Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Enhanced device performances of a new inverted top-emitting OLEDs with relatively thick Ag electrode

Open Access Open Access

Abstract

To improve the device performances of top-emitting organic light emitting diodes (TEOLEDs), we developed a new inverted TEOLEDs structure with silver (Ag) metal as a semi-transparent top electrode. Especially, we found that the use of relatively thick Ag electrode without using any carrier injection layer is beneficial to realize highly efficient device performances. Also, we could insert very thick overlying hole transport layer (HTL) on the emitting layer (EML) which could be very helpful to suppress the surface plasmon polariton (SPP) coupling if it is applied to the common bottom-emission OLEDs (BEOLEDs). As a result, we could realize noteworthy high current efficiency of approximately ~188.1 cd/A in our new inverted TEOLEDs with 25 nm thick Ag electrode.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

The organic light emitting diodes (OLEDs) have received significant attention due to their self-emissive properties, low power consumption, wide viewing angle, color gamut, fast response time, and flexibility [1–9]. Nowadays, most of the active matrix (AM) type OLEDs (e.g., displays for mobile and television application) have been manufactured by applying top-emitting OLEDs (TEOLEDs) structure to realize high-resolution display with low power consumption [10,11]. Those devices typically adopt two metallic electrodes (e.g., a highly reflective bottom electrode and a semi-transparent top electrode) which results in a micro-cavity effect following the Fabry-Pérot resonator model [10–16]. Meanwhile, to improve the device performances of TEOLEDs, it is necessary to control the optical properties of the metallic electrodes (e.g., reflectance, transmittance, absorptivity, etc.) by optimization of the thickness and composition of the electrode materials. And, such optical properties can be further modified or precisely tuned by deposition of additional organic or inorganic capping layer (CL) on (mainly) top electrode [12,17,18]. Indeed, most of AMOLEDs panel makers are applying TEOLEDs structure with their own semi-transparent cathode system covered with organic CL. In this way, the performance of AMOLEDs has improved dramatically, and panel makers are still trying to change the metal electrode itself or its components to improve its properties further while the generally accepted top electrode has been a magnesium:silver (Mg:Ag) system. On the other hand, there have been lots of reports to use the only Ag as a top electrode because it’s suitable as the semi-transparent electrode for TEOELDs because it has excellent optical properties such as high reflectance, higher transmittance, moderately low absorptivity in a visible wavelength region. However, it was very difficult to utilize the Ag metal without using any additional modification as a cathode due to its relatively high work function values (e.g., Ag (Φa) ≈4.3 eV). To solve such problem, many research groups have tried to improve the electron injection by using n-doped electron transport layer (ETL), (e.g., BPhen:Li, BPhen:Cs), multi-layered cathode system (e.g., LiF/Al/Ag), different alloy composition (e.g., Yb:Ag, Ca:Ag) etc [10,14,19,20]. However, the complexity of top electrode applied in TEOLEDs can easily influence the microcavity effects because the change in its composition easily affect the optical parameters of semi-transparent electrode which causes serious color change of the devices. Therefore, the simple metal electrode (e.g., electrode used by only Ag metal) could be best option to control the optical properties of final devices.

So, we focused on the inverted OLEDs structure in this study, because we could use the simple Ag electrode to control the properties of TEOLEDs devices. Indeed, we investigated the device performances for a new inverted TEOLEDs only by controlling the thickness of top electrode (Ag) without any modification of its composition. Especially, we could apply thick overlying layers having multiple junction structures in between cathode and EML to suppress caused charge trapping sites. As a result, we realized highly efficient inverted TEOLEDs showing ~188.1 cd/A of current efficiency without applying additional light extraction technologies.

2. Experimental

2.1 Materials

To realize the highly efficient TEOLEDs with 2nd microcavity structure, we prepared green phosphorescent OLEDs (PHOLEDs) with top-emitting structure. We used indium tin oxide (ITO, 7 nm)/ silver (Ag, 100 nm)/ITO (7 nm) as a reflective cathode (RC); zinc oxide (ZnO)/ poly(ethylenimine) ethoxylated (PEIE) as an electron injection layer (EIL); bis[2-(2-hydroxyphenyl)pyridinato]-beryllium (Bepp2) as a green host (GH) material, bis(2-phenylpyridine) (acetylacetonato)-iridium (III) [Ir(ppy)2(acac)] as a green dopant (GD) for the emitting layer (EML), N,N′-bis(naphthalen-1-yl)-N,N′-bis(phenyl)benzidine (NPB) as a hole transporting layer (HTL) and/or CL, tris(4-carbazoyl-9-ylphenyl) amine (TCTA) as a HTL and/or an exciton blocking layer (EBL) for preventing exciton quenching. Meanwhile, a 1,4,5,8,9,11-hexaazatriphenylene-hexacarbo-nitrile (HATCN) was utilized as a charge generation layer (CGL) and/or electron accepting layer. Besides, Ag electrode as an anode for inverted TEOLEDs. All the materials were purchased from commercial suppliers and used without purification.

2.2 Device fabrication and measurements

The glass substrate with patterned ITO/Ag/ITO with an active area of 4 mm2 was formed by a photolithography process. The substrate was cleaned with sonication in acetone and isopropyl alcohol, rinsed in deionized water, and treated in UV-ozone to eliminate all the remained organic impurities from the previous processes. All organic materials were deposited by vacuum evaporation under a pressure of ∼5 × 10−7 Torr and deposition rate of organic layers was about 0.5 Å s−1. Ag for electrode was deposited at rates of 0.2 Å s−1 under around 10−6 Torr, respectively. The transmittance and reflectance of Ag electrode were obtained by ultraviolet-visible (UV-vis) spectroscopy (Scinco, S-4100) and (Agilent, Cary 5000), respectively. And the current density–voltage (JV) and luminance–voltage (LV) data of the OLEDs were measured by a Keithley 2635A and Minolta CS-100A, respectively. The electroluminescence (EL) spectra and the Commission Internationale de L’Eclairage (CIE) coordinates were obtained using a Minolta CS-2000A spectroradiometer. The power efficiencies were calculated by integrating angular and spectrally resolved emissions. The optical simulation of TEOLEDs was used for SimOLED and SETFOS 4.3 (Fluxim AG). The morphology of thin film was investigated using on the field emission scanning electron microscopy (FE-SEM, HITACHI, S-4700) and an atomic force microscopy (AFM, Park Systems XE-100).

3. Results and discussion

3.1 Analysis of Ag electrode

In general, light intensity of microcavity devices can be strongly affected by two important optical parameters as following equations [13–16]:

Iext(θ,λ,Z0)=Tt1+RtRb2RtRbcos(ΔϕFP)×(1+Rb+2Rbcos(ΔϕTI))×Iint(θ,λ)
ΔϕFP=φtφb+i4πnidicosθorg,iλ
ΔϕTI=φb+4πnorg,EMLZ0cosθorg,EMLλ
where Rt and Rbrepresent the reflectance at interface of top and bottom electrodes in TEOLEDs and Ttis the transmittance of the top electrode. A Z0 is the distance between dipole of an emitter and a reflective electrode. ΔϕFPand ΔϕTIare total phase difference by Fabry-Pérot and two-beam interference, respectively. φtand φbare phase changes which occur upon reflectance at the interfaces of top electrode-organic and bottom electrode-organic, respectively. And ni and diare refractive index and thickness of i-th organic layer, respectively. In addition, θorg,i is the internal observation angle from the surface normal of the microcavity inside the i-th organic layer and θorg,EML is the angle of the light propagation in the emitting layer. Iint(θ,λ)is the intensity of original emission spectrum in the free space. From this equation, we can understand that the extracted light emission strongly depends on the reflectance and transmittance of the used electrodes. In other words, it is very important that the electrode conditions should be optimized with appropriate optical characteristics (e.g., reflectance and transmittance) to maximize the device performance as well as microcavtiy effect of TEOLEDs.

Meanwhile, the conventional TEOLEDs with a top cathode have focused on Mg:Ag (10:1 = wt.%/wt.%) as a semi-transparent electrode [10,16,21]. However, use of a Mg:Ag electrode can easily lead to a performance degradation caused by low transmittance and high sheet resistance as reported previously. Besides, two component cannot be dispersed perfectly to form a very homogeneous mixture of the same proportions in all parts [21]. So, we thought it would be beneficial if we could use a single composition of the metal (e.g. Ag only) in our devices.

Before we optimize an Ag metal condition as a top semi-transparent electrode, the optical simulation was performed to predict the correlation between the thickness of Ag and the intensity of light. Meanwhile, we additionally deposited CL (60 nm) to modify the optical properties of Ag electrode for better performances as reported before [12]. We could predict the change of radiance according to the thickness of Ag through optical simulation as shown in Fig. 1. As a result, we could estimate that we could extract maximum radiance of ~18.1 x 102 W/sr·m2 [at 510 nm, the peak wavelength of the Ir(ppy)2(acac)] when we use up to 25 nm of Ag as shown in Fig. 1. It means that there might be an optimum condition of optical properties (e.g., reflectance, transmittance, and sheet resistance) at the thickness of 25 nm.

 figure: Fig. 1

Fig. 1 The calculated radiance for Ag samples with different thickness.

Download Full Size | PDF

To investigate the optical properties of Ag metal, we prepared the following samples having the thickness of 15, 20, and 25 nm which are pointed out also in the Fig. 1: (Sample A, B, C w/o CL and Sample D, E, F w/ CL)

Sample A: HATCN (7 nm) / Ag (15 nm)

Sample B: HATCN (7 nm) / Ag (20 nm)

Sample C: HATCN (7 nm) / Ag (25 nm)

Sample D: HATCN (7 nm)/ Ag (15 nm) / CL (60 nm)

Sample E: HATCN (7 nm) / Ag (20 nm) / CL (60 nm)

Sample F: HATCN (7 nm) / Ag (25 nm) / CL (60 nm)

Here, we fabricated only 15, 20, and 25 nm thickness of Ag as previous simulation results. Figure 2 and Fig. 3 show transmittance and reflectance data of various samples fabricated as aforementioned. As we expected, the transmittance values at 550 nm decreased when the thickness varies thinner (e.g., 57.9, 44.2, and 29.2% for Sample A, B, and C, respectively, see also Fig. 2(a) and Table 1.

 figure: Fig. 2

Fig. 2 The transmittance for Ag samples with different thickness (a) without CL (b) with CL.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 The reflectance for Ag samples with different thickness (a) without CL (b) with CL.

Download Full Size | PDF

Tables Icon

Table 1. Electrode characteristics of the different Ag thickness

Those values were all significantly improved to 77.3, 68.5, and 52.9% for Sample D, E, and F, respectively, as shown in Fig. 2(b). Mainly, those transmittance values were enhanced in the whole visible wavelength region (e.g., 450 ~700 nm) after deposition of the CL as well-known principles [12].

In contrast, the reflectance values increased in the order of Sample A (Sample D), Sample B (Sample E), and Sample C (Sample F) at 550 nm for without CL (and with CL), respectively, as shown in Fig. 3. As a result, the reflectance of Ag increased as thickness increases while transmittance decreased as a thickness of Ag increases regardless of CL, which means that reflectance and transmittance are in a trade-off relationship. The absorbance values of each sample are shown in Fig. 4.

 figure: Fig. 4

Fig. 4 The absorbance for Ag samples with different thickness (a) without CL (b) with CL.

Download Full Size | PDF

Anyhow, the transmittance and reflectance of 25nm thick Ag were 34.1% (52.2%) and 63.0% (42.1%) for Sample C (Sample F) at 510 nm, respectively. On the other hand, since the conductivity of Ag is a very important characteristic for the application of the display, we investigated the conductivity and sheet resistance characteristics. Indeed we performed 4-probe measurement and calculated by using van der Pauw method [22]. As a result, the conductivity was 0.6 x 106, 3.7 x 106, and 4.6 x 106 S/m for Sample A, B, and C, respectively. Meanwhile, the sheet resistance was 10.7, 1.4, and 0.9 Ω/□ for Sample A, B, and C, respectively. Very importantly, we found that the sheet resistance can be significantly diminished if we used relatively thick Ag (25 nm) electrode which could be applied even in the large area display applications as shown in Fig. 5 and Table 1.

 figure: Fig. 5

Fig. 5 The conductivity and a sheet resistance of different thickness for Ag without CL.

Download Full Size | PDF

3.2 Optimal simulation for highly efficient inverted TEOLEDs

To realize the highly efficient inverted TEOLEDs having a thick overlying layer (HTL), we performed an optical simulation with the software, SimOLED and SETFOS 4.3 for Fig. 6(a) and Fig. 6(b), respectively, where refractive index (n) and extinction coefficient (k) values of the organic materials used were in the ranges of 1.7-2.0 and 0-0.15, respectively. The detailed device configuration for this study was as follows: [RC / ETL (25 nm) / Bepp2:Ir(ppy)2(acac) (3%, 20 nm) / TCTA (15 nm) / HTL (X nm) / Ag (25 nm)/ CL (60 nm)]. Based on previous results, the Ag thickness fixed at 25nm. From this simulation, we found that it could extract maximum light emission at the certain thickness condition of about 25-50 (1st microcavity) or 160-175 nm (2nd microcavity) for HTL, as shown in Fig. 6(a). Among these conditions, we chose a relatively thick HTL about 174 nm to elongate the distance between metallic electrode and EML, and the total thickness of the 1st microcavity is too thin to prevent negative effects related to manufacturing issues [10]. Also, we found that the portion of SPP coupling mode is not changed that serious if we apply a metallic electrode (Ag) as an anode, as shown in Fig. 6(b). Meanwhile, the inverted structure adopting transparent electrode (e.g., TCO: transparent conducting oxide) showed a significant level of SPP coupling as reported by Kim et al [23,24]. However, we thought this structure is still beneficial to get rid of all the possible metal diffusion from the top electrode rather than the suppression of SPP coupling. Based on these results, we designed inverted TEOELDs with different thickness of Ag electrode as an anode.

 figure: Fig. 6

Fig. 6 (a) Simulated radiance contour plots of the microcavity OLED with green emitter as functions of the HTL versus ETL thickness. The solid red box designates the optimal thickness range of HTL/ETL to satisfy the 2nd microcavity condition; (b) optical simulation result according to the thickness of HTL for inverted TEOLEDs.

Download Full Size | PDF

3.4 Device characteristics for Inverted TEOLEDs

To confirm the the effect on the optical properties Ag acting as an anode, we fabricated the inverted TEOLEDs, when we applied different thickness of Ag electrode as follows (Devices A–C, see also Fig. 7):

 figure: Fig. 7

Fig. 7 Energy band diagram of the inverted TEOLEDs in this study.

Download Full Size | PDF

Device A: RC / ZnO (15 nm) / PEIE (10 nm) / Bepp2: Ir(ppy)2(acac) (3%, 20 nm) / TCTA (15 nm) / NPB (60 nm) /HATCN (7 nm) / NPB(100 nm) / HATCN(7 nm) / Ag (15 nm) / CL (60 nm)

Device B: RC / ZnO (15 nm) / PEIE (10 nm) / Bepp2: Ir(ppy)2(acac) (3%, 20 nm) / TCTA (15 nm) / NPB (60 nm) / HATCN (7 nm) / NPB(100 nm) / HATCN (7 nm) / Ag (20 nm) / CL (60 nm)

Device C: RC / ZnO (15 nm) / PEIE (10 nm) / Bepp2: Ir(ppy)2(acac) (3%, 20 nm) / TCTA (15 nm) / NPB (60 nm) / HATCN (7 nm) / NPB(100 nm) / HATCN(7 nm) / Ag (25 nm) / CL (60 nm)

Here, we applied an inverted structure using ZnO/PEIE on the underlying electrode. In addition, we applied p-n-p-n multiple junctions to improve the overall charge balance originating from the thick HTL at the central EML, aforementioned [25,26].

Figure 8(a) shows the JVL characteristics of the devices fabricated in this study. The turn-on voltage (Von) to reach 1cd/m2 were recorded 4.0, 3.5, and 3.5 V for Device A, B, and C, respectively. Meanwhile, the operation voltage (Vop) to reach 1,000 cd/m2 were 6.0, 5.8, and 5.7 V for Device A, B, and C, respectively. (see also Table 2). At a given constant luminance of 1000 cd/m2, the current and power efficiencies were 91.8 cd/A and 48.1 lm/W for Device A, 139.7 cd/A and 75.4 lm/W for Device B, 155.8 cd/A and 85.9 lm/W for Device C, respectively. And the maximum current and power efficiencies were 123.7cd/A and 86.3 lm/W for Device A, 152.1 cd/A and 106.4 lm/W for Device C, 188.1 cd/A and 131.3 lm/W for Device C, respectively, as shown in Fig. 8(b) and Table 2. With this result, we found that the fabricated TEOLEDs with thick overlying layer could relatively realize a good charge balance, which is beneficial to improve the device stability originating from the metal diffusion. Those results are well consistent with our previous simulation results as shown in Fig. 1.

 figure: Fig. 8

Fig. 8 (a) J-V-L, (b) L-efficiency characteristics of Devices A-C.

Download Full Size | PDF

Tables Icon

Table 2. Device characteristics of the inverted TEOLEDs

To understand the reason of the performance change upon variation of Ag thickness, we proceeded FE-SEM and AFM of the samples and obtained the surface morphology as shown in Fig. 9 and Fig. 10, respectively. From this approaches, we found that the 15-nm-thick Ag film contains numbers of dark regions at SEM photographic image, plausibly due to a grain boundary, a crack, or nano-voids, which may affect the device performances related to resistivity, as aforementioned. In other words, the deposition of very thin Ag layer on CL may result in an insufficient island-like growth of Ag as shown in Fig. 9(a).

 figure: Fig. 9

Fig. 9 SEM images of Ag films with (a) 15 nm and (b) 25 nm thick condition (Scale bar: 0.5 μm). The yellow circles in (a) indicate the defect site such as crack, voids, etc.

Download Full Size | PDF

 figure: Fig. 10

Fig. 10 AFM images for each Ag sample.

Download Full Size | PDF

While, the 25 nm-thick Ag film showed relatively uniform Ag film with much denser island-like morphology due to a much better step coverage which also cause an order of magnitude reduced sheet resistance as well as much improved conductivity, as previously commented (see also Fig. 9(b) and Fig. 10) [27–29]. Besides, we could also confirm that the surface morphology of Ag film seems to be much denser if the thickness of Ag film increases as shown in the AFM images of Fig. 10.

Figure 11 shows the normalized EL spectra for inverted TEOELDs in this study. Very interestingly, we observed relatively narrow EL spectra with full width half maximum (FWHM) about 34, 30, and 24 nm, for Device A, B, and C, respectively.(see also Table 2.)

 figure: Fig. 11

Fig. 11 EL spectra fabricated inverted TEOLEDs.

Download Full Size | PDF

Especially, we found that the 25 nm thick Ag electrode exhibited significantly narrow EL emission intensity due to the strongest microcavity effect as compared with others [10–16,30].

Surprisingly, we could obtain efficient TEOLEDs without using any reactive electrode. Instead of, we found that Ag electrode can efficiently act as an anode without any modification.

4. Conclusions

In this study, we demonstrated the device performances depending on the different thickness of Ag electrode. From this results, we found that the simple metal electrode (only used Ag electrode) could effectively act as a semi-transparent top electrode of inverted TEOLEDs, and without any hole injection layers, we obtained relatively good charge balance by using multiple p-n junctions. Consequently, the Device C (with 25 nm thick Ag condition) showed the outstanding performances resultants in their relatively superior optical properties (e.g., high reflectance, good interface conditions, and conductivity) as compared with than others (Device A and B). From this result, we confirm that the usage of the only Ag electrode could utilize the highly efficient inverted TEOELDs, which also could help to resolve the electrode issues for TEOLEDs.

Funding

Ministry of Trade, Industry and Energy (MOTIE) (10079974).

References and links

1. C. W. Tang, S. A. VanSlyke, and C. H. Chen, “Electroluminescence of doped organic thin films,” J. Appl. Phys. 65(9), 3610–3616 (1989). [CrossRef]  

2. T. Furukawa, H. Nakanotani, M. Inoue, and C. Adachi, “Dual enhancement of electroluminescence efficiency and operational stability by rapid upconversion of triplet excitons in OLEDs,” Sci. Rep. 5(1), 8429 (2015). [CrossRef]   [PubMed]  

3. S.-R. Park, S. M. Kim, J.-H. Kang, J.-H. Lee, and M. C. Suh, “Bipolar host materials with carbazole and dipyridylamine groups showing high triplet for blue phosphorescent organic light emitting diodes,” Dyes Pigmemts 141, 217–224 (2017). [CrossRef]  

4. S. J. Cha, Y. R. Cho, and M. C. Suh, “Green phosphorescent organic light emitting diodes with simple structure to realize an extremely low operating voltage,” Synth. Met. 200, 143–147 (2015). [CrossRef]  

5. S.-R. Park, D. H. Shin, S.-M. Park, and M. C. Suh, “Benzoquinoline-based fluoranthene derivatives as electron transport materials for solution-processed red phosphorescent organic light-emitting diodes,” RSC Advances 7(45), 28520–28526 (2017). [CrossRef]  

6. S. Watanabe, N. Ide, and J. Kido, “High-efficiency green phosphorescent organic light-emitting devices with chemically doped layers,” Jpn. J. Appl. Phys. 46(3A), 1186–1188 (2007). [CrossRef]  

7. Y. H. Sohn, G. J. Moon, K. H. Choi, Y. S. Kim, and K. C. Park, “Effect of TFT mobility variation in the threshold voltage compensation circuit of the OLED display,” J. Inform. Displ. 18(1), 25–30 (2017). [CrossRef]  

8. J.-H. Jou, S. Kumar, A. Agrawal, T.-H. Li, and S. Sahoo, “Approaches for fabricating high efficiency organic light emitting diodes,” J. Mater. Chem. C Mater. Opt. Electron. Devices 3(13), 2974–3002 (2015). [CrossRef]  

9. H. Sasabe and J. Kido, “Development of high performance OLEDs for general lighting,” J. Mater. Chem. C Mater. Opt. Electron. Devices 1(9), 1699–1707 (2013). [CrossRef]  

10. B. W. Lim, H. S. Jeon, and M. C. Suh, “Top-emission organic light emitting diodes with lower viewing angle dependence,” Synth. Met. 189, 57–62 (2014). [CrossRef]  

11. S. Chen and H.-S. Kwok, “Alleviate microcavity effects in top-emitting white organic light-emitting diodes for achieving broadband and high color rendition emission spectra,” Org. Electron. 12(12), 2065–2070 (2011). [CrossRef]  

12. J. W. Huh, J. Moon, J. W. Lee, D.-H. Cho, J.-W. Shin, J. Han, J. Hwang, C. W. Joo, J. Lee, and H. Y. Chu, “The Optical Effects of Capping Layers on the Performance of Transparent Organic Light-Emitting Diodes,” IEEE Photonics J. 4(1), 39–47 (2012). [CrossRef]  

13. B. Y. Jung, N. Y. Kim, C. Lee, C. K. Hwangbo, and C. Seoul, “Control of resonant wavelength from organic light-emitting materials by use of a Fabry-Perot microcavity structure,” Appl. Opt. 41(16), 3312–3318 (2002). [CrossRef]   [PubMed]  

14. H. Cho, C. Yun, and S. Yoo, “Multilayer transparent electrode for organic light-emitting diodes: tuning its optical characteristics,” Opt. Express 18(4), 3404–3414 (2010). [CrossRef]   [PubMed]  

15. S. Chen, L. Deng, J. Xie, L. Peng, L. Xie, Q. Fan, and W. Huang, “Recent developments in top-emitting organic light-emitting diodes,” Adv. Mater. 22(46), 5227–5239 (2010). [CrossRef]   [PubMed]  

16. B. Pyo, C. W. Joo, H. S. Kim, B. H. Kwon, J. I. Lee, J. Lee, and M. C. Suh, “A nanoporous polymer film as a diffuser as well as a light extraction component for top emitting organic light emitting diodes with a strong microcavity structure,” Nanoscale 8(16), 8575–8582 (2016). [CrossRef]   [PubMed]  

17. J. H. Lee, S. Hofmann, M. Furno, M. Thomschke, Y. Kim, B. Lüssem, and K. Leo, “Systematic investigation of transparent organic light-emitting diodes depending on top metal electrode thickness,” Org. Electron. 12(8), 1383–1388 (2011). [CrossRef]  

18. M. J. Park, S. K. Kim, R. Pode, and J. H. Kwon, “Low absorption semi-transparent cathode for micro-cavity top-emitting organic light emitting diodes,” Org. Electron. 52, 153–158 (2018). [CrossRef]  

19. R. Springer, B. Y. Kang, R. Lampande, D. H. Ahn, S. Lenk, S. Reineke, and J. H. Kwon, “Cool white light-emitting three stack OLED structures for AMOLED display applications,” Opt. Express 24(24), 28131–28142 (2016). [CrossRef]   [PubMed]  

20. H. S. Jeon, B. Pyo, H. S. Park, S.-R. Park, and M. C. Suh, “Improved out-coupling efficiency of organic light emitting diodes by manipulation of optical cavity length,” Org. Electron. 20, 49–54 (2015). [CrossRef]  

21. S.-K. Kwon, E.-H. Lee, K.-S. Kim, H.-C. Choi, M. J. Park, S. K. Kim, R. Pode, and J. H. Kwon, “Efficient micro-cavity top emission OLED with optimized Mg:Ag ratio cathode,” Opt. Express 25(24), 29906–29915 (2017). [CrossRef]   [PubMed]  

22. R. Chwang, B. J. Smith, and C. R. Crowell, “Contact size effects on the van der Pauw method for resistivity and Hall coefficient measurements,” Solid-State Electron. 17, 1217–1227 (1972).

23. J.-B. Kim, J.-H. Lee, C.-K. Moon, K.-H. Kim, and J.-J. Kim, “Highly enhanced light extraction from organic light emitting diodes with image blurring and good color stability,” Org. Electron. 17, 115–120 (2015). [CrossRef]  

24. J.-B. Kim, J.-H. Lee, C.-K. Moon, S.-Y. Kim, and J.-J. Kim, “Highly enhanced light extraction from surface plasmonic loss minimized organic light-emitting diodes,” Adv. Mater. 25(26), 3571–3577 (2013). [CrossRef]   [PubMed]  

25. W. S. Jeon, J. S. Park, L. Li, D. C. Lim, Y. H. Son, M. C. Suh, and J. H. Kwon, “High current conduction with high mobility by non-radiative charge recombination interfaces in organic semiconductor devices,” Org. Electron. 13(6), 939–944 (2012). [CrossRef]  

26. S. H. Cho, S. W. Pyo, and M. C. Suh, “Low voltage top-emitting organic light emitting devices by using 1,4,5,8,9,11-hexaazatriphenylene-hexacarbonitrile,” Synth. Met. 162(3-4), 402–405 (2012). [CrossRef]  

27. E. Najafabadi, K. A. Knauer, W. Haske, and B. Kippelen, “High-performance inverted top-emitting green electrophosphorescent organic light-emitting didoes with a modified top Ag anode,” Org. Electron. 14(5), 1271–1275 (2013). [CrossRef]  

28. Y. C. Han, M. S. Lim, J. H. Park, and K. C. Choi, “ITO-free flexible organic light-emitting diode using ZnS/Ag/MoO3 anode incorporating a quasi-perfect Ag thin film,” Org. Electron. 14(12), 3437–3443 (2013). [CrossRef]  

29. K.-S. Lee, I. H. Kim, C. B. Yeon, J. W. Kim, S. J. Yun, and G. E. Jabbour, “Thin metal electrodes for semitransparent organic photovoltaic,” ETRI J. 35(4), 587–593 (2013). [CrossRef]  

30. C. Wu, C. Chen, C. Lin, and C. Yang, “Advanced Organic Light-Emitting Devices for Enhancing Display Performances,” J. Displ. Technol. 1(2), 248–266 (2005). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (11)

Fig. 1
Fig. 1 The calculated radiance for Ag samples with different thickness.
Fig. 2
Fig. 2 The transmittance for Ag samples with different thickness (a) without CL (b) with CL.
Fig. 3
Fig. 3 The reflectance for Ag samples with different thickness (a) without CL (b) with CL.
Fig. 4
Fig. 4 The absorbance for Ag samples with different thickness (a) without CL (b) with CL.
Fig. 5
Fig. 5 The conductivity and a sheet resistance of different thickness for Ag without CL.
Fig. 6
Fig. 6 (a) Simulated radiance contour plots of the microcavity OLED with green emitter as functions of the HTL versus ETL thickness. The solid red box designates the optimal thickness range of HTL/ETL to satisfy the 2nd microcavity condition; (b) optical simulation result according to the thickness of HTL for inverted TEOLEDs.
Fig. 7
Fig. 7 Energy band diagram of the inverted TEOLEDs in this study.
Fig. 8
Fig. 8 (a) J-V-L, (b) L-efficiency characteristics of Devices A-C.
Fig. 9
Fig. 9 SEM images of Ag films with (a) 15 nm and (b) 25 nm thick condition (Scale bar: 0.5 μm). The yellow circles in (a) indicate the defect site such as crack, voids, etc.
Fig. 10
Fig. 10 AFM images for each Ag sample.
Fig. 11
Fig. 11 EL spectra fabricated inverted TEOLEDs.

Tables (2)

Tables Icon

Table 1 Electrode characteristics of the different Ag thickness

Tables Icon

Table 2 Device characteristics of the inverted TEOLEDs

Equations (3)

Equations on this page are rendered with MathJax. Learn more.

I e x t ( θ , λ , Z 0 ) = T t 1 + R t R b 2 R t R b cos ( Δ ϕ F P ) × ( 1 + R b + 2 R b cos ( Δ ϕ T I ) ) × I int ( θ , λ )
Δ ϕ F P = φ t φ b + i 4 π n i d i cos θ o r g , i λ
Δ ϕ T I = φ b + 4 π n o r g , E M L Z 0 cos θ o r g , E M L λ
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.