Next Article in Journal
Effect of Synchrotron X-ray Irradiation Time on the Particle Size and DFAFC Performance of Pd/CNT Catalysts
Previous Article in Journal
Selenium Nanoparticles in Protecting the Brain from Stroke: Possible Signaling and Metabolic Mechanisms
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Optimization of Hole Injection Layer in Organic Light-Emitting Diodes

by
Xiaolin Xing
1,2,
Ziye Wu
1,2,
Yingying Sun
1,2,
Yunlong Liu
1,2,*,
Xiaochen Dong
1,2,3,
Shuhong Li
1,2 and
Wenjun Wang
1,2,*
1
School of Physical Science and Information Technology, Liaocheng University, Liaocheng 252059, China
2
Shandong Provincial Key Laboratory of Optical Communication Science and Technology, Liaocheng 252059, China
3
Key Laboratory of Flexible Electronics (KLOFE), Institute of Advanced Materials (IAM), School of Physical and Mathematical Sciences, Nanjing Tech University (NanjingTech), Nanjing 211800, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(2), 161; https://doi.org/10.3390/nano14020161
Submission received: 6 December 2023 / Revised: 2 January 2024 / Accepted: 10 January 2024 / Published: 11 January 2024

Abstract

:
Organic light-emitting diodes (OLEDs) are widely recognized as the forefront technology for displays and lighting technology. Now, the global OLED market is nearly mature, driven by the rising demand for superior displays in smartphones. In recent years, numerous strategies have been introduced and demonstrated to optimize the hole injection layer to further enhance the efficiency of OLEDs. In this paper, different methods of optimizing the hole injection layer were elucidated, including using a suitable hole injection material to minimize the hole injection barrier and match the energy level with the emission layer, exploring new preparation methods to optimize the structure of hole injection layer, and so on. Meanwhile, this article can help people to understand the current research progress and the challenges still faced in relation to the hole injection layer in OLEDs, providing future research directions to enhance the properties of OLEDs.

1. Introduction

Since the significant discovery that Eastman Kodak launched commercial organic light-emitting diodes (OLEDs) with a two-layer structure in 1987, the exploration of OLEDs has begun and increased. In recent years, OLEDs have become the most popular future display and lighting technology due to their superior characteristics, such as higher visual experiences, lower power consumption, design flexibility, and so on [1]. Driven by the rising demand for superior displays in smartphones, the global OLED market is nearly mature now. However, OLEDs have also faced various challenges, such as shorter lifetime, lower efficiency with severe efficiency roll-off, etc. The development of outdoor lighting with OLEDs was impeded by the device degradation from exposure to ultraviolet light and excessive moisture.
In order to enhance the performance of OLEDs, various techniques have been developed. One of the most effective strategies is the incorporation of hole injection layers (HILs) between the hole transport layer (HTL) and electrodes. The HIL promotes the charge injection from the electrode to the emissive layer (EML), while preventing excitons from being quenched by the electrodes at the same time. The typical device configuration that has been widely reported includes indium tin oxide (ITO) (anode)/HIL/HTL/EML/electron transport layer (ETL)/electron injection layer (EIL)/aluminum (Al) (cathode) (Figure 1).
This paper focuses on the latest progress in improving OLED performance by modifying HIL, mainly in terms of materials, structure, preparation methods, and the influence of post-treatment. We mainly reviewed the critical material strategies that were implemented to achieve high-performance OLEDs using organic or inorganic hole injection materials. Subsequently, the recently reported OLEDs based on different structures of preparation and the effect of post-treatment methods on OLEDs were summarized. Finally, it concluded with a discussion of perspectives and remaining issues for the future. The highlights are summarized in Figure 2.

2. Hole Injection Layer Materials

With the development of materials design and synthesis, new-type HILs with different structures and functional groups appeared. An effective hole injection material must be of relatively high glass transition temperature, good hole mobility for high conductivity, and high triplet energy to block triplet exciton. In the quest to avoid charge traps or carrier recombination, energy level alignments and the presence of low interface states are valuable [2,3]. Different materials provide different work functions, hole mobility, and hole injection capabilities. Then, the selection of HIL materials significantly impacts the performance of OLEDs.
To date, a wide range of HIL materials have been developed and extensively employed to enhance the OLED performance. Here, we explored the HIL materials from various categories, including organic materials and inorganic materials.

2.1. Organic Materials

Conductive polymers and organic small molecules are often used as HILs in OLEDs. The requirements for HILs include excellent optical and electrical characteristics, such as high conductivity, transparency, as well as low surface roughness [4,5,6]. In order to optimize the HILs to enhance the OLED performance, the diffusion properties, morphological stability, and environment friendliness of HILs are often considered. Figure 3 presents the structures of organic materials used as HILs.
To reduce the manufacturing cost, solution-processed organic HILs have been implemented to enhance device performance [7,8,9]. Conducting polymers represent a highly promising class of organic semiconductors due to their unique optical and electrical properties. For example, PEDOT:PSS, considered as one of the mostly widespread substances for the preparation of hole injection/transport layers (HITLs) in solution-processed OLEDs, is characterized by suitable energy levels, favorable hole mobility, and an easy ability to form smooth films [10,11]. Amare Benor et al. also proved that carrier injection can be enhanced by adjusting the work function (WF) of PEDOT:PSS, thereby improving the device power efficiency (PE) [10].
However, PEDOT:PSS tends to corrode indium tin oxide (ITO) due to hygroscopicity and strong acidity, which limits the durability and reliability of the devices. Moreover, it has the potential to induce instability and profound deterioration in thin-film optoelectronic devices [2]. Based on the above, Sha Wu et al. introduced a PH-neutral PEDOT:PSS to form double-layer HILs with acidic PEDOT:PSS, which improved the OLED performance [12]. Mingguang Li et al. reduced the acidity of PEDOT:PSS HILs and constructed continuous hole transport channels by using a thermal imprinting and vapor annealing method. Finally, OLEDs with current efficiency (CE) of 36.62 cd A−1, PE of 27.60 lm W−1, and external quantum efficiency (EQE) of 18.80% were obtained [13].
More materials are also proposed to substitute PEDOT:PSS. For instance, polythienothiophene:poly(perfluoroethylene-perfluoroethersulfonic acid) (PTT:PFFSA), which was used as an HIL in OLEDs, can suppress the accumulation of charges at the anode/HIL interface [14]. Compared to OLEDs featuring PEDOT:PSS as the HIL, the device utilizing PTT:PFFSA as an HIL displays enhanced efficiency, slower luminance decay, and slower increases in operating voltage.
The diffusion properties of the HIL materials will change the properties of HTLs and should be considered. For instance, the tetrafluorotetracyanoquinodimethane (F4-TCNQ) used as an HIL with different thicknesses diffuses into the HTL with different lengths and depths and leads to the F4-TCNQ ionization and the p-type doping of F4-TCNQ with the hole transporting material. The application of an electric field has been found to increase both the diffusion depth and length of F4-TCNQ to the HTL. It was found that the diffusion of a 1 nm thickness F4-TCNQ layer induced exciton dissociation within the EML and led to a decrease in the stability of OLEDs [15,16].
Moreover, F4-TCNQ has electron affinity and morphological stability [17]. When it is doped into other HILs, it can increase the HOMO level. Xiao-Lan Huang et al. reported the use of a simple small-molecule organic HIL known as N,N,N′,N′-tetra(4-methoxyphenyl)[2,2′-binaphthalene]-6,6′-diamine for OLEDs. When the material doped with F4-TCNQ, the HOMO level increased from 4.68 eV to 5.27 eV, enhancing the hole conductivity up to 2.64 × 10−4 S m−1 compared to the neat film of 1.68 × 10−4 S m−1 [18].
Biological materials are non-toxic and environmentally friendly, which was explored as HILs in OLEDs. The nucleic acid bases (nucleobases) present in the DNA biopolymer, which possess a diverse range of energy levels, can be easily thermally evaporated and can be used as HILs without additional modification. The intrinsic physical and chemical affinity between the thin-layer adenine and the gold electrode enhances hole injection, leading to a significant increase in CE and luminance [19]. In 2012, Chih-Chia Cheng et al. found that the physically cross-linked low-regioregular nucleobase functionalized poly(3-thiophene)s (P3HTs) incorporating overhanging uracil groups (LR-PU) could significantly smooth the surface and enhance solvent resistance, thermal stability, hole injection, and electron-blocking ability of films. The prepared OLEDs had a maximum luminance of 6310 cd m−2 and a luminescence efficiency of 1.24 cd A−1, outperforming the OLEDs with LR-P3HT and HR-P3HT as a mono HIL [20]. In 2015, PTC-U, a DNA-mimetic π-conjugated poly(triphenylamine-carbazole) with pendent uracil groups, was employed as an HIL. After that, Chih-Chia Cheng et al. prepared new poly(triphenylamine-carbazole)-based adenine (PTC-A) as an HIL, which possessed high self-complementarity in both solution and solid states due to the formation of adenine–adenine (A–A) pairs resulting from induced hierarchical self-assembly. And the device performance based on PTC-A had the highest efficiency compared with those utilizing PTC-U and PEDOT:PSS [21,22]. The structures of OLEDs with biomaterial HILs are shown in Figure 4, and the optical properties of the devices with biomaterial HILs are shown in Table 1.

2.2. Inorganic Materials

Even though organic materials reduced the preparation cost, the improvement in OLED efficiency is still limited. Transition metal oxides (TMOs) and two-dimensional (2D) materials are often used inorganic HIL materials. Transition metal oxides (TMOs), like molybde-num trioxide (MoO3), tungsten trioxide (WO3), and vanadium pentoxide (V2O5), and metal sulfides like tungsten sisulfides (WS2), have been increasingly employed as HILs due to their high technological compatibility, appropriate electronic characterization, and low light absorption in the visible spectrum [23,24].
TMOs have been identified as promising candidates for replacing PEDOT:PSS as HILs for OLEDs [25,26,27,28]. In 1996, Tokito et al. first used vanadium, molybdenum, and ruthenium oxide films to modify the anode, which increased the hole injection and significantly improved the OLED performance [29]. Matsushima et al. implemented 0.75 nm thick MoO3 as HILs for OLEDs, and the ohmic hole injection effect formed by the ultrathin MoO3 HIL significantly reduced the driving voltage [30]. You et al. reported efficient green fluorescent OLEDs with lower turn-on voltage and excellent stability of over 50,000 h by using the MoO3 layer as HILs [31]. Rui Liu et al. prepared a wavelength-tunable multicolor microcavity (μC) OLED device only by controlling the thickness of the MoO3 hole injection layer, and maximal brightness of 140,000 cd m−2 was obtained [32].
The thermally evaporated MoO3 has high conductivity and is beneficial to energy alignment because of the oxygen vacancies, which can also be fabricated through a solution method [32,33]. Xiaowen Zhang et al. used solution-processable MoOx HILs to increase the luminous efficiency of OLEDs. It was found that the tailored surface of WF and proper hole injection arose by the oxygen vacancies, optimizing the carrier balance in OLED and leading to enhanced performance [33].
At the same time, the transmittance and conductivity of TMO also affect the performance of devices. For example, although V2O5 has the highest WF compared with WO3 and MoO3, the device performance based on V2O5 HIL is inferior to the latter due to its high resistance and low transmittance [34]. The structures of OLEDs with transition metal oxide HILs are shown in Figure 5, and the optoelectric properties of OLEDs based on transition metal oxide hole injection materials are shown in Table 2.
Two-dimensional (2D) materials such as transition metal dichalcogenides (TMDs) and graphene have garnered significant attention. TMDs, such as MoS2, WS2, TaS2, and TiS2, have been used in field-effect transistors, solar cells, and catalysis, due to their excellent carrier transport properties [35]. WSx and MoSx nanodots were synthesized and used as hole injection layers by Quyet Van Le et al. [36]. The synthesized MoS2 nanosheets were composed of two layers with suitable surface WF and superior film morphology [37]. In the research of Quyet Van Le et al., the introduction of 250 °C annealed (NH4)2WS4 between ITO and NPB led to a reduction in Φh at the ITO/NPB interface from 1.3 V to 0.7 eV, surpassing the PEDOT:PSS/NPB interface by 0.1 eV. Detailed representations of energy level alignment diagrams are depicted in Figure 6a–c, and Figure 7 shows the characteristics of OLEDs compared with the results. The lower Φh does not cause the higher OLED performance. From the J–V characteristics shown in Figure 7d, at the same applied voltage, the hole injection current density based on 250 °C annealed (NH4)2WS4 is found to be higher compared to that based on PEDOT:PSS. The higher hole injection may lead to a carrier balance as well as balanced electron hole pairs and well-matched energy levels, resulting in better performance of OLEDs based on 250 °C annealed (NH4)2WS4 layers [38].
Phosphomolybdic acid (PMA), a heteropoly acid containing MoO3 units, was also used as an HIL for OLEDs due to its physical and optical properties. Compared with PEDOT:PSS-based devices, lower driving voltage was demonstrated. It is worth mentioning that the annealing temperature and atmosphere will have a significant impact on the WF of PMA films, thus affecting the hole injection ability, resulting in different OLED performances [39,40].
Graphene oxide (GO), because of its high WF and surface coverage, was synthesized by a tip ultrasonic-assisted liquid-phase exfoliation method with a size distribution of 200–400 nm and used as an HIL for OLEDs [35]. The small-sized GO can introduce more oxygen-containing functional groups due to increasing total edge, leading to a higher WF. When used as an HIL in OLEDs, the carrier injection can be balanced by adjusting the thickness of the GO HIL, thus improving the device performance [41]. Reduced graphene oxide (RGO) HIL can be obtained through chemical and thermal treatment to GO, which exhibits high mechanical strength, high thermal conductivity, electrical performance, and electrochemical activity [42,43,44,45].
Perovskites, often used as active materials for LEDs and solar cells, have been used as HILs due to their electronic properties. As we know, if the contact/adhesion between interfaces is poor, the effective migration of holes from the ITO anode to the adjacent organic layer will be hindered [46]. Thus, the better contacts of perovskite at the ITO/HIL and HIL/HTL interfaces are beneficial for hole injection and transmission. A series of different-thickness CH3NH3PbI3 films was prepared on ITO substrates by spin-coating. Maximum brightness of 19,110 cd m−2 and EL efficiency of 3.242 cd A−1 were obtained for the device with a 22 nm thick HIL. At the same time, the operation stability of OLEDs has been significantly improved. All these improvements can be ascribed to a better balance of charge recombination in the OLED light-emitting zone and probably due to the non-aqueous treatment of the HILs, which increases the chemical stability of the contact surfaces of ITO and CH3NH3PbI3 [2].
Inorganic materials have shown great prospects in optoelectronic devices due to favorable thermal stability, high charge carrier mobility, and suitable energy levels. It is unfortunate that a relatively high temperature is needed to anneal the precursor film to fabricate the inorganic HILs, which is detrimental in flexible optoelectronics devices [2]. The structures of OLEDs with inorganic HILs are shown in Figure 8, and the optical properties of OLEDs based on inorganic hole injection materials are shown in Table 3.

3. Doped Hole Injection Layer

To overcome the obstacles like limited charge carrier concentration or insufficient charge balance in the device due to large hole injection barriers, doping HILs is a proven technique [51]. Doped materials can be roughly classified into three categories: polymer materials, transition metal oxide materials, metal and metal compound materials.

3.1. Polymer Materials

In order to enhance the hole injection, electron blocking, and reduce the massive exciton quenching that arises by using PEDOT:PSS as HILs, different materials are doped with PEDOT:PSS [52,53,54]. By adjusting the proportion of PEDOT and poly(3-hexylthiophene) (P3ht) monomer polymerization, Keon-Soo Jang et al. obtained a composite film with controllable band gap and used it as an HIL in OLEDs [55]. Changyeon Lee et al. prepared HILs with corrugated structures by doping PEDOT:PSS with a polystyrene nanosphere (PS NS). The enhancement of light extraction caused by the corrugated structure may be the main factor improving the efficiency of the devices [56]. Tae-Hee Han et al. doped a tetrafluoroethylene-perfluoro-3,6-dioxa-4-methyl-7octene-sulfonic acid copolymer (PFI) to PEDOT:PSS and used it as an HIL (called GraHIL). The self-organized GraHIL has excellent hole injection/electron-blocking capability because of the gradient WF constructed by the self-organization of PFI, which hindered the exciton quenching at the HIL/EML interface. The self-assembled monolayers can also increase the surface contact angle, decrease root-mean-square (RMS) surface roughness, and improve WF [57]. Similarly, Cheng-Chieh Lo et al. enhanced the performance of OLEDs with acetone and deionized water-modified PEDOT:PSS as HILs. The former exhibited a 34% reduction in surface hydrophobicity and a 200% increase in conductivity. Moreover, the modified HIL showed high transmittances and low refractive indices [58].
Except for the PEDOT:PSS polymer, there are some other polymers, which can be used as HILs. WenFen Su et al. balanced the charge injection using a poly (fluorene-co-triphenylamine) (PFO-TPA) crosslinked HIL, and the maximum brightness efficiency of the device was six-times higher than without the PFO-TPA layer [59]. Afsoon Fallahi et al. used Go-doped Poly[(2,5-bis(2-(N,N-diethylammonium bromide)ethoxy)-1,4phenylene)-alt-1,4-phenylene] (PPPNEt2.HBr) as an HIL to reduce the hole injection barrier between ITO and HTL, thereby reducing the driving voltage of OLEDs [60].
As indicated in previous studies, surface roughness plays a significant role in device efficiency [61,62]. In 2011, Anna De Girolamo Del Mauro et al. used pure polyaniline (PANI)-poly(styrene sulfonate) (PSS) and dimethyl sulfoxide (DMSO)-doped PANI-PSS as HILs for OLEDs, respectively. The former exhibited a smoother film and, thus, affected the performance of corresponding OLEDs [63]. Mi-Ri Choi et al. mentioned that the doping of a perfluorinated ionomer (PFI) in conductive polymer compositions could enhance the hydrophobicity and WF. Therefore, a solution-processable HIL with a high WF composed of PANI:PSS and PFI (PANI: PSS/PFI) was introduced, which improved the hole injection efficiency and current efficiency of OLEDs. The results showed that high concentrations of PFI play a decisive role in the increase in WF. In addition, when perfluorinated ionic polymers were preferentially located at a film surface, they significantly increased the surface ionization potential [64].

3.2. Transition Metal Oxide Materials

As mentioned above, the advantages of transition metal oxides can not only be used as hole injection layer alone but also be co-doped with other materials. However, due to the poor solubility of MPcs and transition metal oxides, the fabrication of a thin film typically requires high-vacuum machining and complex operations, resulting in higher production costs [2]. Metal phthalocyanines (MPcs), one of the well-recognized HIL materials in OLEDs, have excellent thermal/chemical stabilities and electronic properties [65]. In 2012, Linsen Li et al. reported a green OLED with MoO3-doped phthalocyanine (CuPc) as an HIL, because of the complex of electrical charge transfer between MoO3 and CuPc, and the devices showed a turn-on voltage of 4.4 V and PE of 4.3 lm W−1 at a brightness of 100 cd m−2. The issue of CuPc raising the device turn-on voltage is solved by the doping of MoO3 [66]. Different concentrations of transition metal oxides doped with PEDOT:PSS as HILs will affect the OLED performance due to the changed WF [67].
Tungsten oxide nanocrystals (WO3NCs) were introduced into PANI:PSS by a solution-processed method and used as a hybrid HIL in OLEDs. The existence of PANI:PSS reduced the surface defects and enhanced the film formation of WO3NCs efficiently without impacting the conductivity of the OLED. WO3NCs improved the WF of the HIL, so that the energy level matched well between the HIL and the anode [68].
In 2018, Weiwei Deng et al. reported an OLED with solution-processed WO3-P2O5, which showed a lower driving voltage for 2.6 V and maximum brightness for 13,553 cd m−2, respectively [62]. In 2019, Mangey Ram Nagar et al. reported that a solution-processed OLED with MoO3:WO3 showed a PE of 22.9 lm W−1 at 1000 cd m−2, which was ascribed to the robust hole transporting ability, balanced charge carrier, and non-acidic nature of transition metal oxide HILs [69].

3.3. Metal and Metal Compound Materials

Active metal materials with low ionization energy are a natural n-type dopant. They are often used as an electron transport material to improve the electron transport capability of devices. However, recent studies have found that the hole injection layer can be doped with metal materials to improve the hole injection capacity.
For example, NiOx has attracted more and more interest in OLEDs due to its good stability, deep valence band, and large band gap. However, the conductivity of a single NiOx film is relatively low, and the energy level mismatch between the ITO and HTL interface hinders the hole injection in OLEDs [70,71]. The doping process is an efficient method to improve the conductivity and carrier mobility of NiOx [72,73]. Cu, Ag, and Li have been used as dopants for NiOx films. This has shown that Cu and Ag can improve the conductivity of NiOx films, and the synergistic effect can enhance the hole injection and improve the efficiency of Li:NiOx-based OLEDs [9,74,75].
Meiling Shan et al. proved that in CuI-doped m-MTDATA layers, the free carriers generated by charge transfer between CuI and m-MTDATA molecules lead to increased conductivity and form ohmic contact at the ITO/HIL interface [76].
Metal nanoparticles (MNPs) developed localized surface-plasmon resonance (LSPR) effects at the metal–dielectric interface and formed a bi-electrical layer on the anode, which can optimize the Fermi level and improve the hole injection effectively [54,77,78]. The introduction of MNPs, nanorods, nanowire arrays, and nanoclusters as HILs improved the performance of OLEDs [79,80,81,82,83]. In 2018, Jayaraman Jayabharathi et al. used the size-controlled Au-Ag NPs-PEDOT:PSS as an HIL, which has high IQE and light out-coupling efficiency, synergistically enhancing OLED performances. Compared with PEDOT:PSS-based devices, the PE, CE, and EQE of Au-Ag NPs co-doped green OLEDs were increased by 67.3%, 96.1%, and 96.0%, respectively [84]. Similarly, the improved photoelectric performance of GO/Au NP-based OLEDs can be attributed to the LSPR excited by the Au NPs, which cause an enhanced emission intensity and effective energy transfer [85]. GO not only helped to balance the hole and electron injection but optimized the energy level structure of the device, further enhancing the OLED performance [86].
Hong Lian et al. prepared 0.5 wt% Fe3O4@Au, 0.5 wt% Fe3O4@G, or 0.25 wt% Fe3O4@SiO2 NPs in mixed solutions. By introducing Au into Fe3O4 NPs, some emerging synergistic effects as well as magnetic effects can be obtained. The combination of light scattering, LSPR, and magnetism effects improved the electroluminescence efficiency of OLEDs. Compared with PEDOT:PSS-based devices, OLEDs with optimal Fe3O4@Au/PEDOT:PSS and Fe3O4@G/PEDOT:PSS composite HILs achieved enhancements of 35% in CE and with Fe3O4@SiO2, NP-PEDOT:PSS HIL obtained a 26.8% increase in CE [87]. Dandan Zhang et al. also mentioned that the Au coating can efficiently stabilize high-moment magnetic NPs. By utilizing the Fe3O4@Au NPs (FOA NPs), the CE of Alq3-based fluorescent OLEDs obtained a 90% increase compared with that without NPs. Additionally, when an external magnetic field (EMF) of 0.5 T was applied, CE can be further increased by approximately 20% due to increased singlet exciton induced by the spin-injected holes via FOA NPs. However, the CE of Ir(ppy)2(acac)-based OLEDs decreased mildly. It indicated that the magnetic field effects depend on the system and mechanism of emitting material [88].

4. Composite Hole Injection Layer Structure

Apart from the type of material, different device structures also affect the OLED performance. So-called composite HIL refers to two or more layers of film acting as HILs together. The performance of OLED devices is improved through the synergistic action between the layers. And the energy level barrier between different layers is small, which is convenient for carrier injection and, thus, improves device efficiency. The advantages of composite hole injection layers are obvious. The composite HIL favors a continuous power function gradient as well as the formation of ohmic contacts to decrease the injection barrier, thus increasing the hole injection capability and device performance.
These structures can intuitively display the above theories, such as the devices with hole injection layers of Ag2O/MoOx, MoO3/NPB/MoO3/CBP, as even the concentration gradient provided a stepped energy level, which greatly facilitated hole injection and, hence, enhanced the injection current [89,90,91]. It has been reported that the presence of charge transfer between MoO3 and NPB is also conductive to hole injection [92,93].
Yoon Ho Huh et al. proved that GO/MoO3/PEDOT:PSS three-layer stacked HIL has better hole injection characteristics than pure PEDOT:PSS [94]. Ling Chen et al. prepared WO3/PEDOT:PSS double-layer HILs. The double-layer structure provides a stepped energy level arrangement and smoother surface. The prepared device has an EQE of 12.47%, and a lifetime of 12,551 h was obtained without decreasing efficiency [95]. Lin Lu et al. demonstrated that OLEDs based on F4-TCNQ doping 4,4′4″-Tris(N-(2-naphthyl)-N-phenyl-amino)-triphenylamine(2T-NATA)/2T-NATA composite HILs had lower operating voltages and longer device lifetimes than single-layer devices [96].
Xiaowen Zhang et al. used a PEDOT:PSS/MoOx bilayer as an HIL for an ultraviolet OLED. The device based on a bilayer HIL exhibited a maximum EQE of 4.6% at 7.28 mA cm−2, an improvement of 59% and 31%, compared to single HILs of MoOx or PEDOT:PSS devices, respectively. This can be ascribed to the variation in HIL WF. The schematic inserted in Figure 9 displays an energy level diagram drawn from the measurements of the UPS. The stepped energy level arrangement makes hole injection easier and, correspondingly, promotes the EQE of devices [97].
Except for using a double-hole injection layer, the synergy between the hole injection layer and other layers can also be used. For example, although the roles of HILs and hole blocking layers were opposite, maximum luminance efficiency and PE could be obtained when their thicknesses were in a specific combination [98].

5. Preparation Methods

At present, the most commonly used methods for preparing hole injection layers are spin-coating and vacuum deposition. Spin-coating has been widely implemented in the fabrication of polymer-based OLEDs due to its cost-effective and easy-to-manufacture nature, but it suffers from low material utilization and difficulties in large-scale industrial production [23,99]. The HIL prepared by vacuum evaporation has the advantages of high purity and density, unique structure, and properties in the film. However, vacuum processing is costly and requires complicated manufacturing equipment [23]. As alternatives, some methods have been developed, like spray-coating, roll-to-roll coating, screen-printing, blade-coating, slot-die coating, Langmuir–Blodgett (LB) deposition technique, nozzle printing technology, and brush coating [99]. Figure 10 shows a schematic diagram of different preparation methods.
LB technology can control the film arrangement and thickness evenly, and the preparation process does not require high temperature and pressure. The arrangement features of carbon sheets lead to significant changes in electrical ability, which greatly influences the performance of OLED devices. The LB deposition technique is superior in inhibiting the wrinkling and folding of GO sheets substantially. The higher surface pressure makes the reduced graphene oxide (RGO) sheets’ arrangement more compact and continuous, forming compact and uniform conductive paths for efficient carrier transport. Yajie Yang et al. used LB technology to incorporate orderly and controllable-thickness RGO flakes between the ITO and HIL, which served to planarize the anode surface. The maximum brightness of the device was about 6232 cd m−2, which was more than the spin-coating RGO devices [104]. However, LB technology is limited in terms of expensive equipment and scarce available materials.
Aqueous polymer-based HILs are usually made by large-area slot-die coating due to the low material waste. Large-area OLEDs for illumination were fabricated by Kwang-Jun Choi using slot-die coating. The performance of OLEDs with optimized slot-die-coated multiple layers is comparable to the performance of OLEDs with spin-coated HIL or vacuum-evaporated HTL [105]. Moreover, Amruth C et al. obtained uniform PEDOT:PSS films by modifying the parameters of slot-die processing, such as the speed of coating, ink flow rate, and substrate temperature [101]. This method can also be used to prepare flexible devices [106]. However, it is tedious to explore suitable parameters in the slot-die technique for different materials, which is more suitable for large-scale production.
For fear of the influence of the high surface tension of water, a PEDOT:PSS solution mixed with various concentrations of ethanol (EtOH) was employed as HILs by using the nozzle printing technique. The CE of the device containing 90% ethanol solution reached 27 cd A−1 [102]. In 2021, Kwon-Yong Shin et al. successfully deposited PEDOT:PSS solution diluted 90 vol% with EtOH as HIL by employing a micro-multi-nozzle spray-coating method [107]. Despite the fact that the HIL preparation based on printing technology has numerous benefits, like simple preparation, low cost, and suitability for large-area processing, the uniformity of the prepared films is still a critical issue.
The brush-coating technique has a material utilization rate close to 100% and has no material constraints. In 2021, red, green, blue (RGB), and white phosphorescent OLEDs were manufactured by sequential brush coating on smooth HILs. The CE of the brush-coated white OLED device was comparable to that of the spin-coated device, which was about 29.3 cd A−1. Furthermore, the peak current efficiencies of the brush-coated red, green, and blue phosphorescent OLEDs were 16.7, 30.4, and 20.9 cd A−1, respectively [99].
It is requisite to select appropriate film preparation methods according to the materials of the functional layer and substrate.

6. Different Treatment Methods

The prepared hole injection layer will be affected by atmosphere, temperature, humidity, and even solar radiation, thus affecting the device performance [84]. Therefore, people try to use different post-treatment methods for hole injection layer to improve the device performance.
Ultraviolet/ozone (UVO) treatment is proved to be an efficient method for HILs to increase the WF, smooth the defects of films, and improve the hole injection [57,108,109]. WSx and MoSx nanodots were synthesized and used as hole injection layers by Quyet Van Le et al. After the UVO treatment, the WF of the nanodots was raised from 4.3–4.4 to 5.0–5.1 eV. The maximum luminance efficiency of the UVO-MoSx based OLED was 14.7 cd A−1, while that of PEDOT:PSS-based devices was 13.1 cd A−1. Moreover, the use of UVO-treated MoSx or WSx nanodots as HILs could prolong the stability of OLEDs in air [36].
HILs prepared on a substrate with plasma treatment such as CF4 and Ar can also improve device performance [110,111]. In a report by Joo Hyung Kim et al., the optimized Ar plasma treatment can improve the physical characteristics of WO3 HIL and, thus, enhance the device performance [111].
The influence of temperature and atmosphere on the preparation of the hole injection layer cannot be ignored. Sul A. Choi et al. discovered that the interfacial morphology as well as the phase state were altered by the substrate temperature, which had an impact on the copper phthalocyanine (CuPc) OLED performance. As the substrate temperature elevated from room temperature to 300 °C, the crystal structure of CuPc films changed from an orthorhombic crystal α-phase to a monoclinic crystal β -phase. The α-phase favors hole transport, and the β -phase increases the surface roughness. Consequently, 10 nm of β -phase CuPc was added on top of the α-phase CuPc layer, which was used as an HIL. This enhanced both the CuPc carrier migration and the device brightness in comparison to the pure β -phase CuPc-based OLEDs [112]. Also, the structure and conductivity of GO were altered by thermal treatment, thus forming RGO with better conductivity [104].
Furthermore, optimizing the carrier injection could employ the polar solvent vapor annealing (PSVA) method, resulting in a significant reduction in charge accumulation at the OLED interface. After the double-layer PSVA treatment, the efficiency roll-off of OLEDs dropped from 33.3% to 26.6% [113].
Light treatment of materials before preparing hole injection layers will also change the performance of devices. Janardan Dagar et al. demonstrated that the device properties with 2D MoO3 nano-flakes as the HIL are closely related to the irradiation time. This is attributed to the transformation of 2D MoO3 nano-flakes from a semiconductor phase to a metal phase after increasing the irradiation time [114]. Hsi-Kang Shih et al. used PTC-U irradiated with UV for 1 h as the hole injection layer, and hole injection, hole transport, and solvent resistance properties of PTC-U photo-crosslinked for 1 h showed a great improvement, as compared with the PTC-U without UV irradiation [22].

7. Conclusions and Outlook

Though OLED televisions and panels are marketized now, there is still room for growth. The relatively lower efficiency of blue OLEDs and the severe efficiency roll-off need to be perfected. The introduction of optimized HILs between electrodes and HTL is a promising solution to facilitate charge injections into EML and prevent the excitons from being quenched by electrodes at the same time.
In this review, we described the materials from 2010 to 2022 that can be used as hole injection layers and classified them into organic materials and inorganic materials. It is clear that organic materials reduce the cost of preparation, but improvements in the efficiency of OLEDs are not as good as for inorganic materials. Moreover, a method of improving device performance by doping two or more materials is described. The roughness and WF of the device surface can be changed by doping. Certain materials can form self-organizing surfaces to reduce exciton quenching. The LSPR effects of mental nanoparticles and the applied magnetic field can also improve the device performance. In addition, the methods used to prepare the hole injection layer are described, such as spin-coating, brush coating, LB deposition, slot-die coating, nozzle printing, evaporation, etc. The comparison of various methods is also pointed out in this paper. Finally, we described the influence of the external environment such as temperature and sunlight on the hole injection layer film.
In short, the performance of OLEDs was enhanced by introducing appropriate HILs. The optimized HIL can provide better energy level matching, reducing the negative influence between adjacent layers, balancing the injection of carriers, and leading to the enhanced optoelectronic properties of OLEDs. However, there are still many pending problems. Different types of composite hole injection layers are waiting to be excavated, as well as whether there are other mechanisms to choose, except forming ladder potential barriers. In addition to the existing hole injection layer preparation methods, there is a demand for the development of more energy-saving, cost-saving, and simpler methods. The implementation of low-cost, large-scale manufacturing will certainly be accelerated because of further awareness and developments.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 61775089), the Shandong Province Natural Science Foundation of China (No. ZR2022MF240, ZR2020KB018), and the Project of Liaocheng University (318011904).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shahnawaz, S.; Sudheendran Swayamprabha, S.; Nagar, M.R.; Yadav, R.A.K.; Gull, S.; Dubey, D.K.; Jou, J.-H. Hole-transporting materials for organic light-emitting diodes: An overview. J. Mater. Chem. C 2019, 7, 7144–7158. [Google Scholar] [CrossRef]
  2. Feng, Y.; Xu, J.; Shan, H.; Dong, L.; Sun, X.; Hu, Q.; Wang, Y.; Roy, V.A.L.; Xu, Z.-X. Organometal halide perovskite as hole injection enhancer in organic light-emitting diode. Org. Electron. 2017, 51, 257–263. [Google Scholar] [CrossRef]
  3. Jou, J.-H.; Sahoo, S.; Dubey, D.K.; Yadav, R.A.K.; Swayamprabha, S.S.; Chavhan, S.D. Molecule-based monochromatic and polychromatic OLEDs with wet-process feasibility. J. Mater. Chem. C 2018, 6, 11492–11518. [Google Scholar] [CrossRef]
  4. Zhao, B.; Miao, Y.; Wang, Z.; Wang, K.; Wang, H.; Hao, Y.; Xu, B.; Li, W. High efficiency and low roll-off green OLEDs with simple structure by utilizing thermally activated delayed fluorescence material as the universal host. Nanophotonics 2017, 6, 1133–1140. [Google Scholar] [CrossRef]
  5. McEwan, J.A.; Clulow, A.J.; Nelson, A.; Wang, R.; Burn, P.L.; Gentle, I.R. Influence of Dopant Concentration and Steric Bulk on Interlayer Diffusion in OLEDs. Adv. Mater. Interfaces 2018, 5, 1700872. [Google Scholar] [CrossRef]
  6. Feng, C.; Zheng, X.; Xu, R.; Zhou, Y.; Hu, H.; Guo, T.; Ding, J.; Ying, L.; Li, F. Highly efficient inkjet printed flexible organic light-emitting diodes with hybrid hole injection layer. Org. Electron. 2020, 85, 105822. [Google Scholar] [CrossRef]
  7. Li, B.; Gan, L.; Cai, X.; Li, X.-L.; Wang, Z.; Gao, K.; Chen, D.; Cao, Y.; Su, S.-J. An Effective Strategy toward High-Efficiency Fluorescent OLEDs by Radiative Coupling of Spatially Separated Electron-Hole Pairs. Adv. Mater. Interfaces 2018, 5, 1800025. [Google Scholar] [CrossRef]
  8. Hu, Y.; Song, L.; Zhang, S.; Lv, Y.; Lin, J.; Guo, X.; Liu, X. Improving the Efficiency of Multilayer Organic Light-Emitting Transistors by Exploring the Hole Blocking Effect. Adv. Mater. Interfaces 2020, 7, 2000657. [Google Scholar] [CrossRef]
  9. Wang, M.; Zhu, W.; Yin, Z.; Huang, L.; Li, J. Synergistic effects of Li-doped NiO film prepared by low-temperature combustion as hole-injection layer for high performance OLED devices. Org. Electron. 2020, 85, 105823. [Google Scholar] [CrossRef]
  10. Benor, A.; Takizawa, S.; Pérez-Bolívar, C.; Anzenbacher, P. Efficiency improvement of fluorescent OLEDs by tuning the working function of PEDOT:PSS using UV–ozone exposure. Org. Electron. 2010, 11, 938–945. [Google Scholar] [CrossRef]
  11. Salsberg, E.; Aziz, H. Degradation of PEDOT:PSS hole injection layers by electrons in organic light emitting devices. Org. Electron. 2019, 69, 313–319. [Google Scholar] [CrossRef]
  12. Wu, S.; Han, S.; Zheng, Y.; Zheng, H.; Liu, N.; Wang, L.; Cao, Y.; Wang, J. pH-neutral PEDOT:PSS as hole injection layer in polymer light emitting diodes. Org. Electron. 2011, 12, 504–508. [Google Scholar] [CrossRef]
  13. Li, M.; Wang, J.; Zhang, Y.; Dai, Y.; Chen, L.; Zheng, C.; Lv, W.; Chen, R.; Huang, W. Thermal imprinting and vapor annealing of interfacial layers for high-performance organic light-emitting diodes. J. Mater. Chem. C 2019, 7, 10281–10288. [Google Scholar] [CrossRef]
  14. Chen, S.; Jiang, X.; So, F. Hole injection polymer effect on degradation of organic light-emitting diodes. Org. Electron. 2013, 14, 2518–2522. [Google Scholar] [CrossRef]
  15. Lee, S.; Ha, H.; Lee, J.Y.; Shon, H.K.; Lee, T.G.; Suh, M.C.; Park, Y. Degradation Mechanism of Solution-Processed Organic Light-Emitting Diodes: Sputter Depth-Profile Study. Appl. Surf. Sci. 2021, 564, 150402. [Google Scholar] [CrossRef]
  16. Tyagi, P.; Dalai, M.K.; Suman, C.K.; Tuli, S.; Srivastava, R. Study of 2,3,5,6-tetrafluoro-7,70,8,80-tetracyano quinodimethane diffusion in organic light emitting diodes using secondary ion mass spectroscopy. RSC Adv. 2013, 3, 24553–24559. [Google Scholar] [CrossRef]
  17. Ma, Y.-Y.; Hua, X.-C.; Zhai, T.-S.; Li, Y.-H.; Lu, X.; Duhm, S.; Fung, M.-K. Doped copper phthalocyanine via an aqueous solution process for high-performance organic light-emitting diodes. Org. Electron. 2019, 68, 236–241. [Google Scholar] [CrossRef]
  18. Huang, X.-L.; Zou, J.-H.; Liu, J.-Z.; Jin, G.; Li, J.-B.; Yao, S.-L.; Peng, J.-B.; Cao, Y.; Zhu, X.-H. A high T g small-molecule arylamine derivative as a doped hole-injection/transport material for stable organic light-emitting diodes. Org. Electron. 2018, 58, 139–144. [Google Scholar] [CrossRef]
  19. Gomez, E.F.; Steckl, A.J. Improved Performance of OLEDs on Cellulose/Epoxy Substrate Using Adenine as a Hole Injection Layer. ACS Photonics 2015, 2, 439–445. [Google Scholar] [CrossRef]
  20. Cheng, C.-C.; Chu, Y.-L.; Huang, P.-H.; Yen, Y.-C.; Chu, C.-W.; Yang, A.C.-M.; Ko, F.-H.; Chen, J.-K.; Chang, F.-C. Bioinspired hole-conducting polymers for application in organic light-emitting diodes. J. Mater. Chem. 2012, 22, 18127. [Google Scholar] [CrossRef]
  21. Cheng, C.-C.; Chu, Y.-L.; Chang, F.-C.; Lee, D.-J.; Yen, Y.-C.; Chen, J.-K.; Chu, C.-W.; Xin, Z. New bioinspired hole injection/transport materials for highly efficient solution-processed phosphorescent organic light-emitting diodes. Nano Energy 2015, 13, 1–8. [Google Scholar] [CrossRef]
  22. Shih, H.-K.; Chen, Y.-H.; Chu, Y.-L.; Cheng, C.-C.; Chang, F.-C.; Zhu, C.-Y.; Kuo, S.-W. Photo-Crosslinking of Pendent Uracil Units Provides Supramolecular Hole Injection/Transport Conducting Polymers for Highly Efficient Light-Emitting Diodes. Polymers 2015, 7, 804–818. [Google Scholar] [CrossRef]
  23. Lee, H.; Kwon, Y.; Lee, C. Improved performances in organic and polymer light-emitting diodes using solution-processed vanadium pentoxide as a hole injection layer: OLEDs and PLEDs with solution-processed V2O5. J. Soc. Inf. Disp. 2012, 20, 640–645. [Google Scholar] [CrossRef]
  24. Meyer, J.; Hamwi, S.; Kröger, M.; Kowalsky, W.; Riedl, T.; Kahn, A. Transition Metal Oxides for Organic Electronics: Energetics, Device Physics and Applications. Adv. Mater. 2012, 24, 5408–5427. [Google Scholar] [CrossRef] [PubMed]
  25. Tokito, S.; Noda, K.; Taga, Y. Strongly directed single mode emission from organic electroluminescent diode with a microcavity. Appl. Phys. Lett. 1996, 68, 2633–2635. [Google Scholar] [CrossRef]
  26. Wu, J.; Hou, J.; Cheng, Y.; Xie, Z.; Wang, L. Efficient top-emitting organic light-emitting diodes with a V2O5 modified silver anode. Semicond. Sci. Tech. 2007, 22, 824–826. [Google Scholar] [CrossRef]
  27. Li, J.; Yahiro, M.; Ishida, K.; Yamada, H.; Matsushige, K. Enhanced performance of organic light emitting device by insertion of conducting/insulating WO3 anodic buffer layer. Synth. Met. 2005, 151, 141–146. [Google Scholar] [CrossRef]
  28. Meyer, J.; Hamwi, S.; Bülow, T.; Johannes, H.-H.; Riedl, T.; Kowalsky, W. Highly efficient simplified organic light emitting diodes. Appl. Phys. Lett. 2007, 91, 113506. [Google Scholar] [CrossRef]
  29. Tokito, S.; Noda, K.; Taga, Y. Metal oxides as a hole-injecting layer for an organic electroluminescent device. J. Phys. D Appl. Phys. 1996, 29, 2750–2753. [Google Scholar] [CrossRef]
  30. Matsushima, T.; Kinoshita, Y.; Murata, H. Formation of Ohmic hole injection by inserting an ultrathin layer of molybdenum trioxide between indium tin oxide and organic hole-transporting layers. Appl. Phys. Lett. 2007, 91, 253504. [Google Scholar] [CrossRef]
  31. You, H.; Dai, Y.; Zhang, Z.; Ma, D. Improved performances of organic light-emitting diodes with metal oxide as anode buffer. J. Appl. Phys. 2007, 101, 026105. [Google Scholar] [CrossRef]
  32. Liu, R.; Xu, C.; Biswas, R.; Shinar, J.; Shinar, R. MoO3 as combined hole injection layer and tapered spacer in combinatorial multicolor microcavity organic light emitting diodes. Appl. Phys. Lett. 2011, 99, 093305. [Google Scholar] [CrossRef]
  33. Zhang, X.; You, F.; Zheng, Q.; Zhang, Z.; Cai, P.; Xue, X.; Xiong, J.; Zhang, J. Solution-processed MoOx hole injection layer towards efficient organic light-emitting diode. Org. Electron. 2016, 39, 43–49. [Google Scholar] [CrossRef]
  34. Chiu, T.-L.; Chuang, Y.-T. Spectral observations of hole injection with transition metal oxides for an efficient organic light-emitting diode. J. Phys. D Appl. Phys. 2015, 48, 075101. [Google Scholar] [CrossRef]
  35. Park, M.; Nguyen, T.P.; Choi, K.S.; Park, J.; Ozturk, A.; Kim, S.Y. MoS2-nanosheet/graphene-oxide composite hole injection layer in organic light-emitting diodes. Electron. Mater. Lett. 2017, 13, 344–350. [Google Scholar] [CrossRef]
  36. Kim, S.Y. Bottom-Up Synthesis of MeSx Nanodots for Optoelectronic Device Applications. Adv. Opt. Mater. 2016, 4, 1796–1804. [Google Scholar]
  37. Liu, L.; Li, W.; Zeng, L.; Wang, Y.; Wang, H.; Miao, Y.; Wang, L.; Lu, Z.; Zhang, X. Isopropanol solvent-treated MoS2 nanosheets from liquid phase exfoliation and their applications to solution-processed anode buffer layer of organic light-emitting diode. J. Mater. Sci. Mater. Electron. 2022, 33, 12137–12146. [Google Scholar] [CrossRef]
  38. Le, Q.V. (NH4)2WS4 precursor as a hole-injection layer in organic optoelectronic devices. Chem. Eng. J. 2016, 284, 285–293. [Google Scholar] [CrossRef]
  39. Ohisa, S.; Kagami, S.; Pu, Y.-J.; Chiba, T.; Kido, J. A Solution-Processed Heteropoly Acid Containing MoO3 Units as a Hole-Injection Material for Highly Stable Organic Light-Emitting Devices. ACS Appl. Mater. Interfaces 2016, 8, 20946–20954. [Google Scholar] [CrossRef]
  40. Dong, D.; Lian, L.; Wang, H.; He, G. An efficient solution-processed hole injection layer with phosphomolybdic acid in quantum dot light-emitting diodes. Org. Electron. 2018, 62, 320–326. [Google Scholar] [CrossRef]
  41. Miao, Y.; Yin, M.; Wang, C.; Wei, X.; Wang, Z.; Zhao, M.; Wang, Y.; Jia, Z.; Wang, H.; Zhu, F. Small-size graphene oxide (GO) as a hole injection layer for high-performance green phosphorescent organic light-emitting diodes. J. Mater. Chem. C 2021, 9, 12408–12419. [Google Scholar] [CrossRef]
  42. Hung, L.S.; Liao, L.S.; Lee, C.S.; Lee, S.T. Sputter deposition of cathodes in organic light emitting diodes. J. Appl. Phys. 1999, 86, 4607–4612. [Google Scholar] [CrossRef]
  43. Li, Y.; Tan, L.-W.; Hao, X.-T.; Ong, K.S.; Zhu, F.; Hung, L.-S. Flexible top-emitting electroluminescent devices on polyethylene terephthalate substrates. Appl. Phys. Lett. 2005, 86, 153508. [Google Scholar] [CrossRef]
  44. Yu, S.-Y.; Huang, D.-C.; Chen, Y.-L.; Wu, K.-Y.; Tao, Y.-T. Approaching Charge Balance in Organic Light-Emitting Diodes by Tuning Charge Injection Barriers with Mixed Monolayers. Langmuir 2012, 28, 424–430. [Google Scholar] [CrossRef] [PubMed]
  45. Wu, C.-I.; Lin, C.-T.; Chen, Y.-H.; Chen, M.-H.; Lu, Y.-J.; Wu, C.-C. Electronic structures and electron-injection mechanisms of cesium-carbonate-incorporated cathode structures for organic light-emitting devices. Appl. Phys. Lett. 2006, 88, 152104. [Google Scholar] [CrossRef]
  46. Lee, H.; Cho, S.W.; Han, K.; Jeon, P.E.; Whang, C.-N.; Jeong, K.; Cho, K.; Yi, Y. The origin of the hole injection improvements at indium tin oxide/molybdenum trioxide/N,N′-bis(1-naphthyl)-N,N′-diphenyl-1,1′-biphenyl- 4,4′-diamine interfaces. Appl. Phys. Lett. 2008, 93, 043308. [Google Scholar] [CrossRef]
  47. Park, Y.; Kim, B.; Lee, C.; Hyun, A.; Jang, S.; Lee, J.-H.; Gal, Y.-S.; Kim, T.H.; Kim, K.-S.; Park, J. Highly Efficient New Hole Injection Materials for OLEDs Based on Dimeric Phenothiazine and Phenoxazine Derivatives. J. Phys. Chem. C 2011, 115, 4843–4850. [Google Scholar] [CrossRef]
  48. Xu, L.; Li, Y.; Chen, L.; Wang, J.; Zheng, C.; Qi, Y.; Chen, R.; Huang, W. Annealing Solution-Processed CuSCN Hole Injection Layer for Blue Phosphorescent Organic Light-Emitting Diodes with Extremely Low Efficiency Roll-Off. ACS Sustain. Chem. Eng. 2018, 6, 17178–17183. [Google Scholar] [CrossRef]
  49. Chavhan, S.D.; Ou, T.H.; Jiang, M.R.; Wang, C.W.; Jou, J.H. Enabling High-Efficiency Organic Light-Emitting Diode with Trifunctional Solution-Processable Copper(I) Thiocyanate. J. Phys. Chem. C. Nanomater. Interfaces 2018, 122, 18836–18840. [Google Scholar] [CrossRef]
  50. Chen, Y.; Wu, X.; Liu, Y.; Chen, L.; Li, H.; Wang, W.; Wang, S.; Tian, H.; Tong, H.; Wang, L. Water-soluble pH neutral triazatruxene-based small molecules as hole injection materials for solution-processable organic light-emitting diodes. J. Mater. Chem. C 2019, 7, 7900–7905. [Google Scholar] [CrossRef]
  51. Özdemir, A.D.; Kaiser, S.; Neumann, T.; Symalla, F.; Wenzel, W. Systematic kMC Study of Doped Hole Injection Layers in Organic Electronics. Front. Chem. 2022, 9, 809415. [Google Scholar] [CrossRef]
  52. Kim, J.-S.; Friend, R.H.; Grizzi, I.; Burroughes, J.H. Spin-cast thin semiconducting polymer interlayer for improving device efficiency of polymer light-emitting diodes. Appl. Phys. Lett. 2005, 87, 023506. [Google Scholar] [CrossRef]
  53. Yan, H.; Scott, B.J.; Huang, Q.; Marks, T.J. Enhanced Polymer Light-Emitting Diode Performance Using a Crosslinked-Network Electron-Blocking Interlayer. Adv. Mater. 2004, 16, 1948–1953. [Google Scholar] [CrossRef]
  54. Zhang, Y.; Aslan, K.; Previte, M.J.R.; Geddes, C.D. Metal-enhanced fluorescence from copper substrates. Appl. Phys. Lett. 2007, 90, 173116. [Google Scholar] [CrossRef]
  55. Jang, K.-S.; Kim, D.O.; Lee, J.-H.; Hong, S.-C.; Lee, T.-W.; Lee, Y.; Nam, J.-D. Synchronous vapor-phase polymerization of poly(3,4-ethylenedioxythiophene) and poly(3-hexylthiophene) copolymer systems for tunable optoelectronic properties. Org. Electron. 2010, 11, 1668–1675. [Google Scholar] [CrossRef]
  56. Lee, C.; Kang, D.J.; Kang, H.; Kim, T.; Park, J.; Lee, J.; Yoo, S.; Kim, B.J. Simultaneously Enhancing Light Extraction and Device Stability of Organic Light-Emitting Diodes using a Corrugated Polymer Nanosphere Templated PEDOT:PSS Layer. Adv. Energy Mater. 2014, 4, 1301345. [Google Scholar] [CrossRef]
  57. Baek, M.-G.; Park, S.-G. Differences in ITO Surfaces According to the Formation of Aromatic Rings and Aliphatic Self-Assembled Monolayers for Organic Light-Emitting Diode Applications. Nanomaterials 2021, 11, 2520. [Google Scholar] [CrossRef]
  58. Lo, C.-C.; Sudheendran Swayamprabha, S.; Hsueh, T.-C.; Chavhan, S.D.; Yadav, R.A.K.; Lee, J.-R.; Kesavan, K.K.; Chen, S.-Z.; Wang, C.-W.; Jou, J.-H. Modification effect of hole injection layer on efficiency performance of wet-processed blue organic light emitting diodes. Org. Electron. 2021, 92, 106084. [Google Scholar] [CrossRef]
  59. Su, W.; Chen, R.; Chen, Y. Thermally crosslinkable hole-transporting poly(fluorene- co -triphenylamine) for multilayer polymer light-emitting diodes. J. Polym. Sci. A Polym. Chem. 2011, 49, 352–360. [Google Scholar] [CrossRef]
  60. Fallahi, A.; Alahbakhshi, M.; Mohajerani, E.; Afshar Taromi, F.; Mohebbi, A.R.; Shahinpoor, M. Cationic Water-Soluble Conjugated Polyelectrolytes/Graphene Oxide Nanocomposites as Efficient Green Hole Injection Layers in Organic Light Emitting Diodes. J. Phys. Chem. C 2015, 119, 13144–13152. [Google Scholar] [CrossRef]
  61. Vacca, P.; Petrosino, M.; Chierchia, R.; Guerra, A.; Minarini, C.; Rubino, A. Influence of Electrical, Chemical and Morphological Properties of Inorganic/Organic Interface on Light Emitting Devices Performance. Macromol. Symp. 2007, 247, 333–339. [Google Scholar] [CrossRef]
  62. Deng, W.; Wang, S.; Xiao, Y.; Zhou, N.; Li, X. Solution-processed phosphorus-tungsten oxide film as hole injection layer for application in efficient organic light-emitting diode. Mater. Sci. Semicond. Process. 2018, 85, 106–112. [Google Scholar] [CrossRef]
  63. De Girolamo Del Mauro, A.; Nenna, G.; Bizzarro, V.; Grimaldi, I.A.; Villani, F.; Minarini, C. Analysis of the performances of organic light-emitting devices with a doped or an undoped polyaniline-poly(4-styrenesulfonate) hole-injection layer. J. Appl. Polym. Sci. 2011, 122, 3618–3623. [Google Scholar] [CrossRef]
  64. Choi, M.-R.; Woo, S.-H.; Han, T.-H.; Lim, K.-G.; Min, S.-Y.; Yun, W.M.; Kwon, O.K.; Park, C.E.; Kim, K.-D.; Shin, H.-K.; et al. Polyaniline-Based Conducting Polymer Compositions with a High Work Function for Hole-Injection Layers in Organic Light-Emitting Diodes: Formation of Ohmic Contacts. ChemSusChem 2011, 4, 363–368. [Google Scholar] [CrossRef]
  65. Boileau, N.T. Metal phthalocyanine organic thin-film transistors: Changes in electrical performance and stability in response to temperature and environment. RSC Adv. 2019, 9, 21478–21485. [Google Scholar] [CrossRef]
  66. Li, L.; Guan, M.; Cao, G.; Li, Y.; Zeng, Y. Low operating-voltage and high power-efficiency OLED employing MoO3-doped CuPc as hole injection layer. Displays 2012, 33, 17–20. [Google Scholar] [CrossRef]
  67. Zheng, Q.; You, F.; Xu, J.; Xiong, J.; Xue, X.; Cai, P.; Zhang, X.; Wang, H.; Wei, B.; Wang, L. Solution-processed aqueous composite hole injection layer of PEDOT:PSSþMoOx for efficient ultraviolet organic light-emitting diode. Org. Electron. 2017, 46, 7–13. [Google Scholar] [CrossRef]
  68. Zhu, W.; Chen, X.-L.; Chang, J.; Yu, R.-M.; Li, H.; Liang, D.; Wu, X.-Y.; Wang, Y.; Lu, C.-Z. Doped polyaniline-hybridized tungsten oxide nanocrystals as hole injection layers for efficient organic light-emitting diodes. J. Mater. Chem. C 2018, 6, 7242–7248. [Google Scholar] [CrossRef]
  69. Nagar, M.R.; Yadav, R.A.K.; Dubey, D.K.; Jou, J.-H. Solution Process Feasible Highly Efficient Organic Light Emitting Diode with Hybrid Metal Oxide Based Hole Injection/Transport Layer. MRS Adv. 2019, 4, 1801–1809. [Google Scholar] [CrossRef]
  70. Chan, I.-M.; Hsu, T.-Y.; Hong, F.C. Enhanced hole injections in organic light-emitting devices by depositing nickel oxide on indium tin oxide anode. Appl. Phys. Lett. 2002, 81, 1899–1901. [Google Scholar] [CrossRef]
  71. Corani, A.; Li, M.-H.; Shen, P.-S.; Chen, P.; Guo, T.-F.; El Nahhas, A.; Zheng, K.; Yartsev, A.; Sundström, V.; Ponseca, C.S. Ultrafast Dynamics of Hole Injection and Recombination in Organometal Halide Perovskite Using Nickel Oxide as p-Type Contact Electrode. J. Phys. Chem. Lett. 2016, 7, 1096–1101. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, K.H.L.; Xi, K.; Blamire, M.G.; Egdell, R.G. P -type transparent conducting oxides. J. Phys. Condens. Mat. 2016, 28, 383002. [Google Scholar] [CrossRef] [PubMed]
  73. Xu, X.; Li, L.; Huang, J.; Jin, H.; Fang, X.; Liu, W.; Zhang, N.; Wang, H.; Wang, X. Engineering Ni3+ Cations in NiO Lattice at the Atomic Level by Li+ Doping: The Roles of Ni3+ and Oxygen Species for CO Oxidation. ACS Catal. 2018, 8, 8033–8045. [Google Scholar] [CrossRef]
  74. Liu, A.; Zhu, H.; Guo, Z.; Meng, Y.; Liu, G.; Fortunato, E.; Martins, R.; Shan, F. Solution Combustion Synthesis: Low-Temperature Processing for p-Type Cu:NiO Thin Films for Transparent Electronics. Adv. Mater. 2017, 29, 1701599. [Google Scholar] [CrossRef]
  75. Zheng, J.; Hu, L.; Yun, J.S.; Zhang, M.; Lau, C.F.J.; Bing, J.; Deng, X.; Ma, Q.; Cho, Y.; Fu, W.; et al. Solution-Processed, Silver-Doped NiOx as Hole Transporting Layer for High-Efficiency Inverted Perovskite Solar Cells. ACS Appl. Energ. Mater. 2018, 1, 561–570. [Google Scholar] [CrossRef]
  76. Shan, M.; Jiang, H.; Guan, Y.; Sun, D.; Wang, Y.; Hua, J.; Wang, J. Enhanced hole injection in organic light-emitting diodes utilizing a copper iodide-doped hole injection layer. RSC Adv. 2017, 7, 13584–13589. [Google Scholar] [CrossRef]
  77. Jayabharathi, J. A dodecanethiol-functionalized Ag nanoparticle-modified ITO anode for efficient performance of organic light-emitting devices. RSC Adv. 2017, 7, 38923–38934. [Google Scholar] [CrossRef]
  78. Tam, F.; Goodrich, G.P.; Johnson, B.R.; Halas, N.J. Plasmonic Enhancement of Molecular Fluorescence. Nano Lett. 2007, 7, 496–501. [Google Scholar] [CrossRef]
  79. Sung, H.; Lee, J.; Han, K.; Lee, J.-K.; Sung, J.; Kim, D.; Choi, M.; Kim, C. Controlled positioning of metal nanoparticles in an organic light-emitting device for enhanced quantum efficiency. Org. Electron. 2014, 15, 491–499. [Google Scholar] [CrossRef]
  80. Ko, S.-J.; Choi, H.; Lee, W.; Kim, T.; Lee, B.R.; Jung, J.-W.; Jeong, J.-R.; Song, M.H.; Lee, J.C.; Woo, H.Y.; et al. Highly efficient plasmonic organic optoelectronic devices based on a conducting polymer electrode incorporated with silver nanoparticles. Energy Environ. Sci. 2013, 6, 1949. [Google Scholar] [CrossRef]
  81. Tanaka, T.; Totoki, Y.; Fujiki, A.; Zettsu, N.; Miyake, Y.; Akai-Kasaya, M.; Saito, A.; Ogawa, T.; Kuwahara, Y. Enhanced Red-Light Emission by Local Plasmon Coupling of Au Nanorods in an Organic Light-Emitting Diode. Appl. Phys. Express 2011, 4, 032105. [Google Scholar] [CrossRef]
  82. Qiu, T.; Kong, F.; Yu, X.; Zhang, W.; Lang, X.; Chu, P.K. Tailoring light emission properties of organic emitter by coupling to resonance-tuned silver nanoantenna arrays. Appl. Phys. Lett. 2009, 95, 213104. [Google Scholar] [CrossRef]
  83. Wu, X.; Liu, L.; Yu, T.; Yu, L.; Xie, Z.; Mo, Y.; Xu, S.; Ma, Y. Gold nanoparticles modified ITO anode for enhanced PLEDs brightness and efficiency. J. Mater. Chem. C 2013, 1, 7020. [Google Scholar] [CrossRef]
  84. Jayabharathi, J.; Abirama Sundari, G.; Thanikachalam, V.; Panimozhi, S. Enhanced internal quantum efficiency of organic light-emitting diodes: A synergistic effect. J. Photochem. Photobiol. A Chem. 2018, 364, 577–587. [Google Scholar] [CrossRef]
  85. Feng, J.; Sun, D.; Mei, S.; Shi, W.; Mei, F.; Zhou, Y.; Xu, J.; Jiang, Y.; Wu, L. Plasmonic-Enhanced Organic Light-Emitting Diodes Based on a Graphene Oxide/Au Nanoparticles Composite Hole Injection Layer. Front. Mater. 2018, 5, 75. [Google Scholar] [CrossRef]
  86. da Silva, W.J.; Mohd Yusoff, A.R.B.; Jang, J. GO:PEDOT:PSS for High-Performance Green Phosphorescent Organic Light-Emitting Diode. IEEE Electron Device Lett. 2013, 34, 1566–1568. [Google Scholar] [CrossRef]
  87. Lian, H.; Tang, Z.; Guo, H.; Zhong, Z.; Wu, J.; Dong, Q.; Zhu, F.; Wei, B.; Wong, W.-Y. Magnetic nanoparticles/PEDOT:PSS composite hole-injection layer for efficient organic light-emitting diodes. J. Mater. Chem. C 2018, 6, 4903–4911. [Google Scholar] [CrossRef]
  88. Zhang, D.; Xu, J. Multiple regulation of efficient organic light-emitting diodes using magnetic composite nanoparticles. Opt. Lett. 2019, 44, 3210. [Google Scholar] [CrossRef]
  89. Lin, H.-P.; Zhou, F.; Zhang, X.-W.; Yu, D.-B.; Li, J.; Zhang, L.; Jiang, X.-Y.; Zhang, Z.-L. Efficient hole injection in blue organic light-emitting devices by using a double hole injection layer to improve chromaticity and electrical characteristics. Curr. Appl. Phys. 2011, 11, 853–859. [Google Scholar] [CrossRef]
  90. Zhang, X.; Mo, B.; You, F.; Zhou, X.; Liu, L.; Wang, H.; Wei, B. Electroluminescence enhancement in ultraviolet organic light-emitting diode with graded hole-injection and -transporting structure: Electroluminescence enhancement in ultraviolet organic light-emitting diode with graded hole-injection and -transporting structure. Phys. Status Solidi RRL 2015, 9, 353–357. [Google Scholar]
  91. Tsai, C.-T.; Liu, Y.-H.; Tang, J.-F.; Kao, P.-C.; Chiang, C.-H.; Chu, S.-Y. Effects of novel transition metal oxide doped bilayer structure on hole injection and transport characteristics for organic light-emitting diodes. Synth. Met. 2018, 243, 121–126. [Google Scholar] [CrossRef]
  92. Xu, Y.; Niu, Y.; Gong, C.; Shi, W.; Yang, X.; Wei, B.; Wong, W. High-Performance Inverted Tandem OLEDs with the Charge Generation Layer based on MoO x and Ag Doped Planar Heterojunction. Adv. Opt. Mater. 2022, 10, 2200984. [Google Scholar] [CrossRef]
  93. Wu, S.P.; Kang, Y.; Liu, T.L.; Jin, Z.H.; Jiang, N.; Lu, Z.H. Formation of charge-transfer-complex in organic:metal oxides systems. Appl. Phys. Lett. 2013, 102, 163304. [Google Scholar] [CrossRef]
  94. Huh, Y.H.; Kwon, O.E.; Park, B.; Ji, S.; Lee, S.S.; Lim, J.; An, K.-S. Improved impedance characteristics of all-water-processable triple-stacked hole-selective layers in solution-processed OLEDs. Opt. Express 2016, 24, A846. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, L.; Wang, S.; Li, D.; Fang, Y.; Shen, H.; Li, L.S.; Du, Z. Simultaneous Improvement of Efficiency and Lifetime of Quantum Dot Light-Emitting Diodes with a Bilayer Hole Injection-Layer Consisting of PEDOT:PSS and Solution-Processed WO3. ACS Appl. Mater. Interfaces 2018, 10, 24232–24241. [Google Scholar] [CrossRef] [PubMed]
  96. Lu, L.; Yu, J.; Long, L.; Yu, F.; Zhang, J.; Zhang, H.; Wei, B. Low-voltage and high-stability p-type doped blue organic light-emitting diodes with bilayer hole-injection layers. Phys. Status Solidi A 2011, 208, 2321–2324. [Google Scholar] [CrossRef]
  97. Zhang, X.; You, F.; Liu, S.; Mo, B.; Zhang, Z.; Xiong, J.; Cai, P.; Xue, X.; Zhang, J.; Wei, B. Exceeding 4% external quantum efficiency in ultraviolet organic light-emitting diode using PEDOT:PSS/MoOx double-stacked hole injection layer. Appl. Phys. Lett. 2017, 110, 043301. [Google Scholar] [CrossRef]
  98. Narayan, K.; Varadharajaperumal, S.; Mohan Rao, G.; Manoj Varma, M.; Srinivas, T. Effect of thickness variation of hole injection and hole blocking layers on the performance of fluorescent green organic light emitting diodes. Curr. Appl. Phys. 2013, 13, 18–25. [Google Scholar] [CrossRef]
  99. Zhang, X.; An, J.; Xu, Y.; Wang, Y.; Lu, Y.; Qin, Y.; Lai, W.-Y.; Chen, Y.; Huang, W. Efficient small molecule organic light-emitting diodes fabricated by brush-coating. J. Mater. Chem. C 2021, 9, 2190–2197. [Google Scholar] [CrossRef]
  100. Hussain, S.A.; Dey, B.; Bhattacharjee, D.; Mehta, N. Unique supramolecular assembly through Langmuir—Blodgett (LB) technique. Heliyon 2018, 4, e01038. [Google Scholar] [CrossRef]
  101. C, A.; Colella, M.; Griffin, J.; Kingsley, J.; Scarratt, N.; Luszczynska, B.; Ulanski, J. Slot-Die Coating of Double Polymer Layers for the Fabrication of Organic Light Emitting Diodes. Micromachines 2019, 10, 53. [Google Scholar] [CrossRef] [PubMed]
  102. Yoon, D.G.; Kang, M.; Kim, J.B.; Kang, K.-T. Nozzle Printed-PEDOT:PSS for Organic Light Emitting Diodes with Various Dilution Rates of Ethanol. Appl. Sci. 2018, 8, 203. [Google Scholar] [CrossRef]
  103. Sakata, T.; Kajiya, D.; Saitow, K. Brush Printing Creates Polarized Green Fluorescence: 3D Orientation Mapping and Stochastic Analysis of Conductive Polymer Films. ACS Appl. Mater. Interfaces 2020, 12, 46598–46608. [Google Scholar] [CrossRef] [PubMed]
  104. Yang, Y.; Yang, X.; Yang, W.; Li, S.; Xu, J.; Jiang, Y. Ordered and ultrathin reduced graphene oxide LB films as hole injection layers for organic light-emitting diode. Nanoscale Res. Lett. 2014, 9, 537. [Google Scholar] [CrossRef]
  105. Choi, K.-J.; Lee, J.-Y.; Park, J.; Seo, Y.-S. Multilayer slot-die coating of large-area organic light-emitting diodes. Org. Electron. 2015, 26, 66–74. [Google Scholar] [CrossRef]
  106. C, A.; Dubey, D.K.; Pahlevani, M.; Welch, G.C. Slot-Die Coating of All Organic/Polymer Layers for Large-Area Flexible OLEDs: Improved Device Performance with Interlayer Modification. Adv. Mater. Technol. 2021, 6, 2100264. [Google Scholar] [CrossRef]
  107. Shin, K.-Y.; Kang, M.; Cho, K.H.; Kang, K.-T.; Lee, S.-H. Micro multi-nozzle jet coating of organic thin film for organic light-emitting diode lighting devices. Micro Nano Syst. Lett. 2021, 9, 15. [Google Scholar] [CrossRef]
  108. Lu, H.-W.; Kao, P.-C.; Juang, Y.-D.; Chu, S.-Y. The effects of ultraviolet-ozone-treated ultra-thin MnO-doped ZnO film as anode buffer layer on the electrical characteristics of organic light-emitting diodes. J. Appl. Phys. 2015, 118, 185501. [Google Scholar] [CrossRef]
  109. Lu, H.-W.; Kao, P.-C.; Chu, S.-Y. Effects of Ultra-Thin Al2O3-Doped ZnO Film as Anode Buffer Layer Grown by Thermal Evaporation for Organic Light-Emitting Diodes. ECS J. Solid State Sci. Technol. 2017, 6, R14–R19. [Google Scholar] [CrossRef]
  110. Park, Y.W.; Choi, H.J.; Choi, J.H.; Park, T.H.; Jeong, J.-W.; Song, E.H.; Ju, B.K. Enhanced Power Efficiency of Organic Light-Emitting Diodes using Pentacene on CF4-Plasma-Treated Indium Tin Oxide Anodes. IEEE Electron Device Lett. 2012, 33, 1156–1158. [Google Scholar] [CrossRef]
  111. Kim, J.H. The effect of Ar plasma bombardment upon physical property of tungsten oxide thin film in inverted top-emitting organic light-emitting diodes. Org. Electron. 2011, 12, 285–290. [Google Scholar] [CrossRef]
  112. Choi, S.A.; Kim, K.; Lee, S.J.; Lee, H.; Babajanyan, A.; Friedman, B.; Lee, K. Effects of thermal preparation on Copper Phthalocyanine organic light emitting diodes. J. Lumin. 2016, 171, 149–153. [Google Scholar] [CrossRef]
  113. Ning, Y.; Zhao, S.; Song, D.; Qiao, B.; Xu, Z.; Zhou, Y.; Chen, J.; Swelm, W.; Al-Ghamdi, A. The Improvement of the Performance of Sky-Blue OLEDs by Decreasing Interface Traps and Balancing Carriers with PSVA Treatment. Polymers 2022, 14, 622. [Google Scholar] [CrossRef] [PubMed]
  114. Dagar, J.; Tyagi, P.; Ahmad, R.; Singh, R.; Sinha, O.P.; Suman, C.K.; Srivastava, R. Application of 2D-MoO3 nano-flakes in organic light emitting diodes: Effect of semiconductor to metal transition with irradiation. RSC Adv. 2015, 5, 8397–8403. [Google Scholar] [CrossRef]
Figure 1. Device structure of a typical OLED.
Figure 1. Device structure of a typical OLED.
Nanomaterials 14 00161 g001
Figure 2. Gist of article.
Figure 2. Gist of article.
Nanomaterials 14 00161 g002
Figure 3. The structures of organic materials used as HILs.
Figure 3. The structures of organic materials used as HILs.
Nanomaterials 14 00161 g003
Figure 4. The structures of OLEDs with biomaterial HILs.
Figure 4. The structures of OLEDs with biomaterial HILs.
Nanomaterials 14 00161 g004
Figure 5. The structures of OLEDs with transition metal oxide HILs.
Figure 5. The structures of OLEDs with transition metal oxide HILs.
Nanomaterials 14 00161 g005
Figure 6. Schematic band diagram of (a) ITO/NPB, (b) ITO/(NH4)2WS4 annealed at 250 °C, and (c) ITO/PEDOT:PSS, (d) The current density–voltage characteristics of the hole-only device without HIL, with PEDOT:PSS, and with the 250 °C annealed (NH4)2WS4.(EF: Fermi level, Φh: hole injection barrier, HOMO: highest occupied molecular orbital). Reproduced with permission [38]. Copyright 2015, Elsevier B.V.
Figure 6. Schematic band diagram of (a) ITO/NPB, (b) ITO/(NH4)2WS4 annealed at 250 °C, and (c) ITO/PEDOT:PSS, (d) The current density–voltage characteristics of the hole-only device without HIL, with PEDOT:PSS, and with the 250 °C annealed (NH4)2WS4.(EF: Fermi level, Φh: hole injection barrier, HOMO: highest occupied molecular orbital). Reproduced with permission [38]. Copyright 2015, Elsevier B.V.
Nanomaterials 14 00161 g006
Figure 7. (a) Current density–voltage, (b) luminance–voltage, (c) luminance efficiency–current density, and (d) Normalized electroluminescence spectra of the PEDOT: PSS- and 250 °C annealed-(NH4)2WS4-based OLEDs. Reproduced with permission [38]. Copyright 2015, Elsevier B.V.
Figure 7. (a) Current density–voltage, (b) luminance–voltage, (c) luminance efficiency–current density, and (d) Normalized electroluminescence spectra of the PEDOT: PSS- and 250 °C annealed-(NH4)2WS4-based OLEDs. Reproduced with permission [38]. Copyright 2015, Elsevier B.V.
Nanomaterials 14 00161 g007
Figure 8. The structures of OLEDs with inorganic HILs.
Figure 8. The structures of OLEDs with inorganic HILs.
Nanomaterials 14 00161 g008
Figure 9. UPS spectra cut-off regions of ITO, ITO/MoOx, ITO/PEDOT: PSS and ITO/PEDOT: PSS/MoOx. Inset: energy level diagram of UV OLED with PEDOT: PSS/MoOx bilayer HIL. Reproduced with permission [97]. Copyright 2017, American Institute of Physics.
Figure 9. UPS spectra cut-off regions of ITO, ITO/MoOx, ITO/PEDOT: PSS and ITO/PEDOT: PSS/MoOx. Inset: energy level diagram of UV OLED with PEDOT: PSS/MoOx bilayer HIL. Reproduced with permission [97]. Copyright 2017, American Institute of Physics.
Nanomaterials 14 00161 g009
Figure 10. (a) LB film technology, (b) slot-die coating, (c) nozzle printer, (d) brush-coating. Reproduced under terms of the CC-BY license [100,101,102]. Copyright 2018, The Authors, published by Elsevier Ltd.; Copyright 2019, The Authors, published by MDPI and Copyright 2018, The Authors, published by MDPI. Reproduced with permission [103]. Copyright 2020, American Chemical Society.
Figure 10. (a) LB film technology, (b) slot-die coating, (c) nozzle printer, (d) brush-coating. Reproduced under terms of the CC-BY license [100,101,102]. Copyright 2018, The Authors, published by Elsevier Ltd.; Copyright 2019, The Authors, published by MDPI and Copyright 2018, The Authors, published by MDPI. Reproduced with permission [103]. Copyright 2020, American Chemical Society.
Nanomaterials 14 00161 g010
Table 1. The optical properties of the devices with biomaterial HILs.
Table 1. The optical properties of the devices with biomaterial HILs.
Commercial SymbolWork
Function [eV]
CE
[cd A−1]
PE [lm W−1]EQE [%]Turn-On Voltage [V]Maximum Brightness
[cd m−2]
Ref.
LR-PU--1.240.373.76310[20]
DNA----4.2545,000[19]
PTC-U--3.82.33.543,652[21]
PTC-A--4.62.4347,226[21]
Table 2. The optical properties of OLEDs based on transition metal oxide hole injection materials.
Table 2. The optical properties of OLEDs based on transition metal oxide hole injection materials.
Commercial SymbolWork
Function [eV]
CE
[cd A−1]
PE
[lm W−1]
EQE [%]Turn-On Voltage [V]Maximum
Brightness [cd m−2]
Ref.
MoO3----2.4140,000 [32]
V2O55.66535-5.8153,600 [23]
WO35.6 12.7--[34]
MoO35.6732.625.113.7--[34]
V2O55.8--11--[34]
MoOx--5-5.2-[33]
Table 3. The optical properties of OLEDs based on inorganic hole injection materials.
Table 3. The optical properties of OLEDs based on inorganic hole injection materials.
Commercial SymbolWork
Function (eV)
CE
(cd A−1)
PE
(lm W−1)
EQE (%)Turn-On Voltage (V)Maximum Brightness (cd m−2)Ref.
1-PNA-BPBPOX4.86-2.8-5.7-[47]
PTT: PFFSA------[14]
MeSx5.1---423,300[36]
CH3NH3PbI3--3.210-2.519,110[2]
CuSCN-17.65.69.25.626,424[48]
CuSCN-53.740.313.94.215,180[49]
TAT-4C--39.2112.7-[50]
GO5.1673.1453.9520.633131,600[41]
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

Xing, X.; Wu, Z.; Sun, Y.; Liu, Y.; Dong, X.; Li, S.; Wang, W. The Optimization of Hole Injection Layer in Organic Light-Emitting Diodes. Nanomaterials 2024, 14, 161. https://doi.org/10.3390/nano14020161

AMA Style

Xing X, Wu Z, Sun Y, Liu Y, Dong X, Li S, Wang W. The Optimization of Hole Injection Layer in Organic Light-Emitting Diodes. Nanomaterials. 2024; 14(2):161. https://doi.org/10.3390/nano14020161

Chicago/Turabian Style

Xing, Xiaolin, Ziye Wu, Yingying Sun, Yunlong Liu, Xiaochen Dong, Shuhong Li, and Wenjun Wang. 2024. "The Optimization of Hole Injection Layer in Organic Light-Emitting Diodes" Nanomaterials 14, no. 2: 161. https://doi.org/10.3390/nano14020161

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