Macromolecules 2006, 39, 3707–3709. [Google Scholar] [CrossRef]
  179. Lai, W.Y.; He, Q.Y.; Zhu, R.; Chen, Q.Q.; Huang, W. Kinked Star-Shaped Fluorene/Triazatruxene Co-Oligomer Hybrids with Enhanced Functional Properties for High-Performance, Solution-Processed, Blue Organic Light-Emitting Diodes. Adv. Funct. Mater. 2008, 18, 265–276. [Google Scholar] [CrossRef]
  180. You, A.; Be, M.A.Y.; In, I. High Efficiency Deep-Blue Organic Light-Emitting Diode with a Blue Dye in Low-Polarity Host. Appl. Phys. Lett. 2008, 92, 193314. [Google Scholar] [CrossRef]
  181. Zhu, M.; Ye, T.; Li, C.G.; Cao, X.; Zhong, C.; Ma, D.; Qin, J.; Yang, C. Efficient Solution-Processed Nondoped Deep-Blue Organic Light-Emitting Diodes Based on Fluorene-Bridged Anthracene Derivatives Appended with Charge Transport Moieties. J. Phys. Chem. C 2011, 115, 17965–17972. [Google Scholar] [CrossRef]
  182. Huang, H.; Fu, Q.; Zhuang, S.; Liu, Y.; Wang, L.; Chen, J.; Ma, D.; Yang, C. Novel Deep Blue OLED Emitters with 1,3,5-Tri(Anthracen-10-Yl)Benzene-Centered Starburst Oligofluorenes. J. Phys. Chem. C 2011, 115, 4872–4878. [Google Scholar] [CrossRef]
  183. Liu, C.; Li, Y.; Zhang, Y.; Yang, C.; Wu, H.; Qin, J.; Cao, Y. Solution-Processed, Undoped, Deep-Blue Organic Light-Emitting Diodes Based on Starburst Oligofluorenes with a Planar Triphenylamine Core. Chemistry 2012, 18, 6928–6934. [Google Scholar] [CrossRef]
  184. Zhi, B.; Gao, Q.; Li, Z.H.; Xia, P.F.; Wong, M.S.; Cheah, K.W.; Chen, C.H. Efficient Deep-Blue Organic Light-Emitting Diodes: Arylamine-Substituted Oligofluorenes. Adv. Funct. Mater. 2007, 17, 3194–3199. [Google Scholar] [CrossRef]
  185. Zhang, T.; Wang, R.; Ren, H.; Chen, Z.; Li, J. Deep Blue Light-Emitting Polymers with Fluorinated Backbone for Enhanced Color Purity and Efficiency. Polymer 2012, 7, 1529–1534. [Google Scholar] [CrossRef]
  186. Meunmart, D.; Prachumrak, N.; Keawin, T.; Jungsuttiwong, S.; Sudyoadsuk, T.; Promark, V. Bis(4-Diphenylaminophenyl)Carbazole End-Capped Fluorene as Solution-Processed Deep-Blue Light-Emitting and Hole-Transporting Materials for Electroluminescent Devices. Tetrahedron Lett. 2012, 28, 3615–3618. [Google Scholar] [CrossRef]
  187. Liu, C.; Gu, Y.; Fu, Q.; Sun, N.; Zhong, C.; Ma, D.; Qin, J.; Yang, C. Nondoped Deep-Blue Organic Light-Emitting Diodes with Color Stability and Very Low Efficiency Roll-off: Solution-Processable Small-Molecule Fluorophores by Phosphine Oxide Linkage. Chem.-A Eur. J. 2012, 18, 13828–13835. [Google Scholar] [CrossRef] [PubMed]
  188. Liu, C.; Li, Y.; Li, Y.; Yang, C.; Wu, H.; Qin, J.; Cao, Y. Efficient Solution-Processed Deep-Blue Organic Light-Emitting Diodes Based on Multibranched Oligofluorenes with a Phosphine Oxide Center. Chem. Mater. 2013, 25, 3320–3327. [Google Scholar] [CrossRef]
  189. Giovanella, U.; Botta, C.; Galeotti, F.; Vercelli, B.; Battiato, S.; Pasini, M. Perfluorinated Polymer with Unexpectedly Efficient Deep Blue Electroluminescence for Full-Colour OLED Displays and Light Therapy Applications. J. Mater. Chem. C Mater. 2013, 1, 5322–5329. [Google Scholar] [CrossRef]
  190. McDowell, J.J.; Maier-Flaig, F.; Wolf, T.J.A.; Unterreiner, A.N.; Lemmer, U.; Ozin, G. Synthesis and Application of Photolithographically Patternable Deep Blue Emitting Poly(3,6-Dimethoxy-9,9-Dialkylsilafluorene)s. ACS Appl. Mater. Interfaces 2014, 6, 83. [Google Scholar] [CrossRef]
  191. Hu, L.; Liu, S.; Xie, G.; Yang, W.; Zhang, B. Bis(Benzothiophene-S,S-Dioxide) Fused Small Molecules Realize Solution-Processible, High-Performance and Non-Doped Blue Organic Light-Emitting Diodes. J. Mater. Chem. C Mater. 2020, 8, 1002–1009. [Google Scholar] [CrossRef]
  192. Liu, X.; Liu, W.; Dongyu, W.; Wei, X.; Wang, L.; Wang, H.; Miao, Y.; Xu, H.; Yu, J.; Xu, B. Deep-Blue Fluorescent Emitter Based on a 9,9-Dioctylfluorene Bridge with a Hybridized Local and Charge-Transfer Excited State for Organic Light-Emitting Devices with EQE Exceeding 8%. J. Mater. Chem. C. 2020, 8, 14117–14124. [Google Scholar] [CrossRef]
  193. Fischer, A.; Forget, S.; Chénais, S.; Castex, M.; Adès, D.; Siove, A.; Denis, C.; Maisse, P.; Geffroy, B. Highly Efficient Multilayer Organic Pure-Blue-Light Emitting Diodes with Substituted Carbazoles Compounds in the Emitting Layer. J. Phys. D Appl. Phys. 2006, 39, 917. [Google Scholar] [CrossRef]
  194. Tseng, R.J.; Chiechi, R.C.; Wudl, F.; Yang, Y. Highly Efficient 7,8,10-Triphenylfluoranthene-Doped Blue Organic Light-Emitting Diodes for Display Application. Appl. Phys. Lett. 2006, 88, 093512. [Google Scholar] [CrossRef]
  195. Jou, J.H.; Chiang, P.H.; Lin, Y.P.; Chang, C.Y.; Lai, C.L. Hole-Transporting-Layer-Free High-Efficiency Fluorescent Blue Organic Light-Emitting Diodes. Appl. Phys. Lett. 2007, 91, 043504. [Google Scholar] [CrossRef]
  196. Li, Z.; Jiao, B.; Wu, Z.; Liu, P.; Ma, L.; Lei, X.; Wang, D.; Zhou, G.; Hu, H.; Hou, X. Fluorinated 9,9′-Spirobifluorene Derivatives as Host Materials for Highly Efficient Blue Organic Light-Emitting Devices. J. Mater. Chem. C Mater. 2013, 1, 2183–2192. [Google Scholar] [CrossRef]
  197. Yu, J.; Chen, Z.; Sakuratani, Y.; Suzuki, H.; Tokita, M.; Miyata, S. A Novel Blue Light Emitting Diode Using Tris(2,3-Methyl-8-Hydroxyquinoline) Aluminum(III) as Emitter. Jpn. J. Appl. Phys. 1999, 38, 6762–6763. [Google Scholar] [CrossRef]
  198. Chan, L.; Lee, R.; Hsieh, C.; Yeh, H.; Chen, C. Optimization of High-Performance Blue Organic Light-Emitting Diodes Containing Tetraphenylsilane Molecular Glass Materials. J. Am. Chem. Soc. 2002, 78, 6469–6479. [Google Scholar] [CrossRef] [PubMed]
  199. Liu, D.; Du, M.; Chen, D.; Ye, K.; Zhang, Z.; Liu, Y.; Wang, Y. A Novel Tetraphenylsilane–Phenanthroimidazole Hybrid Host Material for Highly Efficient Blue Fluorescent, Green and Red Phosphorescent. J. Mater. Chem. C 2015, 3, 4394–4401. [Google Scholar] [CrossRef]
  200. Jou, J.H.; Li, J.L.; Sahoo, S.; Dubey, D.K.; Kumar Yadav, R.A.; Joseph, V.; Thomas, K.R.J.; Wang, C.W.; Jayakumar, J.; Cheng, C.H. Enabling a 6.5% External Quantum Efficiency Deep-Blue Organic Light-Emitting Diode with a Solution-Processable Carbazole-Based Emitter. J. Phys. Chem. C 2018, 122, 24295–24303. [Google Scholar] [CrossRef]
  201. Yang, J.; Liu, X.; Liu, Z.; Wang, L.; Sun, J.; Guo, Z.; Xu, H.; Wang, H.; Zhao, B.; Xie, G. Protonation-Induced Dual Fluorescence of a Blue Fluorescent Material with Twisted A–π–D–π–A Configuration. J. Mater. Chem. C Mater. 2020, 8, 2442–2450. [Google Scholar] [CrossRef]
  202. Wu, Q.; Braveenth, R.; Zhang, H.Q.; Bae, I.J.; Kim, M.; Chai, K.Y. Oxadiazole-Based Highly Efficient Bipolar Fluorescent Emitters for Organic Light-Emitting Diodes. Molecules 2018, 23, 843. [Google Scholar] [CrossRef] [PubMed]
  203. Wang, Y.; Huang, H.; Wang, Y.; Zhu, Q.; Qin, J. Simple Oxadiazole Derivatives as Deep Blue Fluorescent Emitter and Bipolar Host for Organic Light-Emitting Diodes. OptMa 2018, 84, 278–283. [Google Scholar] [CrossRef]
  204. Tang, C.; VanSlyke, S.A. Organic Electroluminescent Diodes. Appl. Phys. Lett. 1987, 51, 913–915. [Google Scholar] [CrossRef]
  205. Wu, C.-C.; Lin, Y.-T.; Wong, K.-T.; Chen, R.-T.; Chien, Y.-Y. Efficient Organic Blue-Light-Emitting Devices with Double Confinement on Terfluorenes with Ambipolar Carrier Transport Properties. Adv. Mater. 2004, 16, 61–65. [Google Scholar] [CrossRef]
  206. Lee, M.-T.; Liao, C.-H.; Tsai, C.-H.; Chen, H. Highly Efficient, Deep-blue Doped Organic Light-emitting Devices. Adv. Mater. 2002, 80, 2493–2497. [Google Scholar] [CrossRef]
  207. Baldo, M.; Thompson, M.; Forrest, S. High-Efficiency Fluorescent Organic Light-Emitting Devices Using a Phosphorescent Sensitizer. Nature 2000, 403, 750–753. [Google Scholar] [CrossRef]
  208. Kido, J.; Kimura, M.; Science, K.N. Multilayer White Light-Emitting Organic Electroluminescent Device. Science 1995, 267, 1332–1334. [Google Scholar] [CrossRef]
  209. Benmansour, B.H.; Shioya, T.; Sato, Y.; Bazan, G.C. Anthracene-Containing Binaphthol Chromophores for Light-Emitting Diode (LED) Fabrication. Adv. Funct. Mater. 2003, 13, 883–886. [Google Scholar] [CrossRef]
  210. Holmes, R.J.; Forrest, S.R.; Sajoto, T.; Tamayo, A.; Djurovich, P.I.; Thompson, M.E.; Brooks, J.; Tung, Y.J.; D’Andrade, B.W.; Weaver, M.S.; et al. Saturated Deep Blue Organic Electrophosphorescence Using a Fluorine-Free Emitter. Appl. Phys. Lett. 2005, 87, 243507. [Google Scholar] [CrossRef]
  211. Xiaohui Yang, B.; Wang, Z.; Madakuni, S.; Li, J.; Jabbour, G.E.; Jabbour, G.E.; Yang, X.; Li, J.; Wang, Z.; Madakuni, S. Efficient Blue- and White-Emitting Electrophosphorescent Devices Based on Platinum(II) [1,3-Difluoro-4,6-Di(2-Pyridinyl)Benzene] Chloride. Adv. Mater. 2008, 20, 2405–2409. [Google Scholar] [CrossRef]
  212. Sasabe, H.; Takamatsu, J.I.; Motoyama, T.; Watanabe, S.; Wagenblast, G.; Langer, N.; Molt, O.; Fuchs, E.; Lennartz, C.; Kido, J. High-Efficiency Blue and White Organic Light-Emitting Devices Incorporating a Blue Iridium Carbene Complex. Adv. Mater. 2010, 22, 5003–5007. [Google Scholar] [CrossRef]
  213. Hang, X.-C.; Fleetham, T.; Turner, E.; Brooks, J.; Li, J.; Hang, X.; Fleetham, T.; Turner, E.; Li, J.; Brooks, J. Highly Efficient Blue-Emitting Cyclometalated Platinum(II) Complexes by Judicious Molecular Design. Angew. Chem. Int. Ed. 2013, 52, 6753–6756. [Google Scholar] [CrossRef]
  214. Ertl, C.D.; Cerdá, J.; Junquera-Hernández, J.M.; Pertegás, A.; Bolink, H.J.; Constable, E.C.; Neuburger, M.; Ortí, E.; Housecroft, C.E. Colour Tuning by the Ring Roundabout: [Ir(C^N)2(N^N)]+ Emitters with Sulfonyl-Substituted Cyclometallating Ligands. RSC Adv. 2015, 5, 42815–42827. [Google Scholar] [CrossRef]
  215. He, L.; Wang, Z.; Duan, L.; Yang, C.; Tang, R.; Song, X.; Pan, C. Toward Fluorine-Free Blue-Emitting Cationic Iridium Complexes: To Generate Emission from the Cyclometalating Ligands with Enhanced Triplet Energy. Dalton Trans. 2016, 45, 5604–5613. [Google Scholar] [CrossRef]
  216. Shavaleev, N.M.; Scopelliti, R.; Grätzel, M.; Nazeeruddin, M.K. Phosphorescent cationic iridium(III) complexes with cyclometalating 1H-indazole and 2H-[1,2,3]-triazole ligands. Inorganica Chim. Acta 2012, 388, 84–87. [Google Scholar] [CrossRef]
  217. Li, J.; Djurovich, P.I.; Alleyne, B.D.; Yousufuddin, M.; Ho, N.N.; Thomas, J.C.; Peters, J.C.; Bau, R.; Thompson, M.E. Synthetic Control of Excited-State Properties in Cyclometalated Ir(III) Complexes Using Ancillary Ligands. Inorg. Chem. 2005, 44, 1713–1727. [Google Scholar] [CrossRef] [PubMed]
  218. Holmes, R.J.; D’Andrade, B.W.; Forrest, S.R.; Ren, X.; Li, J.; Thompson, M.E. Efficient, Deep-Blue Organic Electrophosphorescence by Guest Charge Trapping. Appl. Phys. Lett. 2003, 83, 3818. [Google Scholar] [CrossRef]
  219. Lee, S.J.; Park, K.M.; Yang, K.; Kang, Y. Blue Phosphorescent Ir(III) Complex with High Color Purity: Fac-Tris(2′,6′-Difluoro-2,3′-Bipyridinato-N,C 4′)Iridium(III). Inorg. Chem. 2009, 48, 1030–1037. [Google Scholar] [CrossRef] [PubMed]
  220. Kang, Y.; Chang, Y.L.; Lu, J.S.; Ko, S.B.; Rao, Y.; Varlan, M.; Lu, Z.H.; Wang, S. Highly Efficient Blue Phosphorescent and Electroluminescent Ir(III) Compounds. J. Mater. Chem. C Mater. 2012, 1, 441–450. [Google Scholar] [CrossRef]
  221. Hsieh, C.H.; Wu, F.I.; Fan, C.H.; Huang, M.J.; Lu, K.Y.; Chou, P.Y.; Yang, Y.H.O.; Wu, S.H.; Chen, I.C.; Chou, S.H.; et al. Design and Synthesis of Iridium Bis(Carbene) Complexes for Efficient Blue Electrophosphorescence. Chem.-A Eur. J. 2011, 17, 9180–9187. [Google Scholar] [CrossRef]
  222. Fan, C.; Li, Y.; Yang, C.; Wu, H.; Qin, J.; Cao, Y. Phosphoryl/Sulfonyl-Substituted Iridium Complexes as Blue Phosphorescent Emitters for Single-Layer Blue and White Organic Light-Emitting Diodes by Solution Process. Chem. Mater. 2012, 24, 4581–4587. [Google Scholar] [CrossRef]
  223. Fan, C.; Zhu, L.; Jiang, B.; Zhong, C.; Ma, D.; Qin, J.; Yang, C. Efficient Blue and Bluish-Green Iridium Phosphors: Fine-Tuning Emissions of FIrpic by Halogen Substitution on Pyridine-Containing Ligands. Org. Electron. 2013, 14, 3163–3171. [Google Scholar] [CrossRef]
  224. Kessler, F.; Watanabe, Y.; Sasabe, H.; Katagiri, H.; Nazeeruddin, M.K.; Grätzel, M.; Kido, J. High-Performance Pure Blue Phosphorescent OLED Using a Novel Bis-Heteroleptic Iridium(III) Complex with Fluorinated Bipyridyl Ligands. J. Mater. Chem. C Mater. 2013, 1, 1070–1075. [Google Scholar] [CrossRef]
  225. Park, H.J.; Kim, J.N.; Yoo, H.J.; Wee, K.R.; Kang, S.O.; Cho, D.W.; Yoon, U.C. Rational Design, Synthesis, and Characterization of Deep Blue Phosphorescent Ir(III) Complexes Containing (4′-Substituted-2′-Pyridyl)-1,2,4-Triazole Ancillary Ligands. J. Org. Chem. 2013, 78, 8054–8064. [Google Scholar] [CrossRef] [PubMed]
  226. Lee, S.; Kim, S.O.; Shin, H.; Yun, H.J.; Yang, K.; Kwon, S.K.; Kim, J.J.; Kim, Y.H. Deep-Blue Phosphorescence from Perfluoro Carbonyl-Substituted Iridium Complexes. J. Am. Chem. Soc. 2013, 135, 14321–14328. [Google Scholar] [CrossRef]
  227. Liu, L.; Liu, X.; Wu, K.; Ding, J.; Zhang, B.; Xie, Z.; Wang, L. Efficient Solution-Processed Blue Phosphorescent Organic Light-Emitting Diodes with Halogen-Free Solvent to Optimize the Emissive Layer Morphology. Org. Electron. 2014, 15, 1401–1406. [Google Scholar] [CrossRef]
  228. Sun, D.; Zhou, X.; Li, H.; Sun, X.; Ren, Z.; Yan, S. Multi-3,3 ′ -Bicarbazole-Substituted Arylsilane Host Materials with Balanced Charge Transport for Highly E Ffi Cient Solution-Processed Blue Phosphorescent Organic Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2015, 32, 17802–17810. [Google Scholar] [CrossRef] [PubMed]
  229. Byeon, S.Y.; Choi, J.M.; Lee, J.Y. Dyes and Pigments Pyridoindole Based Intramolecular Charge Transfer Type Host Material for Blue Phosphorescent Organic Light-Emitting Diodes. Dye. Pigment. 2016, 134, 285–290. [Google Scholar] [CrossRef]
  230. Park, S.R.; Kim, S.M.; Kang, J.H.; Lee, J.H.; Suh, M.C. Bipolar Host Materials with Carbazole and Dipyridylamine Groups Showing High Triplet Energy for Blue Phosphorescent Organic Light Emitting Diodes. Dye. Pigment. 2017, 141, 217–224. [Google Scholar] [CrossRef]
  231. Yuan, J.; Jiang, H.; Yang, Q.; Xiang, Y.; Zhang, Y.; Dai, Y.; Li, P.; Zheng, C.; Xie, G.; Chen, R. A solution-processable wholly-aromatic bipolar host material for highly efficient blue electroluminescent devices. J. Mater. Chem. C 2021, 9, 687–692. [Google Scholar] [CrossRef]
  232. Kang, J.; Zaen, R.; Park, K.M.; Lee, K.H.; Lee, J.Y.; Kang, Y. Cyclometalated Platinum(II) β-Diketonate Complexes as Single Dopants for High-Efficiency White OLEDs: The Relationship between Intermolecular Interactions in the Solid State and Electroluminescent Efficiency. Cryst. Growth Des. 2020, 20, 6129–6138. [Google Scholar] [CrossRef]
  233. Liao, K.Y.; Hsu, C.W.; Chi, Y.; Hsu, M.K.; Wu, S.W.; Chang, C.H.; Liu, S.H.; Lee, G.H.; Chou, P.T.; Hu, Y.; et al. Pt(II) Metal Complexes Tailored with a Newly Designed Spiro-Arranged Tetradentate Ligand; Harnessing of Charge-Transfer Phosphorescence and Fabrication of Sky Blue and White OLEDs. Inorg. Chem. 2015, 54, 4029–4038. [Google Scholar] [CrossRef]
  234. Liu, Q.D.; Wang, R.; Wang, S. Blue Phosphorescent Zn(II) and Orange Phosphorescent Pt(II) Complexes of 4,4′-Diphenyl-6,6′-Dimethyl-2,2′-Bipyrimidine. Dalton Trans. 2004, 2073–2079. [Google Scholar] [CrossRef]
  235. Xu, H.; Xu, Z.F.; Yue, Z.Y.; Yan, P.F.; Wang, B.; Jia, L.W.; Li, G.M.; Sun, W.B.; Zhang, J.W. A Novel Deep Blue-Emitting ZnII Complex Based on Carbazole-Modified 2-(2-Hydroxyphenyl)Benzimidazole: Synthesis, Bright Electroluminescence, and Substitution Effect on Photoluminescent, Thermal, and Electrochemical Properties. J. Phys. Chem. C 2008, 112, 15517–15525. [Google Scholar] [CrossRef]
  236. Oh, C.S.; Lee, J.Y. High Triplet Energy Zn Complexes as Host Materials for Green and Blue Phosphorescent Organic Light-Emitting Diodes. Dye. Pigment. 2013, 99, 374–377. [Google Scholar] [CrossRef]
  237. Soo Yook, K.; Yeob Lee, J.; Yook, K.S.; Lee, J.Y. Organic Materials for Deep Blue Phosphorescent Organic Light-emitting Diodes. Adv. Mater. 2012, 24, 3169–3190. [Google Scholar] [CrossRef] [PubMed]
  238. Zhu, M.; Yang, C. Blue Fluorescent Emitters: Design Tactics and Applications in Organic Light-Emitting Diodes. Chem. Soc. Rev. 2013, 42, 4963–4976. [Google Scholar] [CrossRef] [PubMed]
  239. Endo, A.; Ogasawara, M.; Takahashi, A.; Yokoyama, D.; Kato, Y.; Adachi, C.; Adachi, C.; Endo, A.; Yokoyama, D.; Ogasawara, M.; et al. Thermally Activated Delayed Fluorescence from Sn4+–Porphyrin Complexes and Their Application to Organic Light Emitting Diodes—A Novel Mechanism for Electroluminescence. Adv. Mater. 2009, 21, 4802–4806. [Google Scholar] [CrossRef] [PubMed]
  240. Cui, L.-S.; Nomura, H.; Kim, J.U.; Nakanotani, H.; Adachi, C.; Cui, L.; Geng, Y.; Kim, J.U.; Nakanotani, H.; Adachi, C.; et al. Controlling Singlet–Triplet Energy Splitting for Deep-Blue Thermally Activated Delayed Fluorescence Emitters. Angew. Chem. Int. Ed. 2017, 56, 1571–1575. [Google Scholar] [CrossRef]
  241. Cai, X.; Su, S.-J.; Cai, X.Y.; 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]
  242. Su, S.J.; Chiba, T.; Takeda, T.; Kido, J. Pyridine-Containing Triphenylbenzene Derivatives with High Electron Mobility for Highly Efficient Phosphorescent OLEDs. Adv. Mater. 2008, 20, 2125–2130. [Google Scholar] [CrossRef]
  243. Cui, L.S.; Deng, Y.L.; Tsang, D.P.K.; Jiang, Z.Q.; Zhang, Q.; Liao, L.S.; Adachi, C. Controlling Synergistic Oxidation Processes for Efficient and Stable Blue Thermally Activated Delayed Fluorescence Devices. Adv. Mater. 2016, 28, 7620–7625. [Google Scholar] [CrossRef]
  244. Tao, Y.; Yuan, K.; Chen, T.; Xu, P.; Li, H.; Chen, R.; Zheng, C.; Zhang, L.; Huang, W. Thermally Activated Delayed Fluorescence Materials towards the Breakthrough of Organoelectronics. Adv. Mater. 2014, 26, 7931–7958. [Google Scholar] [CrossRef]
  245. Shin, H.; Lee, S.; Kim, K.H.; Moon, C.K.; Yoo, S.J.; Lee, J.H.; Kim, J.J. Blue Phosphorescent Organic Light-Emitting Diodes Using an Exciplex Forming Co-Host with the External Quantum Efficiency of Theoretical Limit. Adv. Mater. 2014, 26, 4730–4734. [Google Scholar] [CrossRef]
  246. Hirata, S.; Sakai, Y.; Masui, K.; Tanaka, H.; Lee, S.Y.; Nomura, H.; Nakamura, N.; Yasumatsu, M.; Nakanotani, H.; Zhang, Q.; et al. Highly Efficient Blue Electroluminescence Based on Thermally Activated Delayed Fluorescence. Nat. Mater. 2014, 14, 330–336. [Google Scholar] [CrossRef]
  247. Li, W.; Pan, Y.; Xiao, R.; Peng, Q.; Zhang, S.; Ma, D.; Li, F.; Shen, F.; Wang, Y.; Yang, B.; et al. Employing ~100% Excitons in OLEDs by Utilizing a Fluorescent Molecule with Hybridized Local and Charge-Transfer Excited State. Adv. Funct. Mater. 2014, 24, 1609–1614. [Google Scholar] [CrossRef]
  248. Zhang, Q.; Li, B.; Huang, S.; Nomura, H.; Tanaka, H.; Adachi, C. Efficient Blue Organic Light-Emitting Diodes Employing Thermally Activated Delayed Fluorescence. Nat. Photonics 2014, 8, 326–332. [Google Scholar] [CrossRef]
  249. Zhang, Q.; Tsang, D.; Kuwabara, H.; Hatae, Y.; Li, B.; Takahashi, T.; Youn Lee, S.; Yasuda, T.; Adachi, C.; Zhang, Q.; et al. Nearly 100% Internal Quantum Efficiency in Undoped Electroluminescent Devices Employing Pure Organic Emitters. Adv. Mater. 2015, 27, 2096–2100. [Google Scholar] [CrossRef] [PubMed]
  250. Sasabe, H.; Seino, Y.; Kimura, M.; Kido, J. A m -Terphenyl-Modifed Sulfone Derivative as a Host Material for High-Efficiency Blue and Green Phosphorescent OLEDs. Chem. Mater. 2012, 24, 1404–1406. [Google Scholar] [CrossRef]
  251. Jeon, S.O.; Earmme, T.; Jenekhe, S.A. New Sulfone-Based Electron-Transport Materials with High Triplet Energy for Highly Efficient Blue Phosphorescent Organic Light-Emitting Diodes. J. Mater. Chem. C Mater. 2014, 2, 10129–10137. [Google Scholar] [CrossRef]
  252. Duan, L.; Qiao, J.; Sun, Y.; Qiu, Y.; Duan, L.; Qiao, J.; Sun, Y.D.; Qiu, Y. Strategies to Design Bipolar Small Molecules for OLEDs: Donor-Acceptor Structure and Non-Donor-Acceptor Structure. Adv. Mater. 2011, 23, 1137–1144. [Google Scholar] [CrossRef]
  253. Hsu, F.M.; Chien, C.H.; Hsieh, Y.J.; Wu, C.H.; Shu, C.F.; Liu, S.W.; Chen, C.T. Highly Efficient Red Electrophosphorescent Device Incorporating a Bipolar Triphenylamine/Bisphenylsulfonyl-Substituted Fluorene Hybrid as the Host. J. Mater. Chem. 2009, 19, 8002–8008. [Google Scholar] [CrossRef]
  254. Zhang, Q.; Li, J.; Shizu, K.; Huang, S.; Hirata, S.; Miyazaki, H.; Adachi, C. Design of Efficient Thermally Activated Delayed Fluorescence Materials for Pure Blue Organic Light Emitting Diodes. J. Am. Chem. Soc. 2012, 134, 14706–14709. [Google Scholar] [CrossRef]
  255. Romain, M.; Tondelier, D.; Geffroy, B.; Shirinskaya, A.; Jeannin, O.; Rault-Berthelot, J.; Poriel, C. Spiro-Configured Phenyl Acridine Thioxanthene Dioxide as a Host for Efficient PhOLEDs. Chem. Commun. 2014, 51, 1313–1315. [Google Scholar] [CrossRef]
  256. Lee, C.W.; Lee, J.Y. A Hole Transport Material with Ortho-Linked Terphenyl Core Structure for High Power Efficiency in Blue Phosphorescent Organic Light-Emitting Diodes. Org. Electron. 2014, 15, 399–404. [Google Scholar] [CrossRef]
  257. Seo, J.A.; Gong, M.S.; Lee, J.Y. Thermally Stable Indoloacridine Type Host Material for High Efficiency Blue Phosphorescent Organic Light-Emitting Diodes. Org. Electron. 2014, 15, 3773–3779. [Google Scholar] [CrossRef]
  258. Park, M.S.; Lee, J.Y. High Efficiency Green and Blue Phosphorescent Organic Light-Emitting Diodes Using Pyrroloacridine Type Hole Transport Material. Mol. Cryst. Liq. Cryst. 2013, 584, 145–152. [Google Scholar] [CrossRef]
  259. Kim, M.; Lee, J.Y. Improved Power Efficiency in Deep Blue Phosphorescent Organic Light-Emitting Diodes Using an Acridine Core Based Hole Transport Material. Org. Electron. 2012, 13, 1245–1249. [Google Scholar] [CrossRef]
  260. Kim, M.; Lee, J.Y. Pyridine-Modified Acridine-Based Bipolar Host Material for Green Phosphorescent Organic Light-Emitting Diodes. Chem. Asian J. 2012, 7, 899–902. [Google Scholar] [CrossRef] [PubMed]
  261. Park, M.S.; Lee, J.Y. Indolo Acridine-Based Hole-Transport Materials for Phosphorescent OLEDs with over 20% External Quantum Efficiency in Deep Blue and Green. Chem. Mater. 2011, 23, 4338–4343. [Google Scholar] [CrossRef]
  262. Seo, J.A.; Jeon, S.K.; Gong, M.S.; Lee, J.Y.; Noh, C.H.; Kim, S.H. Long Lifetime Blue Phosphorescent Organic Light-Emitting Diodes with an Exciton Blocking Layer. J. Mater. Chem. C Mater. 2015, 3, 4640–4645. [Google Scholar] [CrossRef]
  263. Nakanotani, H.; Higuchi, T.; Furukawa, T.; Masui, K.; Morimoto, K.; Numata, M.; Tanaka, H.; Sagara, Y.; Yasuda, T.; Adachi, C. High-Efficiency Organic Light-Emitting Diodes with Fluorescent Emitters. Nat. Commun. 2014, 5, 4016. [Google Scholar] [CrossRef]
  264. Nakao, K.; Sasabe, H.; Komatsu, R.; Hayasaka, Y.; Ohsawa, T.; Kido, J.; Nakao, K.; Sasabe, H.; Komatsu, R.; Hayasaka, Y.; et al. Significant Enhancement of Blue OLED Performances through Molecular Engineering of Pyrimidine-Based Emitter. Adv. Opt. Mater. 2017, 5, 1600843. [Google Scholar] [CrossRef]
  265. Grimsdale, A.C.; Jacob, J. Polymer-Based LEDs and Solar Cells. Polym. Sci. A Compr. Ref. 2012, 8, 261–282. [Google Scholar] [CrossRef]
  266. Liou, G.S.; Yen, H.J. Polyimides. Polym. Sci. A Compr. Ref. 2012, 5, 497–535. [Google Scholar] [CrossRef]
  267. Wu, S.; Aonuma, M.; Zhang, Q.; Huang, S.; Nakagawa, T.; Kuwabara, K.; Adachi, C. High-Efficiency Deep-Blue Organic Light-Emitting Diodes Based on a Thermally Activated Delayed Fluorescence Emitter. J. Mater. Chem. C Mater. 2013, 2, 421–424. [Google Scholar] [CrossRef]
  268. High-Efficiency Diphenylsulfon Derivative-Based Organic Light-Emitting Diode Exhibiting Thermally-Activated Delayed Fluorescence|Request PDF. Available online: https://www.researchgate.net/publication/303448146_High-efficiency_diphenylsulfon_derivative-based_organic_light-emitting_diode_exhibiting_thermally-activated_delayed_fluorescence (accessed on 15 May 2022).
  269. Serevičius, T.; Nakagawa, T.; Kuo, M.C.; Cheng, S.H.; Wong, K.T.; Chang, C.H.; Kwong, R.C.; Xia, S.; Adachi, C. Enhanced Electroluminescence Based on Thermally Activated Delayed Fluorescence from a Carbazole–Triazine Derivative. Phys. Chem. Chem. Phys. 2013, 15, 15850–15855. [Google Scholar] [CrossRef] [PubMed]
  270. Kim, M.; Kyu Jeon, S.; Hwang, S.-H.; Yeob Lee, J.; Kim, M.; Hwang, S.; Jeon, S.K.; Lee, J.Y. Stable Blue Thermally Activated Delayed Fluorescent Organic Light-Emitting Diodes with Three Times Longer Lifetime than Phosphorescent Organic Light-Emitting Diodes. Adv. Mater. 2015, 27, 2515–2520. [Google Scholar] [CrossRef]
  271. 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] [PubMed]
  272. Cho, Y.J.; Yook, K.S.; Lee, J.Y. A Universal Host Material for High External Quantum Efficiency Close to 25% and Long Lifetime in Green Fluorescent and Phosphorescent OLEDs. Adv. Mater. 2014, 26, 4050–4055. [Google Scholar] [CrossRef]
  273. Cho, Y.J.; Yook, K.S.; Lee, J.Y. Cool and Warm Hybrid White Organic Light-Emitting Diode with Blue Delayed Fluorescent Emitter Both as Blue Emitter and Triplet Host. Sci. Rep. 2015, 5, 7859. [Google Scholar] [CrossRef]
  274. Lee, D.R.; Hwang, S.H.; Jeon, S.K.; Lee, C.W.; Lee, J.Y. Benzofurocarbazole and Benzothienocarbazole as Donors for Improved Quantum Efficiency in Blue Thermally Activated Delayed Fluorescent Devices. Chem. Commun. 2015, 51, 8105–8107. [Google Scholar] [CrossRef]
  275. Yi, C.L.; Ko, C.L.; Yeh, T.C.; Chen, C.Y.; Chen, Y.S.; Chen, D.G.; Chou, P.T.; Hung, W.Y.; Wong, K.T. Harnessing a New Co-Host System and Low Concentration of New TADF Emitters Equipped with Trifluoromethyl- and Cyano-Substituted Benzene as Core for High-Efficiency Blue OLEDs. ACS Appl. Mater. Interfaces 2020, 12, 2724–2732. [Google Scholar] [CrossRef]
  276. Numata, M.; Yasuda, T.; Adachi, C. High Efficiency Pure Blue Thermally Activated Delayed Fluorescence Molecules Having 10 H -Phenoxaborin and Acridan Units. Chem. Commun. 2015, 51, 9443–9446. [Google Scholar] [CrossRef]
  277. Tsai, W.L.; Huang, M.H.; Lee, W.K.; Hsu, Y.J.; Pan, K.C.; Huang, Y.H.; Ting, H.C.; Sarma, M.; Ho, Y.Y.; Hu, H.C.; et al. A Versatile Thermally Activated Delayed Fluorescence Emitter for Both Highly Efficient Doped and Non-Doped Organic Light Emitting Devices. Chem. Commun. 2015, 51, 13662–13665. [Google Scholar] [CrossRef]
  278. Li, W.; Li, W.; Gan, L.; Li, M.; Zheng, N.; Ning, C.; Chen, D.; Wu, Y.C.; Su, S.J. J-Aggregation Enhances the Electroluminescence Performance of a Sky-Blue Thermally Activated Delayed-Fluorescence Emitter in Nondoped Organic Light-Emitting Diodes. ACS Appl. Mater. Interfaces 2020, 12, 2717–2723. [Google Scholar] [CrossRef] [PubMed]
  279. Yu, Y.J.; Wang, X.Q.; Liu, J.F.; Jiang, Z.Q.; Liao, L.S. Harvesting Triplet Excitons for Near-Infrared Electroluminescence via Thermally Activated Delayed Fluorescence Channel. iScience 2021, 24, 102123. [Google Scholar] [CrossRef] [PubMed]
  280. Evaristo De Sousa, L.; Dos, L.; Born, S.; Henrique De Oliveira Neto, P.; De Silva, P. Triplet-to-Singlet Exciton Transfer in Hyperfluorescent OLED Materials. J. Mater. Chem. C Mater. 2022, 10, 4914–4922. [Google Scholar] [CrossRef]
  281. Dias, F.B.; Bourdakos, K.N.; Jankus, V.; Moss, K.C.; Kamtekar, K.T.; Bhalla, V.; Santos, J.; Bryce, M.R.; Monkman, A.P. Triplet Harvesting with 100% Efficiency by Way of Thermally Activated Delayed Fluorescence in Charge Transfer OLED Emitters. Adv. Mater. 2013, 25, 3707–3714. [Google Scholar] [CrossRef]
  282. Li, Z.W.; Peng, L.Y.; Song, X.F.; Chen, W.K.; Gao, Y.J.; Fang, W.H.; Cui, G. Room-Temperature Phosphorescence and Thermally Activated Delayed Fluorescence in the Pd Complex: Mechanism and Dual Upconversion Channels. J. Phys. Chem. Lett. 2021, 12, 5944–5950. [Google Scholar] [CrossRef] [PubMed]
  283. Mamada, M.; Inada, K.; Komino, T.; Potscavage, W.J.; Nakanotani, H.; Adachi, C. Highly Efficient Thermally Activated Delayed Fluorescence from an Excited-State Intramolecular Proton Transfer System. ACS Cent. Sci. 2017, 3, 769–777. [Google Scholar] [CrossRef]
  284. Madayanad Suresh, S.; Hall, D.; Beljonne, D.; Olivier, Y.; Zysman-Colman, E.; Madayanad Suresh, S.; Hall, D.; Zysman-Colman, E.; Beljonne, D.; Olivier, Y. Multiresonant Thermally Activated Delayed Fluorescence Emitters Based on Heteroatom-Doped Nanographenes: Recent Advances and Prospects for Organic Light-Emitting Diodes. Adv. Funct. Mater. 2020, 30, 1908677. [Google Scholar] [CrossRef]
  285. Weissenseel, S.; Gottscholl, A.; Bönnighausen, R.; Dyakonov, V.; Sperlich, A.; Maximilian, J. Long-Lived Spin-Polarized Intermolecular Exciplex States in Thermally Activated Delayed Fluorescence-Based Organic Light-Emitting Diodes. Sci. Adv. 2021, 7, eabj9961. [Google Scholar] [CrossRef]
  286. Sakai, Y.; Sagara, Y.; Nomura, H.; Nakamura, N.; Suzuki, Y.; Miyazaki, H.; Adachi, C. Zinc Complexes Exhibiting Highly Efficient Thermally Activated Delayed Fluorescence and Their Application to Organic Light-Emitting Diodes. Chem. Commun. 2015, 51, 3181–3184. [Google Scholar] [CrossRef]
  287. Kanno, H.; Ishikawa, K.; Nishio, Y.; Endo, A.; Adachi, C.; Shibata, K. Highly Efficient and Stable Red Phosphorescent Organic Light-Emitting Device Using Bis[2-(2-Benzothiazoyl)Phenolato]Zinc(II) as Host Material. Appl. Phys. Lett. 2007, 90, 123509. [Google Scholar] [CrossRef]
  288. Hamada, Y.; Sano, T.; Fujii, H.; Nishio, Y.; Takahashi, H.; Shibata, K. White-Light-Emitting Material for Organic Electroluminescent Devices. Jpn. J. Appl. Phys. 1996, 35, L1339. [Google Scholar] [CrossRef]
  289. Norikazu, N.; Shinichi, W.; Keiichi, M.; Tsuneo, F. A Novel Blue Light Emitting Material Prepared from 2-(o-Hydroxyphenyl)Benzoxazole. Chem. Lett. 2006, 23, 1741–1742. [Google Scholar] [CrossRef]
  290. Osawa, M. Highly Efficient Blue-Green Delayed Fluorescence from Copper(i) Thiolate Complexes: Luminescence Color Alteration by Orientation Change of the Aryl Ring. Chem. Commun. 2014, 50, 1801–1803. [Google Scholar] [CrossRef]
  291. Kang, L.; Chen, J.; Teng, T.; Chen, X.L.; Yu, R.; Lu, C.Z. Experimental and Theoretical Studies of Highly Emissive Dinuclear Cu(i) Halide Complexes with Delayed Fluorescence. Dalton Trans. 2015, 44, 11649–11659. [Google Scholar] [CrossRef]
  292. Zhang, Q.; Chen, J.; Wu, X.Y.; Chen, X.L.; Yu, R.; Lu, C.Z. Outstanding Blue Delayed Fluorescence and Significant Processing Stability of Cuprous Complexes with Functional Pyridine–Pyrazolate Diimine Ligands. Dalton Trans. 2015, 44, 6706–6710. [Google Scholar] [CrossRef]
  293. Zink, D.M.; Bächle, M.; Baumann, T.; Nieger, M.; Kühn, M.; Wang, C.; Klopper, W.; Monkowius, U.; Hofbeck, T.; Yersin, H.; et al. Synthesis, Structure, and Characterization of Dinuclear Copper(I) Halide Complexes with P^N Ligands Featuring Exciting Photoluminescence Properties. Inorg. Chem. 2013, 52, 2292–2305. [Google Scholar] [CrossRef] [PubMed]
  294. Deaton, J.C.; Switalski, S.C.; Kondakov, D.Y.; Young, R.H.; Pawlik, T.D.; Giesen, D.J.; Harkins, S.B.; Miller, A.J.M.; Mickenberg, S.F.; Peters, J.C. E-Type Delayed Fluorescence of a Phosphine-Supported Cu 2(μ-Nar 2) 2 Diamond Core: Harvesting Singlet and Triplet Excitons in OLEDs. J. Am. Chem. Soc. 2010, 132, 9499–9508. [Google Scholar] [CrossRef] [PubMed]
  295. Igawa, S.; Hashimoto, M.; Kawata, I.; Yashima, M.; Hoshino, M.; Osawa, M. Highly Efficient Green Organic Light-Emitting Diodes Containing Luminescent Tetrahedral Copper(i) Complexes. J. Mater. Chem. C Mater. 2012, 1, 542–551. [Google Scholar] [CrossRef]
  296. Osawa, M.; Kawata, I.; Ishii, R.; Igawa, S.; Hashimoto, M.; Hoshino, M. Application of Neutral d 10 Coinage Metal Complexes with an Anionic Bidentate Ligand in Delayed Fluorescence-Type Organic Light-Emitting Diodes. J. Mater. Chem. C Mater. 2013, 1, 4375–4383. [Google Scholar] [CrossRef]
  297. Volz, D.; Zink, D.M.; Bocksrocker, T.; Friedrichs, J.; Nieger, M.; Baumann, T.; Lemmer, U.; Bräse, S. Molecular Construction Kit for Tuning Solubility, Stability and Luminescence Properties: Heteroleptic MePyrPHOS-Copper Iodide-Complexes and Their Application in Organic Light-Emitting Diodes. Chem. Mater. 2013, 25, 3414–3426. [Google Scholar] [CrossRef]
  298. Ma, Y.; Che, C.M.; Chao, H.Y.; Zhou, X.; Chan, W.H.; Shen, J. High Luminescence Gold(I) and Copper(I) Complexes with a Triplet Excited State for Use in Light-Emitting Diodes. Adv. Mater. 1999, 11, 852–857. [Google Scholar] [CrossRef]
  299. Ma, Y.G.; Chan, W.H.; Zhou, X.M.; Che, C.M. Light-Emitting Diode Device from a Luminescent Organocopper(I) Compound. New J. Chem. 1999, 23, 263–265. [Google Scholar] [CrossRef]
  300. Cu(I) Complexes–Thermally Activated Delayed Fluorescence. Photophysical Approach and Material Design|Request PDF. Available online: https://www.researchgate.net/publication/304779001_CuI_complexes_-_Thermally_activated_delayed_fluorescence_Photophysical_approach_and_material_design (accessed on 17 May 2022).
  301. Tomé, L.I.N.; Domínguez-Pérez, M.; Cláudio, A.F.M.; Freire, M.G.; Marrucho, I.M.; Oscar Cabeza, O.C.; Coutinho, J.A.P. On the Interactions between Amino Acids and Ionic Liquids in Aqueous Media. J. Phys. Chem. B 2009, 113, 13971–13979. [Google Scholar] [CrossRef] [PubMed]
  302. Housecroft, C.E.; Constable, E.C. TADF: Enabling Luminescent Copper(I) Coordination Compounds for Light-Emitting Electrochemical Cells. J. Mater. Chem. C Mater. 2022, 10, 4456–4482. [Google Scholar] [CrossRef] [PubMed]
  303. Teng, T.; Xiong, J.; Cheng, G.; Zhou, C.; Lv, X.; Li, K. Solution-Processed Oleds Based on Thermally Activated Delayed Fluorescence Copper(I) Complexes with Intraligand Charge-Transfer Excited State. Molecules 2021, 26, 1125. [Google Scholar] [CrossRef]
  304. Lv, L.L.; Yuan, K.; Wang, Y.C. Theoretical Studying of Basic Photophysical Processes in a Thermally Activated Delayed Fluorescence Copper(I) Complex: Determination of Reverse Intersystem Crossing and Radiative Rate Constants. Org. Electron. 2017, 51, 207–219. [Google Scholar] [CrossRef]
  305. Czerwieniec, R.; Leitl, M.J.; Homeier, H.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]
  306. Czerwieniec, R.; Yersin, H. Diversity of Copper(I) Complexes Showing Thermally Activated Delayed Fluorescence: Basic Photophysical Analysis. Inorg. Chem. 2015, 54, 4322–4327. [Google Scholar] [CrossRef]
  307. Linfoot, C.L.; Leitl, M.J.; Richardson, P.; Rausch, A.F.; Chepelin, O.; White, F.J.; Yersin, H.; Robertson, N. Thermally Activated Delayed Fluorescence (TADF) and Enhancing Photoluminescence Quantum Yields of [CuI(Diimine)(Diphosphine)]+ Complexes-Photophysical, Structural, and Computational Studies. Inorg. Chem. 2014, 53, 10854–10861. [Google Scholar] [CrossRef]
  308. Xu, H.; Yang, T.; Wang, F.; Zhang, J.; Zhang, X.; Wang, H.; Xu, B. Thermally Activated Delayed Fluorescence of Copper(I) Complexes Using N, N′-Heteroaromatic of 2-(5-Phenyl-1,2,3-Triazole)Pyridine as Ligand. J. Lumin. 2019, 205, 82–86. [Google Scholar] [CrossRef]
  309. Stoïanov, A.; Gourlaouen, C.; Vela, S.; Daniel, C. Luminescent Dinuclear Copper(I) Complexes as Potential Thermally Activated Delayed Fluorescence (TADF) Emitters: A Theoretical Study. J. Phys. Chem. A 2018, 122, 1413–1421. [Google Scholar] [CrossRef] [PubMed]
  310. 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]
  311. Leitl, M.J.; Krylova, V.A.; Djurovich, P.I.; Thompson, M.E.; Yersin, H. Phosphorescence versus Thermally Activated Delayed Fluorescence. Controlling Singlet-Triplet Splitting in Brightly Emitting and Sublimable Cu(I) Compounds. J. Am. Chem. Soc. 2014, 136, 16032–16038. [Google Scholar] [CrossRef] [PubMed]
  312. Chen, X.L.; Yu, R.; Zhang, Q.K.; Zhou, L.J.; Wu, X.Y.; Zhang, Q.; Lu, C.Z. Rational Design of Strongly Blue-Emitting Cuprous Complexes with Thermally Activated Delayed Fluorescence and Application in Solution-Processed OLEDS. Chem. Mater. 2013, 25, 3910–3920. [Google Scholar] [CrossRef]
  313. 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]
  314. Lee, C.F.; Chin, K.F.; Peng, S.M.; Che, C.M. A Luminescent Tetrameric Zinc(II) Complex Containing the 7-Azaindolate Ligand. Photophysical Properties and Crystal Structure. J. Chem. Soc. Dalton Trans. 1993, 3, 467–470. [Google Scholar] [CrossRef]
  315. Yu, G.; Yin, S.; Liu, Y.; Shuai, Z.; Zhu, D. Structures, Electronic States, and Electroluminescent Properties of a Zinc(II) 2-(2-Hydroxyphenyl)Benzothiazolate Complex. J. Am. Chem. Soc. 2003, 125, 14816–14824. [Google Scholar] [CrossRef]
  316. Xu, X.; Liao, Y.; Yu, G.; You, H.; Di, C.; Su, Z.; Ma, D.; Wang, Q.; Li, S.; Wang, S.; et al. Charge Carrier Transporting, Photoluminescent, and Electroluminescent Properties of Zinc(II)-2-(2-Hydroxyphenyl)Benzothiazolate Complex. Chem. Mater. 2007, 19, 1740–1748. [Google Scholar] [CrossRef]
  317. Singh, S.P.; Mohapatra, Y.N.; Qureshi, M.; Manoharan, S.S. White Organic Light-Emitting Diodes Based on Spectral Broadening in Electroluminescence Due to Formation of Interfacial Exciplexes. Appl. Phys. Lett. 2005, 86, 113505. [Google Scholar] [CrossRef]
  318. Hao, Y.; Meng, W.; Xu, H.; Wang, H.; Liu, X.; Xu, B. White Organic Light-Emitting Diodes Based on a Novel Zn Complex with High CRI Combining Emission from Excitons and Interface-Formed Electroplex. Org. Electron. 2011, 12, 136–142. [Google Scholar] [CrossRef]
  319. Roh, S.G.; Kim, Y.H.; Seo, K.D.; Lee, D.H.; Kim, H.K.; Park, Y.I.; Park, J.W.; Lee, J.H. Synthesis, Photophysical, and Electroluminescent Device Properties of Zn(II)-Chelated Complexes Based on Functionalized Benzothiazole Derivatives. Adv. Funct. Mater. 2009, 19, 1663–1671. [Google Scholar] [CrossRef]
  320. Li, Z.; Dellali, A.; Malik, J.; Motevalli, M.; Nix, R.M.; Olukoya, T.; Peng, Y.; Ye, H.; Gillin, W.P.; Hernández, I.; et al. Luminescent Zinc(II) Complexes of Fluorinated Benzothiazol-2-Yl Substituted Phenoxide and Enolate Ligands. Inorg. Chem. 2013, 52, 1379–1387. [Google Scholar] [CrossRef] [PubMed]
  321. Wang, R.; Deng, L.; Fu, M.; Cheng, J.; Li, J. Novel Zn II Complexes of 2-(2-Hydroxyphenyl)Benzothiazoles Ligands: Electroluminescence and Application as Host Materials for Phosphorescent Organic Light-Emitting Diodes. J. Mater. Chem. 2012, 22, 23454–23460. [Google Scholar] [CrossRef]
  322. Son, H.J.; Han, W.S.; Chun, J.Y.; Kang, B.K.; Kwon, S.N.; Ko, J.; Su, J.H.; Lee, C.; Sung, J.K.; Sang, O.K. Generation of Blue Light-Emitting Zinc Complexes by Band-Gap Control of the Oxazolylphenolate Ligand System: Syntheses, Characterizations, and Organic Light Emitting Device Applications of 4-Coordinated Bis(2-Oxazolylphenolate) Zinc(II) Complexes. Inorg. Chem. 2008, 47, 5666–5676. [Google Scholar] [CrossRef]
  323. Zhang, Q.; Komino, T.; Huang, S.; Matsunami, S.; Goushi, K.; Adachi, C. Triplet Exciton Confinement in Green Organic Light-Emitting Diodes Containing Luminescent Charge-Transfer Cu(I) Complexes. Adv. Funct. Mater. 2012, 22, 2327–2336. [Google Scholar] [CrossRef]
  324. Hashimoto, M.; Igawa, S.; Yashima, M.; Kawata, I.; Hoshino, M.; Osawa, M. Highly Efficient Green Organic Light-Emitting Diodes Containing Luminescent Three-Coordinate Copper(I) Complexes. J. Am. Chem. Soc. 2011, 133, 10348–10351. [Google Scholar] [CrossRef]
  325. Cheng, G.; So, G.K.M.; To, W.P.; Chen, Y.; Kwok, C.C.; Ma, C.; Guan, X.; Chang, X.; Kwok, W.M.; Che, C.M. Luminescent Zinc(II) and Copper(I) Complexes for High-Performance Solution-Processed Monochromic and White Organic Light-Emitting Devices. Chem. Sci. 2015, 6, 4623–4635. [Google Scholar] [CrossRef]
  326. Nam, S.; Kim, J.W.; Bae, H.J.; Maruyama, Y.M.; Jeong, D.; Kim, J.; Kim, J.S.; Son, W.J.; Jeong, H.; Lee, J.; et al. Improved Efficiency and Lifetime of Deep-Blue Hyperfluorescent Organic Light-Emitting Diode Using Pt(II) Complex as Phosphorescent Sensitizer. Adv. Sci. 2021, 8, 2100586. [Google Scholar] [CrossRef]
  327. 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]
  328. Bulović, V.; Deshpande, R.; Thompson, M.E.; Forrest, S.R. Tuning the Color Emission of Thin Film Molecular Organic Light Emitting Devices by the Solid State Solvation Effect. Chem. Phys. Lett. 1999, 308, 317–322. [Google Scholar] [CrossRef]
  329. Chen, W.C.; Yuan, Y.; Ni, S.F.; Tong, Q.X.; Wong, F.L.; Lee, C.S. Achieving Efficient Violet-Blue Electroluminescence with CIEy <0.06 and EQE >6% from Naphthyl-Linked Phenanthroimidazole–Carbazole Hybrid Fluorophores. Chem. Sci. 2017, 8, 3599–3608. [Google Scholar] [CrossRef]
  330. Lin, S.L.; Chan, L.H.; Lee, R.H.; Yen, M.Y.; Kuo, W.J.; Chen, C.T.; Jeng, R.J. Highly Efficient Carbazole-π-Dimesitylborane Bipolar Fluorophores for Nondoped Blue Organic Light-Emitting Diodes. Adv. Mater. 2008, 20, 3947–3952. [Google Scholar] [CrossRef]
  331. Reig, M.; Puigdollers, J.; Velasco, D. Molecular Order of Air-Stable p-Type Organic Thin-Film Transistors by Tuning the Extension of the π-Conjugated Core: The Cases of Indolo[3,2-b]Carbazole and Triindole Semiconductors. J. Mater. Chem. C Mater. 2014, 3, 506–513. [Google Scholar] [CrossRef]
  332. Chen, S.; Sun, B.; Hong, W.; Yan, Z.; Aziz, H.; Meng, Y.; Hollinger, J.; Seferos, D.S.; Li, Y. Impact of N-Substitution of a Carbazole Unit on Molecular Packing and Charge Transport of DPP–Carbazole Copolymers. J. Mater. Chem. C Mater. 2014, 2, 1683–1690. [Google Scholar] [CrossRef]
  333. Yu, M.; Wang, S.; Shao, S.; Ding, J.; Wang, L.; Jing, X.; Wang, F. Starburst 4,4′,4″-Tris(Carbazol-9-Yl)-Triphenylamine-Based Deep-Blue Fluorescent Emitters with Tunable Oligophenyl Length for Solution-Processed Undoped Organic Light-Emitting Diodes. J. Mater. Chem. C Mater. 2015, 3, 861–869. [Google Scholar] [CrossRef]
  334. Xiao, J.; Yang, S. Sequential Crystallization of Sea Urchin-like Bimetallic (Ni, Co) Carbonate Hydroxide and Its Morphology Conserved Conversion to Porous NiCo2O4 Spinel for Pseudocapacitors. RSC Adv. 2011, 1, 588–595. [Google Scholar] [CrossRef]
  335. Sahoo, S.; Dubey, D.K.; Singh, M.; Joseph, V.; Thomas, K.R.J.; Jou, J.H. Highly Efficient Deep-Blue Organic Light Emitting Diode with a Carbazole Based Fluorescent Emitter. Jpn. J. Appl. Phys. 2018, 57, 04FL08. [Google Scholar] [CrossRef]
  336. Huang, C.C.; Xue, M.M.; Wu, F.P.; Yuan, Y.; Liao, L.S.; Fung, M.K. Deep-Blue and Hybrid-White Organic Light Emitting Diodes Based on a Twisting Carbazole-Benzofuro[2,3-b]Pyrazine Fluorescent Emitter. Molecules 2019, 24, 353. [Google Scholar] [CrossRef] [PubMed]
  337. Pal, A.K.; Krotkus, S.; Fontani, M.; Mackenzie, C.F.R.; Cordes, D.B.; Slawin, A.M.Z.; Samuel, I.D.W.; Zysman-Colman, E.; Pal, A.K.; Fontani, M.; et al. High-Efficiency Deep-Blue-Emitting Organic Light-Emitting Diodes Based on Iridium(III) Carbene Complexes. Adv. Mater. 2018, 30, 1804231. [Google Scholar] [CrossRef]
  338. Wang, R.; Liu, Y.; Hu, T.; Wei, X.; Liu, J.; Li, Z.; Hu, X.; Yi, Y.; Wang, P.; Wang, Y. Solution-Processed White Organic Light-Emitting Diodes with Bi-Component Emitting Layer Based on Symmetry Blue Spiro-Sulfone Derivative. Org. Electron. 2019, 71, 24–30. [Google Scholar] [CrossRef]
  339. Weissenseel, S.; Drigo, N.A.; Kudriashova, L.G.; Schmid, M.; Morgenstern, T.; Lin, K.H.; Prlj, A.; Corminboeuf, C.; Sperlich, A.; Brütting, W.; et al. Getting the Right Twist: Influence of Donor-Acceptor Dihedral Angle on Exciton Kinetics and Singlet-Triplet Gap in Deep Blue Thermally Activated Delayed Fluorescence Emitter. J. Phys. Chem. C 2019, 123, 27778–27784. [Google Scholar] [CrossRef]
  340. Chan, C.-Y.; Cui, L.-S.; Uk Kim, J.; Nakanotani, H.; Adachi, C.; Chan, C.; Cui, L.; Kim, J.U.; Nakanotani, H.; Adachi, C. Rational Molecular Design for Deep-Blue Thermally Activated Delayed Fluorescence Emitters. Adv. Funct. Mater. 2018, 28, 1706023. [Google Scholar] [CrossRef]
  341. Luo, Y.; Li, S.; Zhao, Y.; Li, C.; Pang, Z.; Huang, Y.; Yang, M.; Zhou, L.; Zheng, X.; Pu, X.; et al. An Ultraviolet Thermally Activated Delayed Fluorescence OLED with Total External Quantum Efficiency over 9%. Adv. Mater. 2020, 32, 2001248. [Google Scholar] [CrossRef] [PubMed]
  342. Gudeika, D.; Bezvikonnyi, O.; Volyniuk, D.; Skuodis, E.; Lee, P.H.; Chen, C.H.; Ding, W.C.; Lee, J.H.; Chiu, T.L.; Grazulevicius, J.V. Oxygen Sensing and OLED Applications of Di-Tert-Butyl-Dimethylacridinyl Disubstituted Oxygafluorene Exhibiting Long-Lived Deep-Blue Delayed Fluorescence. J. Mater. Chem. C Mater. 2020, 8, 9632–9638. [Google Scholar] [CrossRef]
  343. Yang, J.; Guo, Q.; Wang, J.; Ren, Z.; Chen, J.; Peng, Q.; Ma, D.; Li, Z.; Yang, J.; Wang, J.; et al. Rational Molecular Design for Efficient Exciton Harvesting, and Deep-Blue OLED Application. Adv. Opt. Mater. 2018, 6, 1800342. [Google Scholar] [CrossRef]
  344. Fu, C.; Luo, S.; Li, Z.; Ai, X.; Pang, Z.; Li, C.; Chen, K.; Zhou, L.; Li, F.; Huang, Y.; et al. Highly Efficient Deep-Blue OLEDs Based on Hybridized Local and Charge-Transfer Emitters Bearing Pyrene as the Structural Unit. Chem. Commun. 2019, 55, 6317–6320. [Google Scholar] [CrossRef]
  345. Yu, Y.; Zhao, R.; Liu, H.; Zhang, S.; Zhou, C.; Gao, Y.; Li, W.; Yang, B. Highly Efficient Deep-Blue Light-Emitting Material Based on V-Shaped Donor-Acceptor Triphenylamine-Phenanthro[9,10-d]Imidazole Molecule. Dye. Pigment. 2020, 180, 108511. [Google Scholar] [CrossRef]
  346. Usta, H.; Alimli, D.; Ozdemir, R.; Dabak, S.; Zorlu, Y.; Alkan, F.; Tekin, E.; Can, A. Highly Efficient Deep-Blue Electroluminescence Based on a Solution-Processable A-π-D-π-A Oligo(p-Phenyleneethynylene) Small Molecule. ACS Appl. Mater. Interfaces 2019, 11, 44474–44486. [Google Scholar] [CrossRef]
  347. Zhang, H.; Xue, J.; Li, C.; Zhang, S.; Yang, B.; Liu, Y.; Wang, Y.; Zhang, H.; Xue, J.; Li, C.; et al. Novel Deep-Blue Hybridized Local and Charge-Transfer Host Emitter for High-Quality Fluorescence/Phosphor Hybrid Quasi-White Organic Light-Emitting Diode. Adv. Funct. Mater. 2021, 31, 2100704. [Google Scholar] [CrossRef]
  348. Kondakov, D.Y.; Pawlik, T.D.; Hatwar, T.K.; Spindler, J.P. Triplet Annihilation Exceeding Spin Statistical Limit in Highly Efficient Fluorescent Organic Light-Emitting Diodes. J. Appl. Phys. 2009, 106, 124510. [Google Scholar] [CrossRef]
  349. Turshatov, A.A.; Baluschev, S.B. Triplet-Triplet Annihilation Assisted Upconversion: All-Optical Tools for Probing Physical Parameter of Soft Matter. In Handbook of Coherent-Domain Optical Methods; Springer: New York, NY, USA, 2012. [Google Scholar] [CrossRef]
  350. Hu, J.Y.; Pu, Y.J.; Satoh, F.; Kawata, S.; Katagiri, H.; Sasabe, H.; Kido, J. Bisanthracene-Based Donor–Acceptor-Type Light-Emitting Dopants: Highly Efficient Deep-Blue Emission in Organic Light-Emitting Devices. Adv. Funct. Mater. 2014, 24, 2064–2071. [Google Scholar] [CrossRef]
  351. Chen, Y.H.; Lin, C.C.; Huang, M.J.; Hung, K.; Wu, Y.C.; Lin, W.C.; Chen-Cheng, R.W.; Lin, H.W.; Cheng, C.H. Superior Upconversion Fluorescence Dopants for Highly Efficient Deep-Blue Electroluminescent Devices. Chem. Sci. 2016, 7, 4044–4051. [Google Scholar] [CrossRef]
  352. Zeng, W.; Zhao, Y.; Ning, W.; Gong, S.; Zhu, Z.; Zou, Y.; Lu, Z.H.; Yang, C. Efficient Non-Doped Fluorescent OLEDs with Nearly 6% External Quantum Efficiency and Deep-Blue Emission Approaching the Blue Standard Enabled by Quaterphenyl-Based Emitters. J. Mater. Chem. C Mater. 2018, 6, 4479–4484. [Google Scholar] [CrossRef]
  353. Peng, L.; Yao, J.W.; Wang, M.; Wang, L.Y.; Huang, X.L.; Wei, X.F.; Ma, D.G.; Cao, Y.; Zhu, X.H. Efficient Soluble Deep Blue Electroluminescent Dianthracenylphenylene Emitters with CIE y (Y ≤ 0.08) Based on Triplet-Triplet Annihilation. Sci. Bull. 2019, 64, 774–781. [Google Scholar] [CrossRef] [PubMed]
  354. Wu, Z.; Song, S.; Zhu, X.; Chen, H.; Chi, J.; Ma, D.; Zhao, Z.; Tang, B.Z. Highly Efficient Deep-Blue Fluorescent OLEDs Based on Anthracene Derivatives with a Triplet–Triplet Annihilation Mechanism. Mater. Chem. Front. 2021, 5, 6978–6986. [Google Scholar] [CrossRef]
  355. 《擁抱暗黑》新書講座(憑書報名入場)|ACCUPASS. Available online: https://www.accupass.com/event/2011030819584976501710 (accessed on 15 May 2022).
Figure 1. Global revenues of LCD- and OLED-based displays from 2016 to 2024. Keyword: Global revenue OLED display 2020. Source: DSCC’s Quarterly Display Capex and Equipment Market Share Report [15].
Figure 1. Global revenues of LCD- and OLED-based displays from 2016 to 2024. Keyword: Global revenue OLED display 2020. Source: DSCC’s Quarterly Display Capex and Equipment Market Share Report [15].
Nanomaterials 13 02521 g001
Figure 2. Global LED Lighting Revenues and Forecast by Statista Research Development. Keyword: Global revenue LED lighting. Source: https://www.statista.com/statistics/753939/global-led-luminaire-market-size [16].
Figure 2. Global LED Lighting Revenues and Forecast by Statista Research Development. Keyword: Global revenue LED lighting. Source: https://www.statista.com/statistics/753939/global-led-luminaire-market-size [16].
Nanomaterials 13 02521 g002
Figure 3. Global OLED Lighting Revenues and Forecast by ElectroniCast. Keyword: Global revenue OLED lighting. Source: https://www.oled-info.com/electronicast-sees-fast-growing-oled-lighting-market-starting-2015 [17].
Figure 3. Global OLED Lighting Revenues and Forecast by ElectroniCast. Keyword: Global revenue OLED lighting. Source: https://www.oled-info.com/electronicast-sees-fast-growing-oled-lighting-market-starting-2015 [17].
Nanomaterials 13 02521 g003
Figure 4. Research papers on blue fluorescent OLEDs per year since 1990. Keywords: Fluorescent blue organic light-emitting device. Source: Web of science.
Figure 4. Research papers on blue fluorescent OLEDs per year since 1990. Keywords: Fluorescent blue organic light-emitting device. Source: Web of science.
Nanomaterials 13 02521 g004
Figure 5. Research papers on blue phosphorescent OLEDs per year; the first paper appeared in 1999 and rapidly peaked to 31 papers in 2009. Keywords: Phosphorescent blue organic light-emitting device. Source: Web of science.
Figure 5. Research papers on blue phosphorescent OLEDs per year; the first paper appeared in 1999 and rapidly peaked to 31 papers in 2009. Keywords: Phosphorescent blue organic light-emitting device. Source: Web of science.
Nanomaterials 13 02521 g005
Figure 6. Research papers on blue TADF OLEDs per year since 2015; the first paper appeared in 2012 and rapidly peaked to 71 papers in 8 years. Keywords: TADF blue organic light-emitting device. Source: Web of science.
Figure 6. Research papers on blue TADF OLEDs per year since 2015; the first paper appeared in 2012 and rapidly peaked to 71 papers in 8 years. Keywords: TADF blue organic light-emitting device. Source: Web of science.
Nanomaterials 13 02521 g006
Figure 7. An overall chart depicting the rise of blue OLED papers using fluorescent (Gen-1), phosphorescent (Gen-2), and TADF (Gen-3) emitters. The yearly paper amount for Gen-2 emitter-composed devices surpassed that of Gen-1 counterparts in 2008 (after the first paper was published in 2001), while the yearly paper amount for Gen-3-based devices surpassed that of Gen-2 in 2020 (after the first paper was published in 2012). Keywords: Blue organic light-emitting device. Source: Web of science.
Figure 7. An overall chart depicting the rise of blue OLED papers using fluorescent (Gen-1), phosphorescent (Gen-2), and TADF (Gen-3) emitters. The yearly paper amount for Gen-2 emitter-composed devices surpassed that of Gen-1 counterparts in 2008 (after the first paper was published in 2001), while the yearly paper amount for Gen-3-based devices surpassed that of Gen-2 in 2020 (after the first paper was published in 2012). Keywords: Blue organic light-emitting device. Source: Web of science.
Nanomaterials 13 02521 g007
Figure 9. EQE performance vs. CIEy of fluorescent blue OLEDs with dry and wet processes. The horizontal black dashed lines represent the theoretical 5% upper limit. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively. Some deep-blue devices were reported with a greater than 5% EQE at 1000 cd/m2 with a CIEy less than 0.1 by using the dry process.
Figure 9. EQE performance vs. CIEy of fluorescent blue OLEDs with dry and wet processes. The horizontal black dashed lines represent the theoretical 5% upper limit. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively. Some deep-blue devices were reported with a greater than 5% EQE at 1000 cd/m2 with a CIEy less than 0.1 by using the dry process.
Nanomaterials 13 02521 g009
Figure 10. Power efficacy (PE) performance vs. CIEy of fluorescent blue OLEDs with dry and wet processes. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively. None of the deep-blue devices shows a greater than 10 lm/W power efficacy at any reported luminance, regardless of whether dry or wet processes were used.
Figure 10. Power efficacy (PE) performance vs. CIEy of fluorescent blue OLEDs with dry and wet processes. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively. None of the deep-blue devices shows a greater than 10 lm/W power efficacy at any reported luminance, regardless of whether dry or wet processes were used.
Nanomaterials 13 02521 g010
Figure 11. Current efficacy (CE) vs. CIEy for dry- and wet-processed fluorescent blue OLEDs. From display application perspective, commercially viable ones are those with a CIEy lying on or above the dashed line with a CE equal to 70*CIEy. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Figure 11. Current efficacy (CE) vs. CIEy for dry- and wet-processed fluorescent blue OLEDs. From display application perspective, commercially viable ones are those with a CIEy lying on or above the dashed line with a CE equal to 70*CIEy. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Nanomaterials 13 02521 g011
Figure 12. Lifetime and current efficacy vs. CIEy for fluorescent blue OLEDs. The shaded area represents a CE ≥ 70*CIEy [30,37,41,42,43]. The different shades of blue denote the emission-color at CIEy.
Figure 12. Lifetime and current efficacy vs. CIEy for fluorescent blue OLEDs. The shaded area represents a CE ≥ 70*CIEy [30,37,41,42,43]. The different shades of blue denote the emission-color at CIEy.
Nanomaterials 13 02521 g012
Figure 13. EQE performance vs. CIEy of phosphorescent blue OLEDs with dry and wet processes. Some deep-blue devices were reported with a greater than 20% EQE at a maximum luminance dimmer than 100 nits with a CIEy less than 0.1 by using dry processes. The horizontal black dashed lines represent the theoretical upper limit. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Figure 13. EQE performance vs. CIEy of phosphorescent blue OLEDs with dry and wet processes. Some deep-blue devices were reported with a greater than 20% EQE at a maximum luminance dimmer than 100 nits with a CIEy less than 0.1 by using dry processes. The horizontal black dashed lines represent the theoretical upper limit. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Nanomaterials 13 02521 g013
Figure 14. Power efficacy performance vs. CIEy of phosphorescent blue OLEDs with dry and wet processes. No deep-blue devices show a power efficacy greater than 10 lumens per watt at 100 or 1000 nits. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Figure 14. Power efficacy performance vs. CIEy of phosphorescent blue OLEDs with dry and wet processes. No deep-blue devices show a power efficacy greater than 10 lumens per watt at 100 or 1000 nits. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Nanomaterials 13 02521 g014
Figure 15. Current efficacy (CE) performance vs. CIEy for dry- and wet-processed phosphorescent blue OLEDs. Many dry-processed devices exhibit a CE much greater than 70*CIE= in the typical blue region, while only a few exhibits deep-blue emission. The black dashed lines represent the minimum CE requirement for display. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Figure 15. Current efficacy (CE) performance vs. CIEy for dry- and wet-processed phosphorescent blue OLEDs. Many dry-processed devices exhibit a CE much greater than 70*CIE= in the typical blue region, while only a few exhibits deep-blue emission. The black dashed lines represent the minimum CE requirement for display. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Nanomaterials 13 02521 g015
Figure 16. Lifetime and current efficacy vs. CIEy for phosphorescent blue OLEDs. Some phosphorescent OLEDs show relatively high current efficacy and comparatively long lifetime in the sky-blue region. Those that fall into the grey area are plausibly commercially viable. The shaded area represents a CE ≥ 70*CIEy [30,37,41,42,43]. The different shades of blue denote the emission-color at CIEy.
Figure 16. Lifetime and current efficacy vs. CIEy for phosphorescent blue OLEDs. Some phosphorescent OLEDs show relatively high current efficacy and comparatively long lifetime in the sky-blue region. Those that fall into the grey area are plausibly commercially viable. The shaded area represents a CE ≥ 70*CIEy [30,37,41,42,43]. The different shades of blue denote the emission-color at CIEy.
Nanomaterials 13 02521 g016
Figure 17. EQE performance vs. CIEy of TADF blue OLEDs with dry and wet processes. Some deep-blue devices were successfully fabricated with an EQE greater than 20% via dry-process. The shaded area represents the upper and lower limits expected for a TADF-based device. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Figure 17. EQE performance vs. CIEy of TADF blue OLEDs with dry and wet processes. Some deep-blue devices were successfully fabricated with an EQE greater than 20% via dry-process. The shaded area represents the upper and lower limits expected for a TADF-based device. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Nanomaterials 13 02521 g017
Figure 18. Power efficacy performance vs. CIEy of TADF blue OLEDs with dry and wet processes. One device shows an around 20 lm/W power efficacy at 100 nits with deep blue emission. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Figure 18. Power efficacy performance vs. CIEy of TADF blue OLEDs with dry and wet processes. One device shows an around 20 lm/W power efficacy at 100 nits with deep blue emission. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Nanomaterials 13 02521 g018
Figure 19. Current efficacy (CE) performance vs. CIEy for dry- and wet-processed TADF blue OLEDs. One dry-processed device exhibits a near 25 cd/A maximum current efficacy with relatively deep-blue emission. Greater than 20 cd/A is also reported at 100 nits. The black dashed lines represent the minimum CE requirement for display. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Figure 19. Current efficacy (CE) performance vs. CIEy for dry- and wet-processed TADF blue OLEDs. One dry-processed device exhibits a near 25 cd/A maximum current efficacy with relatively deep-blue emission. Greater than 20 cd/A is also reported at 100 nits. The black dashed lines represent the minimum CE requirement for display. The vertical dark and light blue dashed lines represent the region for deep blue and greenish blue, respectively.
Nanomaterials 13 02521 g019
Figure 20. Lifetime and current efficacy vs. CIEy for TADF blue OLED. A couple TADF-based OLEDs show a fair lifetime with high current efficacy in the sky-blue region. The shaded area represents a CE ≥ 70*CIEy [80].
Figure 20. Lifetime and current efficacy vs. CIEy for TADF blue OLED. A couple TADF-based OLEDs show a fair lifetime with high current efficacy in the sky-blue region. The shaded area represents a CE ≥ 70*CIEy [80].
Nanomaterials 13 02521 g020
Figure 21. Numerous series of fluorescent blue emitters based on different core structure units.
Figure 21. Numerous series of fluorescent blue emitters based on different core structure units.
Nanomaterials 13 02521 g021
Figure 22. Chemical structures of some typical blue emitters, OXD-8, PPP, and SA-PC, for OLED devices.
Figure 22. Chemical structures of some typical blue emitters, OXD-8, PPP, and SA-PC, for OLED devices.
Nanomaterials 13 02521 g022
Figure 23. Molecular structures of typical DSA derivative-based blue emitters with electron donor groups.
Figure 23. Molecular structures of typical DSA derivative-based blue emitters with electron donor groups.
Nanomaterials 13 02521 g023
Figure 24. Molecular structures of non-planar (a) DSA diphenyl and (b) IDB phenyl, diphenyl, and triphenyl derivative-based blue emitters.
Figure 24. Molecular structures of non-planar (a) DSA diphenyl and (b) IDB phenyl, diphenyl, and triphenyl derivative-based blue emitters.
Nanomaterials 13 02521 g024
Figure 25. Molecular structures of pyrene-based blue emitters.
Figure 25. Molecular structures of pyrene-based blue emitters.
Nanomaterials 13 02521 g025
Figure 26. Molecular structures of anthracene-based blue emitters.
Figure 26. Molecular structures of anthracene-based blue emitters.
Nanomaterials 13 02521 g026
Figure 27. Molecular structures of fluorene-based blue emitters.
Figure 27. Molecular structures of fluorene-based blue emitters.
Nanomaterials 13 02521 g027
Figure 28. Molecular structures of biaryl-based blue emitters.
Figure 28. Molecular structures of biaryl-based blue emitters.
Nanomaterials 13 02521 g028
Figure 29. Molecular structures of spiro-based blue emitters.
Figure 29. Molecular structures of spiro-based blue emitters.
Nanomaterials 13 02521 g029
Figure 30. Molecular structures of silane-based blue emitters.
Figure 30. Molecular structures of silane-based blue emitters.
Nanomaterials 13 02521 g030
Figure 31. Molecular structures of carbazole-based blue emitters.
Figure 31. Molecular structures of carbazole-based blue emitters.
Nanomaterials 13 02521 g031
Figure 32. Molecular structures of oxadiazole-based blue emitters.
Figure 32. Molecular structures of oxadiazole-based blue emitters.
Nanomaterials 13 02521 g032
Figure 33. Molecular structures of Ir-complex-based blue emitters.
Figure 33. Molecular structures of Ir-complex-based blue emitters.
Nanomaterials 13 02521 g033
Figure 34. Molecular structure of platinum complex-based blue emitters.
Figure 34. Molecular structure of platinum complex-based blue emitters.
Nanomaterials 13 02521 g034
Figure 35. Molecular structure of Zn(Lc)2 blue emitter.
Figure 35. Molecular structure of Zn(Lc)2 blue emitter.
Nanomaterials 13 02521 g035
Figure 36. Types of aromatic TADF blue emitters based on different donor–acceptor pairs. All abbreviations are given as follows: PA (phenylamine), DPS (diphenylsulfone), ACR (acridine), Cz (carbazole), PN (phthalonitrile), PXB (phenoxaborin), PXZ (phenoxazines), TAZ (triazole), and TRZ (triazine).
Figure 36. Types of aromatic TADF blue emitters based on different donor–acceptor pairs. All abbreviations are given as follows: PA (phenylamine), DPS (diphenylsulfone), ACR (acridine), Cz (carbazole), PN (phthalonitrile), PXB (phenoxaborin), PXZ (phenoxazines), TAZ (triazole), and TRZ (triazine).
Nanomaterials 13 02521 g036
Figure 37. Molecular structures of (a) electron donor moities and (b) electron acceptor moities, frequently used in aromatic-based blue TADF emitters.
Figure 37. Molecular structures of (a) electron donor moities and (b) electron acceptor moities, frequently used in aromatic-based blue TADF emitters.
Nanomaterials 13 02521 g037
Figure 38. Molecular structures of DPA-DPS-based blue emitters.
Figure 38. Molecular structures of DPA-DPS-based blue emitters.
Nanomaterials 13 02521 g038
Figure 39. Molecular structures of ACR-DPS-based blue emitters.
Figure 39. Molecular structures of ACR-DPS-based blue emitters.
Nanomaterials 13 02521 g039
Figure 40. Molecular structures of Cz-DPS-based blue emitters.
Figure 40. Molecular structures of Cz-DPS-based blue emitters.
Nanomaterials 13 02521 g040
Figure 41. Molecular structures of Cz-TRZ-based blue emitters.
Figure 41. Molecular structures of Cz-TRZ-based blue emitters.
Nanomaterials 13 02521 g041
Figure 42. Molecular structures of Cz-PN-based blue emitters.
Figure 42. Molecular structures of Cz-PN-based blue emitters.
Nanomaterials 13 02521 g042
Figure 43. Molecular structures of Cz-PXB-based blue emitters.
Figure 43. Molecular structures of Cz-PXB-based blue emitters.
Nanomaterials 13 02521 g043
Figure 44. Molecular structures of PXZ-TAZ-based blue emitters.
Figure 44. Molecular structures of PXZ-TAZ-based blue emitters.
Nanomaterials 13 02521 g044
Figure 45. Molecular structures of ACR-TRZ-based blue emitters.
Figure 45. Molecular structures of ACR-TRZ-based blue emitters.
Nanomaterials 13 02521 g045
Figure 46. Types of non-precious metal-based TADF blue emitters.
Figure 46. Types of non-precious metal-based TADF blue emitters.
Nanomaterials 13 02521 g046
Figure 47. Molecular structures of copper (Cu)-based blue emitters.
Figure 47. Molecular structures of copper (Cu)-based blue emitters.
Nanomaterials 13 02521 g047
Figure 48. Molecular structures of zinc (Zn)-based blue emitters.
Figure 48. Molecular structures of zinc (Zn)-based blue emitters.
Nanomaterials 13 02521 g048
Figure 49. Molecular structures of Lithium (Li) and Magnesium (Mg)-based blue emitters.
Figure 49. Molecular structures of Lithium (Li) and Magnesium (Mg)-based blue emitters.
Nanomaterials 13 02521 g049
Figure 50. Jablonski diagram of a postulated TTA mechanism. The abbreviations used are S0: ground state, S1 and S2: excited singlet state, T1: triplet state, ISC: intersystem crossing, TTA: triplet–triplet annihilation, and hv: emitted photon.
Figure 50. Jablonski diagram of a postulated TTA mechanism. The abbreviations used are S0: ground state, S1 and S2: excited singlet state, T1: triplet state, ISC: intersystem crossing, TTA: triplet–triplet annihilation, and hv: emitted photon.
Nanomaterials 13 02521 g050
Table 1. Commercial OLED lighting panels.
Table 1. Commercial OLED lighting panels.
CompanyDimensions (mm)ηp
(lm/W)
Luminance
(cd/m2)
Lifetime
(h)
Reference
Philips124 × 124.5 × 3.316.7400010,000 @ L50[32]
OsramΦ75 mm35.0200010,000 @ L50[33]
Lumiotech145 × 14545.0300040,000 @ L70[36]
Konica Minolta21.4 × 25.9-500-[38]
Mitsubishi Chem.55 × 55 × 1.08-2000-[39]
Kaneka80 × 80 × 1.05-300050,000 @ -[40]
LG Chem.200 × 50 × 0.8860.0300040,000 @ L70[34]
Table 2. Efficiency and lifetime performance of dry-processed fluorescent blue OLEDs. All the lifetime data have been converted to the same 1000 nits for easy comparison.
Table 2. Efficiency and lifetime performance of dry-processed fluorescent blue OLEDs. All the lifetime data have been converted to the same 1000 nits for easy comparison.
CIEyEQEmax
(%)
L0 (cd/m2)Lifetime (h)Lifetime (h)
@L0:1000
(cd/m2)
CEmax
(cd/A)
YearReference
0.0911.81000125 (LT90)125 (LT90)9.22018[72]
0.118.61176110 (LT50)121 (LT50)4.92010[83]
0.158.25000168 (LT50)458 (LT50)9.12010[82]
0.173.01005600 (LT50)1330 (LT50)4.02005[84]
0.224.9100078 (LT80)78 (LT80)7.12020[85]
0.321.51007000 (LT50)1662 (LT50)9.72005[84]
L0: Initial luminance.
Table 3. Efficiency and lifetime performance of dry-processed phosphorescent blue OLEDs. All the lifetime data have been converted to the same 1000 nits for easy comparison.
Table 3. Efficiency and lifetime performance of dry-processed phosphorescent blue OLEDs. All the lifetime data have been converted to the same 1000 nits for easy comparison.
CIEyEQEmax
(%)
L0 (cd/m2)Lifetime (h)Lifetime (h)
@L0:1000
(cd/m2)
CEmax
(cd/A)
YearReference
0.1327.610010,000 (LT50)2375 (LT50)-2020[99]
0.16253272203 (LT50)1095 (LT50)42.62017[87]
0.235.52002.9 (LT75)1 (LT75)6.32019[100]
0.235.92002.6 (LT75)0.9 (LT75)9.12019
0.249.72004 (LT75)1.4 (LT75)-2019
0.2412.1-0.8 (LT95)--2019[101]
0.2417.1-1.3 (LT95)--2019
0.2416.1-13.8 (LT95)--2019
0.2516.92046 (LT70)4 (LT70)14.52019[102]
0.2516.91000602 (LT70)602 (LT70)-2019
0.2811.61000553 (LT50)553 (LT50)20.72020[103]
0.2817.62043 (LT70)3.7 (LT70)-2019[102]
0.2817.61000709 (LT70)709 (LT70)-2019
0.2814.7-2.1 (LT95)--2019[101]
0.291120017,500 (LT50)6410 (LT50)212006[97]
0.29182039.5 (LT70)3.4 (LT70)-2019[102]
0.29181000761 (LT70)761 (LT70)-2019
0.3225.6100020 (LT70)20 (LT70)52.12020[98]
0.3814200100,000 (LT50)36,630 (LT50)322006[97]
0.3812.62037.8 (LT70)3.2 (LT70)-2019[102]
0.3812.61000466 (LT70)466 (LT70)-2019
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

Siddiqui, I.; Kumar, S.; Tsai, Y.-F.; Gautam, P.; Shahnawaz; Kesavan, K.; Lin, J.-T.; Khai, L.; Chou, K.-H.; Choudhury, A.; et al. Status and Challenges of Blue OLEDs: A Review. Nanomaterials 2023, 13, 2521. https://doi.org/10.3390/nano13182521

AMA Style

Siddiqui I, Kumar S, Tsai Y-F, Gautam P, Shahnawaz, Kesavan K, Lin J-T, Khai L, Chou K-H, Choudhury A, et al. Status and Challenges of Blue OLEDs: A Review. Nanomaterials. 2023; 13(18):2521. https://doi.org/10.3390/nano13182521

Chicago/Turabian Style

Siddiqui, Iram, Sudhir Kumar, Yi-Fang Tsai, Prakalp Gautam, Shahnawaz, Kiran Kesavan, Jin-Ting Lin, Luke Khai, Kuo-Hsien Chou, Abhijeet Choudhury, and et al. 2023. "Status and Challenges of Blue OLEDs: A Review" Nanomaterials 13, no. 18: 2521. https://doi.org/10.3390/nano13182521

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