Enhanced light extraction efficiency from organic light emitting diodes by insertion of a two-dimensional photonic crystal structure

Similar documents
Ming-Lung CHEN, An-Chi WEI 1, and Han-Ping D. SHIEH

UniMCO 4.0: A Unique CAD Tool for LED, OLED, RCLED, VCSEL, & Optical Coatings

High Efficiency White OLEDs for Lighting

Microcavity OLED using Ag electrodes

Silole Derivative Properties in Organic Light Emitting Diodes

Supplementary Figure 1. OLEDs/polymer thin film before and after peeled off from silicon substrate. (a) OLEDs/polymer film fabricated on the Si

P-224: Damage-Free Cathode Coating Process for OLEDs

Development of OLED Lighting Panel with World-class Practical Performance

Development of Extremely High Efficacy White OLED with over 100 lm/w

Nanostructured super-period gratings and photonic crystals for enhancing light extraction efficiency in OLEDs

Performance Comparison of Bilayer and Multilayer OLED

Organic Electronic Devices

VARIOUS DISPLAY TECHNOLOGIESS

OLED Technology Introduction

Low-haze light extraction from organic light-emitting diode lighting with auxiliary electrode by selective microlens arrays

Spectroscopy on Thick HgI 2 Detectors: A Comparison Between Planar and Pixelated Electrodes

Solution Processable OLEDs. Anna Hayer EuroDisplay /09/2013

High contrast tandem organic light emitting devices employing transparent intermediate nano metal layers and a phase shifting layer

Development of OLED Lighting Applications Using Phosphorescent Emission System

High Power Efficiencies at Record Lifetimes: NOVALED s PIN-OLEDs

Fundamentals of Organic Light Emitting Diode

Organic Electronics 11 (2010) Contents lists available at ScienceDirect. Organic Electronics. journal homepage:


An Overview of OLED Display Technology

Sep 09, APPLICATION NOTE 1193 Electronic Displays Comparison

High Performance White OLEDs Technologies for Lighting

Comparative Analysis of Organic Thin Film Transistor Structures for Flexible E-Paper and AMOLED Displays

Emission behavior of dual-side emissive transparent white organic light-emitting diodes

The Company. A leading OLED player

Single-layer organic-light-emitting devices fabricated by screen printing method

Stacked OLEDs for Lighting Applications - Improvement of the yellow building block

Content. Core Technology (Short introduction) LCMO (Light Controlled Molecular Orientation) technology

Organic light emitting diode (OLED) displays

Novel film patterned retarder utilizing in-plane electric field

Application note. Materials. Introduction. Authors. Travis Burt, Huang ChuanXu*, Andy Jiang* Agilent Technologies Mulgrave, Victoria, Australia

Computer Graphics Hardware

Impact of DMD-SLMs errors on reconstructed Fourier holograms quality

DESIGN AND OPTIMIZATION OF LARGE-AREA OLEDS

OLED for Lighting. Outline

These are used for producing a narrow and sharply focus beam of electrons.

Full down-conversion of amber-emitting phosphor-converted light-emitting diodes with powder phosphors and a long-wave pass filter

Organic light emitting diodes for display technology

Emiflective Display with Integration of Reflective Liquid Crystal Display and Organic Light Emitting Diode

White top-emitting organic light-emitting diodes using one-emissive layer of the DCJTB doped DPVBi layer

Process Dependent Performance of Slot Die Coated OLED-Multilayers (TALK)

Modulation transfer function of a liquid crystal spatial light modulator

JOURNAL OF DISPLAY TECHNOLOGY, VOL. 5, NO. 12, DECEMBER

Development and Mass-Production of an OLED Lighting Panel - Most-Promising Next-Generation Lighting -

Technology White Paper Plasma Displays. NEC Technologies Visual Systems Division

High-resolution screens have become a mainstay on modern smartphones. Initial. Displays 3.1 LCD

The feasible application of low-cost Al/Cu bimetal semitransparent cathode in top-emitting organic light-emitting diode

Advanced Display Technology (continued) Lecture 13 October 4, 2016 Imaging in the Electronic Age Donald P. Greenberg

Design of Organic TFT Pixel Electrode Circuit for Active-Matrix OLED Displays

:: Reduce needs for heat dissipation components. :: Extend battery life in mobile products. :: Save power and reduce heat generation in TVs


PLEASE SCROLL DOWN FOR ARTICLE

[1.9] AMOLED 공정 Introduction OLED Materials Patterning Process Process Equipments

An Overview of the Performance Envelope of Digital Micromirror Device (DMD) Based Projection Display Systems

Phosphorescent OLED Technologies: The Next Wave. Plastic Electronics Conference Oct 9, 2012

Wavelength selective electro-optic flip-flop

Overview of All Pixel Circuits for Active Matrix Organic Light Emitting Diode (AMOLED)

Journal of Organometallic Chemistry

Design Studies For The LCLS 120 Hz RF Gun Injector

Screen investigations for low energetic electron beams at PITZ

Polarizer-free, high-contrast inverted top-emitting organic light emitting diodes: effect of the electrode structure

Liquid Crystal Display (LCD)

Comparison of SONY ILX511B CCD and Hamamatsu S10420 BT-CCD for VIS Spectroscopy

Experimental Study on Dual-Wavelength Distributed Feedback Fiber Laser

2.2. VIDEO DISPLAY DEVICES

Simulation of Mixed-Host Emitting Layer based Organic Light Emitting Diodes

Scalable self-aligned active matrix IGZO TFT backplane technology and its use in flexible semi-transparent image sensors. Albert van Breemen

All-Optical Flip-Flop Based on Coupled Laser Diodes

2x1 prototype plasma-electrode Pockels cell (PEPC) for the National Ignition Facility

P_02_1011:A Novel Pixel Circuit to Compensate for the Degradation of OLED Luminance in High-Resolution AMOLED Displays

Pseudospark-sourced Micro-sized Electron Beams for High Frequency klystron Applications

Reducing tilt errors in moiré linear encoders using phase-modulated grating

IOSR Journal of Engineering (IOSRJEN) ISSN (e): , ISSN (p): Volume 2, PP Organic Led. Figure 1.

Organic Light-Emittin g Devices

Organic Light Emitting Devices

IGM. Development of Vapor Deposition Processes for OLEDs. Bachelor Thesis. Prof. Dr.-Ing. N. Frühauf. 28th of September Alexandru Andrei Lungu

Display Technologies CMSC 435. Slides based on Dr. Luebke s slides

Commissioning the TAMUTRAP RFQ cooler/buncher. E. Bennett, R. Burch, B. Fenker, M. Mehlman, D. Melconian, and P.D. Shidling

Empirical Equations for the Analysis of the Time Dependence of the Luminance Properties of LCD Panels and Backlights for TV Applications

Processing. Electrical Engineering, Department. IIT Kanpur. NPTEL Online - IIT Kanpur

Abstract. Keywords INTRODUCTION. Electron beam has been increasingly used for defect inspection in IC chip

New Pixel Circuit Compensating Poly-si TFT Threshold-voltage Shift for a Driving AMOLED

High Repetition Rate USP Lasers Improve OLED Cutting Results

CNT FIELD EMISSION CATHODE CATALOG. XinRay Systems Inc. April 2014

SINCE more than two decades, Organic Light Emitting

Journal of Luminescence

Principles of Electrostatic Chucks 6 Rf Chuck Edge Design

Semiconductors Displays Semiconductor Manufacturing and Inspection Equipment Scientific Instruments

Joint Development of Ultra-Bright, Inorganic EL Light-Emitting Materials. November 2, 2005 KURARAY CO., LTD.

WITH the rapid development of Gallium Nitride

DELTA MODULATION AND DPCM CODING OF COLOR SIGNALS

1. Publishable summary

Fabrication of Lithium Niobate nanopillars using Focused Ion Beam (FIB)

This work was supported by FINEP (Research and Projects Financing) under contract

OLED display technology

LEDs, New Light Sources for Display Backlighting Application Note

Transcription:

JOURNAL OF APPLIED PHYSICS VOLUME 96, NUMBER 12 15 DECEMBER 2004 Enhanced light extraction efficiency from organic light emitting diodes by insertion of a two-dimensional photonic crystal structure Young Rag Do a) Department of Chemistry, Kookmin University, Seoul 136-702, Korea Yoon-Chang Kim and Young-Woo Song Corporate Research & Development Center, Samsung SDI Co., Ltd., Kyeonggi-Do 449-902, Korea Yong-Hee Lee Department of Physics, Korea Advanced Institute of Science and Technology, Taejon 350-710, Korea (Received 9 July 2004; accepted 16 September 2004) We studied the characteristics of organic light emitting diode (OLED) devices containing two-dimensional (2D) SiO 2 /SiN x photonic crystal (PC) layers. The finite-difference time-domain (FDTD) method was employed for the design and analysis of the PC OLED. Based on the design parameters derived from the FDTD calculations, a 2D PC layer was introduced on the glass substrate of a typical OLED structure by two-step irradiated hologram lithography and reactive ion etching. Experiments showed that incorporation of the PC layer improved the light extraction efficiency by over 50% compared to the conventional OLED, without noticeable degradation in electrical characteristics, under typical operating conditions. This improvement originates from the liberation of the photons trapped in the high-index guiding layers. 2004 American Institute of Physics. [DOI: 10.1063/1.1815049] I. INTRODUCTION Organic light emitting diode (OLED) displays have been actively investigated in recent years on account of their potential applications in flat panel displays and flexible displays. The OLED is a very promising organic electronics component created by combining organic and inorganic materials, such as organic luminescent layers, organic electronhole injection layers, a metal electrode, and a transparent oxide electrode. 1,2 However, OLEDs have yet to be developed that meet the overall efficiency standards required by the industry. Various efforts have been made to improve the performance of OLEDs by elucidating the role, operating mechanism, and fabrication method of each of the organic and inorganic materials used to construct the OLED. Improving the luminous efficiency of OLEDs is the key issue that must be resolved if OLEDs are to be widely commercialized, because both low power consumption and long operational lifetime are required for mobile displays. The external quantum efficiency ex of an OLED device 1 is related to the internal quantum efficiency in and the light extraction efficiency ext ex = in ext. The internal quantum efficiency in of OLEDs has been optimized such that it now approaches the theoretical limit. 3 12 On the other hand, for thin-film structures, only a small fraction of the total photons generated inside the film are generally useable, because of the total internal reflection and wave-guiding effects of high-index layers. In fact, the light extraction efficiency of OLEDs constructed to date is a) Author to whom correspondence should be addressed; FAX: 82-2-910-4415; electronic mail: yrdo@kookmin.ac.kr typically only 20% of the internal efficiency. 13 15 Therefore, in order to further improve the external quantum efficiency, one has to find ways to increase the light extraction efficiency. Thus, if full-color-mobile OLED displays are to become commercially viable, the problem of low extraction efficiency must be overcome. Several out-coupling schemes have been implemented in the hope of improving light extraction from OLEDs. 13 34 Recently, the insertion of a photonic crystal (PC) structure into the semiconductor slab LED structure was proposed and successfully implemented. 35 38 Prompted by those studies on semiconductor LEDs, we have devoted considerable research effort to apply a similar approach to OLEDs without degrading the spatial resolution of the display device. Very recently, we proved that the introduction of two-dimensional (2D) PC layers into OLEDs is an effective way to solve the light-trapping problem. 33,34 In this paper, we provide an overview of the optical consequences of introducing a 2D PC structure into an OLED. This article is organized as follows. In the first section, we summarize previous attempts to improve the extraction efficiency from OLEDs. In the second section, we describe the device architecture used in the simulations and experiments. Analyses based on the finite-difference time-domain (FDTD) method for the design of efficient PC OLEDs are also discussed. In the third part, we discuss the experimental procedure, results, analyses, and approaches that can be taken to optimize the effects of the integrated 2D photonic crystal patterns. Finally, we present our conclusions. II. ENHANCEMENT OF LIGHT EXTRACTION EFFICIENCY Various methodologies have been tried in the ongoing effort to improve the extraction efficiency of thin-film OLED 0021-8979/2004/96(12)/7629/8/$22.00 7629 2004 American Institute of Physics

7630 J. Appl. Phys., Vol. 96, No. 12, 15 December 2004 Do et al. FIG. 1. Schematic diagram of a conventional OLED device and of the light paths of the three modes of emitted light: external mode, glass guided modes, and high-index guided modes. devices. 13 34 Analyses based on the quantum mechanical microcavity model suggested that external coupling efficiencies in excess of 50% should be possible for certain structures. 16 19 This model was used to examine how the external efficiency varies with (i) exciton-to-cathode distance, (ii) the thickness of the indium tin oxide (ITO) layer, and (iii) the refractive index of the substrate in the normal OLED structure. A simple method for improving OLED extraction efficiency is to make the ITO layer sufficiently thin so that the formation of waveguide modes is prevented. 18,19 Simple dielectric capping of top-emitting OLED devices 20,21 has also been proven to be effective for this purpose. Light emission from OLEDs can be understood in terms of interplay between two different interference effects. For example, the microcavity effect can modify the angular distribution of the emitted light, potentially making it possible to funnel more photons within a narrow angle of the normal direction. In addition to the two approaches outlined above (the addition of a layer and the modification of the layer thickness), other techniques have also been developed based on the perturbation of the guided light. These techniques can be divided into two main categories: (1) Reduction of the total internal reflection at the glass-air interface, and (2) modification of the relative number of organic layers to encourage the escape of the photons trapped in substrate modes (glassguided modes) and the high-index guided modes. One simple method to perturb the glass-guided mode is to increase the surface roughness at the substrate-air interface. In this case, photons that would normally undergo total internal reflection can escape from the glass by finding different incident angles upon the second reflection. In LED devices, this effect has been shown to increase the output coupling by as much as 50%. 22 However, this scheme is not applicable to display applications because of the resolution problem. The use of microlenses attached to the front surface has been shown to be a successful method to extract the guided light from the glass substrate. 23 Attaching an ordered array of 10 m diameter microlenses to the glass substrate of an OLED has been shown to increase the light output of the OLED by a factor of 1.5 2.0 compared to the lens-free substrate. Shaping of the device into a mesa structure has also been shown to increase the extraction efficiency by a factor of 2.0. 24 Other simple technique that has been used to perturb the substrate guided mode of OLEDs and thereby enhance the light output is the introduction onto the glass substrate of monolayers of silica spheres as a scattering medium. 25 In another recent study on OLEDs, Tsutsui et al. showed that the external quantum efficiency could be doubled by incorporating a thin layer of very-low-index porous silica n 1.03 in the device. 26 The techniques described above address the problem of enhancing the output coupling of the waveguide modes that are confined to the substrate. The other major fraction of light generated from OLED source flows into the waveguide modes that are confined to the organic/ito layers. If the problem of electrical leakage in OLED devices were less critical, the best way to reduce the high-index-guided modes would be to modify the optical structure at the ITO-glass interface. Several techniques based on this concept have been proposed to modify the distribution of optical modes in the organic/ito layer. 27 32 A direct method to perturb both the ITO and the organic layer is by using a corrugated substrate, which induces a periodic undulation of the ITO/organic thin-film structure. The introduction of one-dimensional gratings with a period of several hundred nanometers has been shown to improve the extraction efficiency by a factor close to 2.0. 28,29 Moreover, Hobson et al. found that further recovery of the trapped light could be provided by the surface plasmon with the help of a periodic grating formed on substrate. 30 Gifford and Hall demonstrated that one-dimensional grating-assisted Ag surface plasmon cross coupling can be used to transmit organic photoluminescence and electroluminescence through an opaque silver layer. 31,32 In reality, the modification of the guided light through the ITO/organic layer is of little practical use in electrically pumped OLEDs due to the strong effect of the surface roughness of ITO on the electrical properties of OLEDs. Therefore, to improve light extraction from OLEDs without causing the electrical problems associated with electrically pumped OLEDs, we introduced a 2D PC slab into the interface between the glass substrate and ITO layer. 33,34 The introduction of a 2D PC into OLEDs to enhance the extraction efficiency of OLEDs is a different research area; hence in the present study we focused our attention on the physical background, fabrication processes, and display applications of 2D PC OLEDs. III. DESIGN OF THE PC OLED STRUCTURE We used the FDTD method to model and design the OLEDs. The FDTD method has been shown to be very effective for modeling complex multilayer OLED devices in which the electron-transporting layer (ETL) is only a few tens of nanometers away from a metallic layer. 38 Figure 1 shows a schematic diagram of the conventional OLED structure used in both the FDTD modeling and OLED fabrication in the present work. The conventional OLED is made up of a multilayer sandwich structure that comprises a glass substrate, an ITO anode, several organic emitting layers, and a metal cathode. The FDTD calculations assume no optical losses from the organic/inorganic transparent layers, and that emitted light is reflected at the metal interface. 39 The light emitting

J. Appl. Phys., Vol. 96, No. 12, 15 December 2004 Do et al. 7631 excitons inside the OLED are modeled as 20 fs long Gaussian oscillating dipole pulses that have a center frequency of 530 nm and a full bandwidth at half maximum of 50 nm. Because the spatial distribution of excitons in real OLEDs is such that they can be termed as incoherent sources, the pulses used in the FDTD calculations will mimic the emission properties provided the spectral distribution is similar to that of the real source. In our FDTD calculations, sufficiently many dipole sources with equal numbers of mutually orthogonal x-, y-, z- polarizations were distributed randomly throughout the active area. In these calculations, the contribution of the surface plasmon mode was automatically included by using a complex refractive index n=0.867 +i6.49 of the Al electrode for both the conventional and PC OLED structures. 29,30 The refractive indices of all of the layers of the conventional OLED used in the FDTD simulations are shown in Fig. 1. As a first step, we calculated the extraction efficiency of a conventional OLED structure, using FDTD analysis. As mentioned above, the light emitted from an OLED device is funneled into three modes: the external propagation modes, the glass guided modes, and high-index guided (ITO/organic guided) modes. The light paths and the calculated fractions of the three modes are shown in Fig. 1. Previous calculations on Lambertian light sources with classical ray optics implied that the fractions of the energy in the air propagation, glass guided, and high-index guided modes are 0.2, 0.3, and 0.5, respectively. 17 Recently, a quantum-mechanical weak microcavity theory was proposed to precisely calculate the efficiencies of each mode for OLED layered structures. 16 According to our FDTD calculations, the external modes that escape the substrate make up only 26.4% of the light emitted from the OLED device. The precise ratio of the three modes in OLEDs is dependent on the optical parameters of each organic or inorganic layer. Similar to the ray optics results reported previously, 17 the FDTD findings indicate that the conventional OLED device has the low external efficiency and high trapping efficiency. For this OLED device, the two confined modes do not contribute to the extracted light because they are either absorbed by the various layers or emitted from the edge of the substrate; these modes must be minimized if high efficiency is to be achieved. Previously we reported 34 that the radiation profiles of both the horizontal dipole source d x,y and the vertical dipole source d z vary with the distance between the active layer and the metallic cathode D a. Given that he previous study showed that most of the radiation from the d z dipoles is emitted below the critical angle and thus cannot escape from the glass into the air, we concentrated on the in-plane dipoles d x and d y. Note that since the image dipole of d x,y induced by the metallic cathode is out of phase with the original d x,y, constructive interference was expected when D a / 4n ETL. Clear enhancement of the vertical radiation was obtained for D a 80 nm, indicating that this thickness satisfies the condition for constructive interference. Thus, we confirmed that the optimal ETL thickness is about 80 nm. The thickness of the ETL and the position of the emitting layer are considered critical design parameters for OLEDs. To enhance the extraction efficiency of the OLED, we FIG. 2. Three-dimensional FDTD calculation structure and parameter definitions for the square-lattice 2D PC OLED. Left diagram, side view; right diagram, top view. introduced into the conventional OLED a 2D PC pattern made of two transparent materials of different refractive index n. As shown schematically in Fig. 2, the 2D PC layer was inserted at the interface between the glass substrate and the ITO electrode. The PC pattern was characterized by a lattice constant of 300 700 nm with a diameter d of 150 500 nm and a height h of 50 400 nm. The FDTD method was employed to estimate the contribution from the leaky modes and/or the diffraction modes of the 2D PC layer. The FDTD analyses showed that the light extraction efficiency depends on the structural parameters of the 2D PC pattern and the thickness of the over-coated high-index layer and ITO. It is thus desirable to design a 2D PC layer with dimensions that maximize the extraction efficiency of the OLED. It has been reported 34 that, regardless of the depth of the pattern, there is a cutoff for the lattice constant cutoff of about 200 nm, below which leaky waves remain trapped in the glass substrate. From the phase matching condition at the PC layer, the cutoff lattice constant can be expressed as follows: cutoff / n eff +1, 1 where n eff is the effective refractive index of the PC layer [ n eff 1.7 1.8 for a SiO 2 n=1.48 SiN x n=1.95 PC layer]. In fact, we observed the same cutoff phenomenon in the FDTD calculations. We found that the extraction efficiency increases with the pattern depth, and takes on the largest values when the lattice constant is similar to the vacuum wavelength. A detailed understanding of the modes and light paths produced in the OLED containing the 2D PC layer can be gained by studying the variation of the calculated intensity of propagating light as a function of time, which enables the creation of a light propagation diagram. Data were acquired for the system with incident light from a vertical dipole source. It has been reported that most of the radiation from vertical d z dipoles is emitted below the critical angle in organic light-emitting diodes, and thus cannot escape from the glass into the air. 34 To confirm the effect of the introduction of a 2D PC structure on the light paths, it is best to examine the main components of the guided modes, such as the vertical dipoles. The calculation results are shown in Fig. 3 as gray-scale maps showing the energy density of the propagating wave. Immediately after the commencement of the emission process in the emitting layer, unguided light escapes from the glass substrate in both the conventional and 2D PC OLEDs. From the time dependence of the propagation path

7632 J. Appl. Phys., Vol. 96, No. 12, 15 December 2004 Do et al. FIG. 3. The results of the numerical model for the time dependence of the light-propagation path in (a) the conventional OLED and (b) the 2D PC OLED =500 nm. in the conventional OLED [Fig. 3(a)], it is clear that highindex guided light has little chance to leak into the glass substrate. The propagation diagram for the 2D PC OLED [Fig. 3(b)] shows the time dependence of the emission process, in which light is leaked by the corrugation in the leaky and/or diffraction modes. This diagram provides qualitative evidence of the enhancement of the time-integrated radiation power produced by the corrugated structure. It is worth pointing out that in this computation the light emitting dipole pulses last only 20 fs and the total photon number available from the pulses is finite and fixed. It is interesting to observe that the time-integrated photon energy extracted into the air increases asymptotically with time. The photons in the ultrashort pulses leak out of the waveguide as the guided photons propagate along the high-index layer. However, one cannot wait indefinitely to collect all the photons. In real OLED displays, the finite pixel size of the device places an upper limit on the photon pulse travel time after which all the photons are effectively outside a given pixel. One simple way to take this into consideration is use the values of the integration at the time when the photon generated at the center of a pixel arrives at the boundary of the pixel. 33 To compare the extraction efficiency of OLEDs, a standard light propagation time of 400 fs, calculated as the average light propagation time needed to reach the boundary of a pixel of size 200 50 m 2, is used. The extraction efficiency can be improved by 50% 80% by adjusting the structural parameters of the 2D PC slab and the pixel size. IV. FABRICATION OF THE PC OLED STRUCTURE The fabrication processes for 2D nanoarrays of lowindex materials 40 and 2D PC slabs have been described in detail elsewhere. 33 The process used for coating low-index materials depends on both the refractive index of the material and the procedure used to prepare the material. Two different types of low-index materials SiO 2 and silicate derivatives were used in the present work. SiO 2 films were coated onto a glass substrate by plasma enhanced chemical vapor deposition (PECVD). Then, the photoresist (PR) film was spin coated on top of the low-index layer. The patterning of the 2D nanoarray of low-index films was carried out using the interference method with two laser beams. A 325 nm linearly polarized He-Cd laser was used as an irradiation source. Fabrication of a 2D periodic structure was carried out by first irradiating the substrate from one angle, and then rotating it by 90 and irradiating it a second time under the same conditions as the first irradiation. In this process, the intensity per exposure lies under the threshold for development, whereas the sum of the two exposures lies above this threshold. With positive PR, development resulted in a 2D FIG. 4. (a) FE-SEM micrographs of the tilted cross-sectional view of the 2D low-index material nanopattern 40 000. (b) FIB-SEM image of a PC OLED device on a 2D PC layer 65 000. square lattice pattern of rods or holes in the resist. The irradiated PR films were developed with tetramethylammonium hydroxide, and SiO 2 films masked with patterned PR films were etched with O 2 +CF 4 gas by using reactive ion etching. Then, low-index films with a square nanorod or nanohole pattern were obtained. Figure 4(a) shows scanning electron microscopy (SEM) images of a tilted cross-sectional view of the 2D nanorod patterns of low-index SiO 2. The nanorods of low-index material look regular and uniform. To form a 2D PC slab, a layer of a high-index material such as SiN x is PECVD deposited onto the 2D low-index nanoarray. PECVD gives pure, homogeneous, planar, low-temperature processed films. Crack-free, thin SiN x films are transparent and have relatively high refractive indexes 1.90 1.95. The image [Fig. 4(b)] of OLED containing a 2D PC slab shows that the fabrication procedure successfully achieved filling and planarization over the low-index dotted structure. As reported, 33 high-index film coatings have rms surface roughness values of 40 50 Å, as measured by AFM topographic imaging. To further decrease the rms surface roughness of the 2D PC slabs used in the present work, we employed a chemical mechanical polishing (CMP) process with a silicate polishing slurry. After polishing, the OLED surface was acceptably smooth for the insertion of a 2D PC layer between the ITO layer and the glass substrate. To investigate the effect of the 2D PC layer experimentally, we fabricated OLEDs on both a conventional substrate and a PC layer covered substrate, side by side for comparison. The ITO layer (thickness 150 nm) was deposited by rf magnetron sputtering with no intentional heating. The sheet resistance of the ITO was 30 /sq. The measured transmission of the ITO-deposited glass substrate at 530 nm was 85%. This value agrees well with the transmission ob-

J. Appl. Phys., Vol. 96, No. 12, 15 December 2004 Do et al. 7633 FIG. 6. Emission spectra of OLEDs with and without a 2D PC layer at the interface between the glass layer and the ITO electrode. FIG. 5. (a) Current density-brightness and current density-efficiency characteristics and (b) Current-voltage curves of OLEDs with and without a 2D PC layer at the interface between the glass layer and the ITO electrode. tained by the thin-film calculation assuming zero absorption in the ITO film. Subsequent application of photolithographical patterning and etching processes to the ITO produced the ITO stripe patterns. The layers of N,N -di naphthalene- 1-y1 -N,N -diphenyl-benzidine, tris 8-hydroxyquinoline aluminum Alq 3 coumarin 6 C6, and Alq 3 were prepared by vacuum sublimation with an open mask. The 30 50 nm thick LiF/Al cathode and the Al cap layer were formed by thermal evaporation using a shadow mask with 3 mm wide stripes oriented at an angle of 90 to the ITO stripes. The OLEDs were encapsulated with 0.7 mm Corning cover glass. Figure 4(b) shows a focused ion beam-type SEM (FIB-SEM) image of the tilted cross-sectional view of the OLED with the 2D SiO 2 /SiN x PC layer. This image shows that the 2D SiO 2 /SiN x PC layer was successfully inserted at the interface between the ITO layer and the glass substrate to give an OLED device with the design shown in Fig. 2. V. RESULTS AND DISCUSSION We first investigated the EL of the conventional and 2D PC OLED devices to confirm that the inclusion of the 2D PC had given the predicted enhancement of extraction efficiency. We then measured the optical properties of the PCcontaining OLEDs to assess their suitability for display applications. EL analyses were carried out to study the electro-optical properties of the 2D SiO 2 /SiN x PC slab. Figure 5(a) shows the current density-luminance (right axis) characteristics of the OLEDs with and without the 2D SiO 2 /SiN x PC layer. The luminance values of the 2D PC and conventional OLEDs measured at the normal angle =0 under 20 ma/cm 2 DC excitation were 3280 and 2180 cd/m 2, respectively. Figure 5(a) also shows the relationship between the current efficiency and the current density of three OLEDs, the conventional OLED and 2D PC OLEDs with nanorod array periods of =350 and 500 nm. The current efficiency in the normal direction at 20 ma/cm 2 of the conventional OLED was 10.9 cd/ A. In contrast, the corresponding current efficiencies for the =350 nm 2D PC OLED was 16.4 cd/a and that of the =500 nm 2D PC was 14.6 cd/a, indicating that the efficiency along the normal direction is 52% and 35% higher in the OLEDs containing PCs. Figure 5(b) shows the I-V characteristics of the OLEDs with and without the 2D PC layer. The I-V curve of the OLED containing the 2D SiO 2 /SiN x photonic slab is very similar to that of the conventional OLED, due to the improved flatness of the PC slab achieved by CMP. The similarity of the electron injection characteristics of the conventional and 2D PC OLEDs demonstrates that the techniques used to introduce the 2D SiO 2 /SiN x PC layer into the OLED are electrically acceptable. The results also clearly confirm that the insertion of the 2D PC layer between the ITO layer and glass substrate improves the external efficiency of OLEDs. The emission spectra measured at constant current density 20 ma/cm 2 for the conventional OLED and two PC OLEDs (lattice constant, =350 and 500 nm) are shown in Fig. 6. The results show that, at the main peak wavelength, the 2D PC OLEDs exhibit much stronger spectra than the conventional OLED. Moreover, the EL spectra of the two 2D PCs are very different, indicating that the 2D PC effect and therefore the structure parameter influence the spectral characteristics. For the conventional OLED, the EL is greatest at 520 nm. A simple analysis based on grating scattering can be applied to the 2D PC layer at the interface between the ITO layer and the glass substrate. On the flat interface inside OLEDs, the wave vector (k ITO : parallel to the surface) of the guided mode in the ITO/organic layer is larger than that k glass of the emitted mode in the glass substrate. According to Snell s law, the guided mode of ITO/organic layers propagates inside of thin films: k glass =2 n glass /,k ITO =2 n ITO /, 2 where n glass =1.50 and n ITO =1.85 are the refractive indices of glass and ITO, respectively. To free the photons trapped in the high-index guided mode, the 2D PC layer is inserted at the glass-ito interface. The guided light can be extracted

7634 J. Appl. Phys., Vol. 96, No. 12, 15 December 2004 Do et al. into the air when the wave vector of the guided mode at the 2D PC layer k eff becomes smaller than k glass. Bragg scattering is governed by the following equation: k eff =2 n eff /, k glass k eff ± m 2 î ± n2 ĵ, 3 where m and n are integers, n eff is the refractive index of the 2D PC layer, is the wavelength at which the EL intensity of the OLED is greatest, and is the lattice constant of the 2D PC. The Bragg scattering modes produced by the 2D PC provide a simple set of characteristics on which to base efforts to enhance the extraction efficiency of OLEDs. Equation (3) indicates that the lattice constant of the 2D PC strongly influences the peak shape of EL spectra such as those shown in Fig. 6. However, although the scattering characteristics can be used as a guide to enhance the extraction efficiency and to alter the emission spectrum, Bragg scattering alone cannot explain the peak shape of the emission spectrum of a 2D PC OLED. In addition to the Bragg scattering and/or leaky modes of a PC, the interference effects of weak microcavity are also believed to influence the emission spectrum of 2D PC OLEDs. The correlation between the emission spectrum and the structural parameters of a 2D PC OLED is considerably more complex. To explain the emission spectra from 2D PC OLEDs qualitatively, the microcavity modes as well as the leaky and/or Bragg scattering modes must be considered. We additionally examined the angular dependence of the emission spectrum because variation in the color with changing viewing angle is undesirable for display applications. Figures 7(a) and 7(b) show the angle dependence of the emission spectra of a conventional OLED and a 2D PC OLED at room temperature, respectively. The results show negligible angle dependence in the peak shape of the conventional OLED, but some variations in the emission spectrum of the 2D PC OLED. The perceived color of the emission spectrum was expressed in terms of the chromaticity coordinates developed by the Commission Internationale de Eclairage (CIE). CIE coordinates were utilized to assess the color changes in the OLEDs. Figure 8 shows the angular dependence of CIE color coordinates for the conventional OLED and the 2D PC green OLED =350 nm on a chromaticity diagram. When viewed from the normal direction, the CIE color coordinates for the conventional OLED and 2D PC green OLED are (0.306, 0.642) and (0.278, 0.666) respectively. This indicates that the light emitted by the 2D PC OLED is of much higher color purity than the conventional standard. The results additionally show that, as the viewing angle is changed over angles up to 70 of normal, the color change ratio of the 2D PC OLED x=6.2%, y =3.3% is similar to that of the conventional OLED x =7.6%, y=2.0%. Improved color purity is therefore another advantage of 2D PC OLEDs over conventional OLEDs. Further improvement of color purity and angular dependence can be achieved by tuning the structural parameters of the 2D PC slab. FIG. 7. Angular dependence of the emission spectra of (a) the conventional OLED and (b) the 2D PC OLED =350 nm. The far-field pattern was measured to further elucidate the physical characteristics underlying the enhancement of extraction efficiency. 33 The data measured over the curved surface of the spheres, were represented on a 2D flat surface x,y by a simple mapping defined by Shin. 41 All of the 2D far-field radiation patterns presented here are based on this mapping method. The far-field photon radiation from the OLED devices was measured directly with a photodiode of area 1 mm 2, located 30 cm from the radiation source. Figure 9 compares the far-field patterns of conventional and PC OLEDs. Figure 9(a) shows that the conventional OLED radiates into air in a manner consistent with a Lambertian source. The far-field radiation of each of the two PC OLEDs ( =350 and 500 nm) shows an emission pattern with four- FIG. 8. Angular dependence of Commission Internationale de Eclairage (CIE) color coordinates of OLEDs with and without a 2D PC layer =350 nm.

J. Appl. Phys., Vol. 96, No. 12, 15 December 2004 Do et al. 7635 FIG. 9. (a c) Measured 2D far-field intensity profiles of (a) a conventional OLED, (b) a PC OLED =350 nm, and (c) a PC OLED =500 nm. (d f) Intensity profiles along the horizontal line (A B) of the 2D far-field intensity profiles for (d) a conventional OLED, (e) a PC OLED =350 nm, and (f) a PC OLED =500 nm. fold symmetry, resulting from the square lattice pattern with fourfold symmetry. The measured 2D far-field radiation pattern of the 2D PC OLEDs can be explained by the simulated 2D pattern obtained by FDTD based simulation. The details of this computational analysis are given elsewhere. 42 The relationship between the experimental and computational diffraction fringes consistently indicates that the 2D far-field radiation pattern of the 2D PC OLED results from the diffraction of light propagating inside the 2D SiO 2 /SiN x PC slab waveguide. Moreover, the 2D PC OLED device shows different angular profiles depending on the lattice constant of the 2D nanoarray. For the sample with a lattice constant of 350 nm, the intensity profile exhibits a double cross shape and the emission is predominantly in the direction normal to the surface. In contrast, the sample with a lattice constant of 500 nm gives rise to a complex intensity profile that comprises a mixture of concentric square, circle, and cross patterns. The far-field profiles cut along the horizontal line (A-B) are plotted in Figs. 9(d), 9(e), and 9(f). Comparison of these profiles reveals that the angle of maximum intensity of emitted light depends strongly on the lattice constant of the 2D periodic structure. Therefore, the far-field images of 2D PC OLEDs are complex due to mixing of the effects of Bragg scattering modes and weak microcavity modes. To compare the extraction efficiencies of the conventional and PC OLEDs, we calculated the total amount of emitted light over all angles by using an integration sphere. As shown in Table I the PC OLEDs show enhanced extraction efficiencies relative to the conventional OLED. This enhancement of the total amount of emitted light is somewhat different from that in the extraction along the normal direction, because the integrated value includes contributions from the anisotropic diffraction fringes in the far-field radiation patterns of the PC OLEDs [Figs. 9(b) and 9(c)]. Inaddition, although the amount of light emitted in the normal TABLE I. Measured relative extraction efficiency of a conventional OLED and two PC OLEDs ( =350 and 500 nm). Relative efficiency normal =0 Relative efficiency integrated sphere Conventional OLED PC OLED =350 nm PC OLED =500 nm 100% 152% 135% 10.9 cd/a 16.4 cd/a 14.6 cd/a 100% 141% 155% FIG. 10. Normalized EL intensity at 50 ma/cm 2 as a function of aging time of the conventional OLED (dotted line), and of the OLED containing a 2D PC layer (solid line). direction simply decreased with increasing the lattice constant from 350 to 500 nm, the integrated values for the two PC OLEDs showed a more complex relationship. For low viewing angles with respect to the substrate surface normal, the 350 nm PC OLED gives a larger integrated extraction efficiency than the 500 nm PC OLED. At other viewing angles, however, the opposite behavior is observed. Therefore, the far-field profile of 2D PC OLEDs is an important determinant of whether such devices can be used in display or lighting applications. To evaluate the reliability of the 2D PC OLED, DC lifetime measurements under accelerated conditions 50 ma/cm 2 were performed on a PC OLED =350 nm and a conventional OLED (Fig. 10). The initial luminescences of the PC OLED and the conventional OLED were 8400 and 5500 cd/m 2, respectively. It is apparent that the insertion of a 2D SiO 2 /SiN x PC layer into an OLED does not alter the aging rate. This result implies that the slight change in surface roughness produced by the insertion of the 2D SiO 2 /SiN x PC layer has little effect on the aging process. This reliability test and the EL data demonstrate that the fabrication technique described above for introducing a 2D SiO 2 /SiN x PC layer into an OLED is electrically acceptable and commercially viable. VI. CONCLUSION We have studied the effects of the 2D PC layer inserted in an OLED with the aim of increasing its extraction efficiency. First we used the FDTD method to calculate the extraction efficiency while varying the lattice constant, depth, and rod size. Introduction of a PC layer into the OLED structure was demonstrated to be an effective way to liberate the photons trapped within the organic ITO waveguide layer. The theoretical predictions of the FDTD calculations were confirmed through the realization of highly efficient OLEDs containing a 2D PC. Two-step irradiated hologram lithography and reactive ion etching were used to create a 2D PC pattern on the glass substrate. To date, the maximum enhancement of the extraction efficiency of an electrically pumped PC OLED that has been achieved is about 50%.

7636 J. Appl. Phys., Vol. 96, No. 12, 15 December 2004 Do et al. However, we believe that further improvement in the OLED efficiency and the far-field radiation pattern can be achieved by optimizing 2D PC OLED structural characteristics such as the period, depth, filling factor, and refractive index of the 2D low-index material dotted array, the thickness and degree of planarization of the high-index layer, and the reflectivity of the anode metal mirror. ACKNOWLEDGMENTS This work was supported by the new faculty research program 2004 of Kookmin University in Korea. The authors would like to thank the members of the electronic materials development team of Samsung SDI, Co. Ltd. for their great help. 1 C. W. Tang and S. A. Van Slike, Appl. Phys. Lett. 51, 913 (1987). 2 J. H. Burroughes, D. D. C. Bradley, A. R. Brown, R. N. Marks, K. Mackay, R. H. Friend, P. L. Burns, and A. B. Holmes, Nature (London) 347, 539(1990). 3 L. S. Hung, L. R. Zneng, and M. G. Mason, Appl. Phys. Lett. 78, 673 (2001). 4 P. K. H. Ho, M. Granström, R. H. Friend, and N. C. Greenham, Adv. Mater. (Weinheim, Ger.) 10, 769 (1998). 5 H. Jiang, Y. Zhou, B. S. Ooi, Y. Chen, T. Wee, Y. L. Lam, J. Huang, and S. Liu, Thin Solid Films 363, 25(2000). 6 K. Murata, S. Ciná, and N. Greenham, Appl. Phys. Lett. 79, 1193 (2001). 7 T. M. Brown, R. H. Friend, I. S. Millard, D. J. Lacey, J. H. Burroughes, and F. Cacialli, Appl. Phys. Lett. 79, 174(2001). 8 L. S. Hung, C. W. Tang, and M. G. Mason, Appl. Phys. Lett. 70, 152 (1997). 9 M. A. Baldo, D. F. O Brien, Y. You, A. Shoustikov, S. Sibley, M. E. Thompson, and S. R. Forrest, Nature (London) 395, 151 (1998). 10 V. Cleave, G. Yahioglu, P. Le Barny, R. H. Friend, and N. Tessler, Adv. Mater. (Weinheim, Ger.) 11, 285(1999). 11 M. A. Baldo, S. M. Lamansky, P. E. Burrows, M. E. Thompson, and S. R. Forrest, Appl. Phys. Lett. 75, 4(1999). 12 Y. Cao, I. Parker, G. Yu, Z. Gang, and A. Heeger, Nature (London) 397, 414 (1999). 13 T. Tsutsui, E. Aminaka, C. P. Lin, and D.-U. Kim, Philos. Trans. R. Soc. London, Ser. A 355, 801 (1997). 14 N. Patel, K. S. J. Cina, and H. Burroughes, IEEE J. Sel. Top. Quantum Electron. 8, 346(2002). 15 K. Meerholz and D. C. Muller, Adv. Funct. Mater. 11, 251 (2001). 16 V. Bulovic, V. B. Khalfin, G. Gi, P. E. Burrows, D. Z. Garbuzov, and S. R. Forrest, Phys. Rev. B 58, 3730 (1998). 17 C. F. Madigan, M.-H. Lu, and J. C. Strum, Appl. Phys. Lett. 76, 1650 (2000). 18 M.-H. Lu and J. C. Sturm, Appl. Phys. Lett. 78, 1927 (2001). 19 M.-H. Lu and J. C. Sturm, J. Appl. Phys. 91, 595 (2002). 20 H. Riel, S. Karg, T. Beierlin, B. Ruhstaller, and W. Rie, Appl. Phys. Lett. 82, 466(2003). 21 H. Riel, S. Karg, T. Beierlin, W. Rie, and K. Neyts, J. Appl. Phys. 94, 5290 (2003). 22 R. Windisch, P. Heremans, A. Knobloch, P. Kiesel, G. H. Döhler, B. Dutta, and G. Borghs, Appl. Phys. Lett. 74, 2256 (1999). 23 S. Möller and S. R. Forrest, J. Appl. Phys. 91, 3324 (2002). 24 G. Gu, D. Z. Garbuzov, P. E. Burrows, S. Vendakesh, S. R. Forrest, and M. E. Thompson, Opt. Lett. 22, 396 (1997). 25 T. Yamasaki, K. Sumioka, and T. Tsutsui, Appl. Phys. Lett. 76, 1243 (2000). 26 T. Tsutsui, M. Yahiro, H. Yokogawa, K. Kawano, and M. Yokoyama, Adv. Mater. (Weinheim, Ger.) 13, 123 (2001). 27 J. M. Lupton, B. J. Matterson, I. D. W. Samuel, M. J. Jory, and W. L. Barnes, Appl. Phys. Lett. 77, 3340 (2000). 28 B. J. Matterson, J. M. Lupton, A. F. Safonov, M. G. Salt, W. L. Barnes, and I. D. W. Samuel, Adv. Mater. (Weinheim, Ger.) 13, 123(2001). 29 P. A. Hobson, S. Wedge, J. A. E. Wasey, I. Sage, and W. L. Barnes, Adv. Mater. (Weinheim, Ger.) 14, 1393 (2002). 30 P. A. Hobson, J. A. E. Wasey, I. Sage, and W. L. Barnes, IEEE J. Sel. Top. Quantum Electron. 8, 378 (2002). 31 D. K. Gifford and D. G. Hall, Appl. Phys. Lett. 80, 3679 (2002). 32 D. K. Gifford and D. G. Hall, Appl. Phys. Lett. 81, 4315 (2002). 33 Y. R. Do, Y. C. Kim, Y.-W. Song, C.-O. Cho, H. Jeon, Y.-J. Lee, S.-H. Kim, and Y.-H. Lee, Adv. Mater. (Weinheim, Ger.) 15, 1214 (2003). 34 Y.-J. Lee, S.-H. Kim, J. Huh, G.-H. Kim, Y.-H. Lee, S.-H. Cho, Y. C. Kim, and Y. R. Do, Appl. Phys. Lett. 82, 3779 (2003). 35 J. D. Joannopoulos, P. R. Villeneuve, and S. Fan, Nature (London) 386, 143 (1997). 36 M. Boroditsky, T. F. Krauss, R. Coccioli, R. Vrijen, R. Bhat, and E. Yablonovitch, Appl. Phys. Lett. 75, 1036 (1999). 37 H.-Y. Ryu, Y.-H. Lee, R. L. Sellin, and D. Bimberg, Appl. Phys. Lett. 79, 3573 (2001). 38 J.-K. Hwang, H.-Y. Ryu, and Y.-H. Lee, Phys. Rev. B 60, 4688 (1999). 39 J.-K. Hwang, H.-Y. Ryu, Y.-J. Lee, and Y.-H. Lee, IEEE J. Sel. Top. Quantum Electron. 8, 238 (2002). 40 Y. R. Do, D.-H. Park, Y.-C. Kim, and Y.-D. Huh, J. Electrochem. Soc. 150, H260 (2003). 41 D.-J. Shin, S.-H. Kim, J.-K. Hwang, H.-Y. Ryu, H.-G. Park, D.-S. Song, and Y.-H. Lee, IEEE J. Quantum Electron. 38, 857 (2002). 42 J. Vučković, M.Lončar, M. Mabuchi, and A. Asherer, IEEE J. Quantum Electron. 38, 850 (2002).