investigation of blue phosphorescent organic light-emitting diode instability

160
Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability By Kevin P. Klubek Submitted in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy Supervised by Professor Ching W. Tang Department of Chemical Engineering Arts, Sciences and Engineering Edmund A. Hajim School of Engineering and Applied Sciences University of Rochester Rochester, New York 2014

Upload: others

Post on 11-Sep-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

Investigation of Blue Phosphorescent Organic Light-Emitting

Diode Instability

By

Kevin P. Klubek

Submitted in Partial Fulfillment of the

Requirements for the Degree

Doctor of Philosophy

Supervised by

Professor Ching W. Tang

Department of Chemical Engineering

Arts, Sciences and Engineering

Edmund A. Hajim School of Engineering and Applied Sciences

University of Rochester

Rochester, New York

2014

Page 2: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

ii

Dedicated To

My Wonderful Wife Kimberly

And

Amazing Children Nathaniel, Zachary, and Julia

Page 3: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

iii

BIOGRAPHICAL SKETCH

Kevin Paul Klubek was born in 1972 in Lackawanna, New York. In 1996 he

received a Bachelor of Science degree in Medicinal Chemistry from the University of

Buffalo, after which he accepted a position as synthetic chemist at the Kodak Research

Laboratories in Rochester, New York to work on the synthesis of organic materials for

organic light emitting diode applications. From 2007, he began pursuing his doctorate

part-time in Chemical Engineering at the University of Rochester under the supervision

of Professor Ching W. Tang. In 2010 he left Kodak to purse his doctorate full-time. In the

same year, he was awarded a NSF Graduate Research Fellowship. In 2012 he received

his Master of Science Degree. In 2013, he received a NSF East Asia and Pacific Summer

Institute award and spent eight weeks conducting OLED research with Professor Liang-

Sheng Liao at Suzhou University, China. His field of study was in organic materials and

device physics related to Organic Light-Emitting Diodes.

PUBLICATIONS AND PAPERS SUBMITTED FOR

PUBLICATION

1. H. Wang, K. P. Klubek, C.W. Tang, Appl. Phys. Lett., 93 (2008) 093306.

2. K.P. Klubek, C.W. Tang, L.J. Rothberg, Org. Electron., 15 (2014) 1312.

3. K.P. Klubek, S.C. Dong, L.S. Liao, C.W. Tang, L.J. Rothberg, Org. Electron.,

submitted June 2014.

Page 4: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

iv

ACKNOWLEDGEMENTS

I would like to first thank my thesis advisor and mentor Professor Ching W. Tang

for providing me with his guidance, support and remarkable insights in conducting

research related to OLED materials, device fabrication, and device physics. I have

learned so much in all the time we have worked together. It was truly an honor and a

privilege to have had the opportunity to work with Professor Tang.

I would like to thank Professor Shaw H. Chen for his support and encouragement

in pursuing my PhD studies at the Department of Chemical Engineering and for the

research collaboration we have undertaken in the DOE project. I also thank Professor

Lewis J. Rothberg, Professor Alexander A. Shestopalov, and Professor Todd D. Krauss

for serving on my thesis committee. I would also like to express my gratitude to Professor

Rothberg for his perspicacity and support throughout our collaboration.

I would like to thank Mr. Joseph Madathil for teaching and helping me with

everything related to vacuum systems/equipment and electronics. It was a pleasure

working with Joe. I could not have imagined building the KOMET coater without all of

his help, guidance and support. The KOMET most definitely did not come from Mars. I

also would like to thank Mike Culver, who was extremely helpful in sharing his technical

and equipment building experience with me. He was always willing to help in any way,

especially with providing all types of lab equipment and supplies. I also like to thank Dr.

Ralph H. Young for our many stimulating conversations. He was extremely patient and

helpful in guiding me in the field of electrostatics. I would like to express my sincere

Page 5: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

v

gratitude to Professor Liang-Sheng (Larry) Liao and Dr. Shou-Cheng Dong of Suzhou

University for hosting my visit in Suzhou, and for the continuing collaboration. I would

also like to thank Dr. David Weiss for providing valuable advice related to experiments

and for providing help with proofreading this thesis and in presentations.

My gratitude and appreciation also goes to Master Machinists Mr. John Miller,

Mr. John Gresty, Mr. Richard Fellows and Mr. Jim Despard for all their assistance and

training that they provided to me in the machine shop.

To all the Professors in the Chemical Engineering Department and to Mrs. Sandra

Willison, Mrs. Gina Eagan, Mr. Larry Kuntz, Mrs. Tiffany Landers, Ms. Jennifer Condit,

and Victoria Heberling, I would like to say “thank you” for your support and assistance

with administrative issues.

I also want to acknowledge my fellow lab-mates for their collaboration and

camaraderie: Minlu Zhang, Mohan Ahluwalia, Wei Xia, Hao Lin, Hui Wang, Hsiang Ning

(Sunny) Wu, Felipe Angel, Yung-Hsin (Thomas) Lee, Mathew Smith, Charles Chan, Sang-

Min Lee, Lichang Zeng, Jason Wallace, Prashant Kumar Singh, Laura Ciammaruchi,

Jonathan Welt, Guy Mongelli, Chris Favaro, Erik Glowacki, Lisong Xu, Aanand

Thiyagarajan, Michael Beckley, Thao Nguyen, Meng-huan (Kinneas) Ho, William Finnie,

Sihan (Jonas) Xie, and all of our group alumni.

To Eastman Kodak Company I am grateful for the financial support and “time-off” in

my pursuit of graduate studies.

In addition to all the academic fun, I truly enjoyed all our Group events, going to

the Cantonese House, The Distillery, Sheridans, and cookout Tuesdays. Tuesday will

Page 6: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

vi

forever be known as burger day! Mike Culver’s “burger time, burger time” will ring like

a bell. Mike, thank you for the awesome cookies. I’ll never be able to eat another burger

or hotdog without getting a cookie afterwards. And there was tennis! It was great fun

playing tennis with “Bazooka” Joe Madathil, Felipe Angel, Minlu Zhang, Lisong Xu,

Prashant Kumar Singh, and Aanand Thiyagarajan. And to Laura Ciammaruchi, I would

like to thank you for introducing me to squash and for all the fun we had. I’m still not

sure if we’ve really played the “Queen” match yet. Oopha…

To my Mom and Dad, Marian and Gerald, Mother-in-law Diana, and all my

siblings and their families, I thank you all for your support and encouragement.

This thesis would not have been possible without the unconditional love, support

and encouragement provided by my wife Kimberly and children Nathaniel, Zachary, and

Julia. I am extremely grateful to have such an amazing and wonderful family. I love you

all so much.

Page 7: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

vii

ABSTRACT

Phosphorescent emitters have played a critical role in the advance of organic

light-emitting diode (OLED) technology. Among the phosphorescent emitters, the

cyclometalated complexes with iridium (III) as the central atom are the most well

developed and successfully commercialized. They are also among the most efficient.

With chelating ligands specifically designed for color tuning, RGB (red, green and blue)

emitters with nearly 100% (internal) quantum efficiency, defined as photons generated

per injected electron, have been demonstrated. The operating lifetime of the

phosphorescent OLED devices, however, remains an issue, particularly for the blue

devices.

This thesis focuses on two phosphorescent blue dopants based on iridium; bis(4,6-

difluorophenyl-pyridinato-N,C2) picolinate iridium (III) (FIrpic), and tris[1-(2,6-

diisopropylphenyl)-2-phenyl-1H-imidazole] iridium (III) (Ir(iprpmi)3). FIrpic is perhaps

the most well-known blue phosphorescent dopant that has demonstrated high quantum

efficiency, but it has been found to produce short operational lifetimes. Ir(iprpmi)3 was

chosen for this study because there were claims in patent literature that it provides

excellent stability in OLED devices.

Using FIrpic as the blue dopant in the emitter layer, we have investigated the

dependence of OLED performance, including device lifetime, on the compositions of the

emitter layer (host-dopant) and the adjacent electron/hole transport layers. Regardless of

the choice of the host materials in the emitter layer, or the materials in the transport

Page 8: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

viii

layers, it is shown that the stability of OLED devices is poor (<12 hours) for devices

using FIrpic as the blue phosphorescent dopant. Recombination that occurs on the host or

transport materials is also found to be detrimental to device lifetime.

By tracking voltage and efficiency during operation for model devices with well

known tris(8-hydroxyquinolinato) aluminum (Alq) as the emitter, it is shown that in

addition to its instability in charge recombination processes, FIrpic is also unstable with

respect to hole-transport processes. This indicates that the FIrpic radical cation itself is

unstable. It is also found that the host 3,3'-bis(N-carbazolyl)biphenyl (mCBP) is unstable

to electron-hole recombination processes.

The photophysical properties of Ir(iprpmi)3 were studied. This material has a low

ionization potential of 4.8 eV, which is indicative of a material possessing strong electron

donor characteristics. Electron donating materials are typically associated with hole

injecting or hole transporting materials. By adjusting the concentration of Ir(iprpmi)3

within the host mCBP, we show that Ir(iprpmi)3 is capable of trapping (at low

concentration) and transporting (at high concentration) holes. In this way, the location of

the recombination zone (RZ) is modulated. At low concentrations the RZ is confined near

the hole-transport layer whereas at high concentration the RZ is shifted towards the

electron-transport layer. The external quantum efficiency, EQE, defined as photons

exiting the device per injected electron, can reach >20% as long as the RZ is adjacent a

charge transport material that has a higher triplet energy than Ir(iprpmi)3. Due to

molecular structural differences, device lifetime using Ir(iprpmi)3 as the emitting dopant

is significantly improved compared to FIrpic. However, the lifetime is also highly

Page 9: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

ix

dependent on the choice of material for the host and charge transport layers.

Page 10: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

x

CONTRIBUTORS AND FUNDING SOURCES

This work was supervised by a dissertation committee consisting of Professors Ching

W. Tang (advisor) and Alexander A. Shestopalov of the Department of Chemical

Engineering, Professors Lewis J. Rothberg and Todd D. Krauss (chair) of the Department of

Chemistry.

For Chapter 2: The organic boat design was based on an initial design by fellow lab-

member Sang-min Lee. Mr. Joseph Madathil and Professor Tang both contributed to the

design and building of the vacuum coater for fabricating OLED devices.

For Chapter 3: My fellow lab-mate Lisong Xu fabricated the undoped TCTA device.

The data analysis was conducted in part by Professor Ching W. Tang.

For Chapter 4: The data analyses were conducted in part by Professor Ching W. Tang

and Professor Lewis J. Rothberg and were published in 2014, in an article listed in the

Biographical Sketch.

For Chapter 5: Dr. Shou-Cheng Dong of Suzhou University in China helped to

acquire photophysical, electrochemical, and theoretical data related to the iridium

cyclometalated dopant Ir(iprpmi)3. The analyses were conducted in part by Dr. Dong,

Professor Tang, Professor Rothberg, and Professor Liang-Sheng Liao of Suzhou University in

China and were submitted for publication in 2014, in an article listed in the Biographical

Sketch.

All other work conducted for this dissertation was completed by Kevin P. Klubek

independently.

This thesis was funded by the National Science Foundation (NSF) under Grant

Page 11: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xi

No. 1316969, the NSF Graduate Research Fellowship Program under Grant No. 0935947

and the Department of Energy (DOE) under Award No. DE-EE0003296. Any opinions,

findings, and conclusions or recommendations expressed in this material are those of the

author and do not necessarily reflect the views of the NSF or DOE.

Page 12: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xii

TABLE OF CONTENTS

Biographical Sketch iii

Acknowledgements iv

Abstract vii

Contributors and Funding Source x

Table of Contents xii

List of Figures xvii

List of Tables xxvi

CHAPTER 1 BACKGROUND AND INTRODUCTION 1

1.1 Organic Electroluminescence 1

1.2 Organic Light-Emitting Diode Operating Principle 1

1.3 Exciton Formation 4

1.4 Device Efficiency 5

1.5 Iridium Cyclometalated Complexes 6

1.6 Device Lifetime 8

1.7 Degradation 10

1.8 Objectives and Outline of Thesis 12

Page 13: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xiii

References 15

CHAPTER 2 DEVICE FABRICATION EXPERIMENTAL DETAILS 21

2.1 Thermal Evaporation Coater 21

2.2 The KOMET Coater 23

2.3 Organic Boats 29

2.4 Device Fabrication 31

2.5 OLED Performance Measurements 34

2.6 Device Lifetime 34

2.7 Materials 37

References 42

CHAPTER 3 FIRPIC AS THE LIGHT-EMITTING DOPANT FOR

PHOSPHORESCENT ORGANIC LIGHT-EMITTING DIODES 45

3.1 Introduction 45

3.2 Experimental 47

3.3 Mixed Host 47

3.3.1 TCTA:UGH3 (1:1) Mixed Host with Varied FIrpic Concentration 49

Page 14: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xiv

3.3.2 Mixed TCTA:UGH3 (x:y) Hosts with Fixed Flrpic Concentration 52

3.3.3 Effect of Electron TransportMaterials 54

3.3.4 Device Lifetime 57

3.4 Single Host 58

3.4.1 TCTA Single Host and FIrpic Concentration Series 59

3.4.2 TCTA Single Host and UGH3 as Electron Transport Material 62

3.4.3 TCTA Host and BAlq as Electron Transport Material 67

3.4.4 mCBP Host and FIrpic Concentration Series 69

3.4.5 mCBP Host and BAlq as Electron Transport Material 73

3.4.6 mCBP Host and NPB as Hole Transport Layer 76

3.5 Conclusions 78

References 81

CHAPTER 4 CHARGE TRANSPORT THROUGH FIRPIC 86

4.1 Introduction 86

4.2 Experimental 87

4.3 Results and Discussion 88

Page 15: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xv

4.3.1 Undoped Hole Transport Layers 89

4.3.2 FIrpic Doped into mCBP Adjacent to the Anode 91

4.3.3 FIrpic Doped mCBP Sandwiched Between Undoped Hole Transport

Materials 94

4.3.4 FIrpic Doped into NPB 96

4.4 Conclusions 97

References 98

CHAPTER 5 INVESTIGATING BLUE PHOSPHORESCENT IRIDIUM DOPANT

WITH PHENYL-IMIDAZOLE LIGANDS 100

5.1 Introduction 100

5.2 Materials and Methods 101

5.3 Results and Discussion 103

5.3.1 Photophysical and Electrochemical Measurements 103

5.3.2 Low EQE OLEDs 105

5.3.3 High EQE OLEDs 107

5.3.4 Probing the Recombination Zone 111

Page 16: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xvi

5.3.5 High EQE, Low Voltage OLEDs 114

5.3.6 Device Lifetime 116

5.4 Conclusions 119

References 120

CHAPTER 6 SUMMARY AND FUTURE RESEARCH 124

6.1 Summary 124

6.2 Future Research 127

References 132

Page 17: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xvii

LIST OF FIGURES

Figure 1.1 OLED configuration and molecular structures used by

Tang and VanSlyke. 2

Figure 1.2 State-of-the-art OLED device structure. 3

Figure 1.3 Blue [Ir(F2ppy)3], green [Ir(ppy)3] and red [Ir(piq)3] Ir-

based phosphorescent dopants and maximum emission

peaks (max). 7

Figure 1.4 Commission Internationale De L'Eclairage (CIE) 1931

color coordinates. 9

Figure 2.1 Basic design of a thermal evaporation coater. 22

Figure 2.2 KOMET coater. 24

Figure 2.3 Inside the KOMET coater chamber. (a) Top of the

chamber, (b) bottom of the chamber, (c) area for depositing

organic materials, and (d) area for depositing

metals/inorganics. 25

Figure 2.4 KOMET coater top plate. 26

Figure 2.5 Features below the KOMET bottom plate. 27

Page 18: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xviii

Figure 2.6 Control cabinet 28

Figure 2.7 Picture of components used to build organic boat. 30

Figure 2.8 Components for assembling organic boat. 30

Figure 2.9 ITO pattern on substrate. 32

Figure 2.10 Pictures of OLED active area and encapsulation technique

(a) ITO pattern, organic deposited through circular mask,

and metal deposited through L-shaped mask, (b) An OLED

device with 4 lit pixels, and (c) Vacuum assembly for

encapsulating devices. 33

Figure 2.11 Output from John Burtis software program (a) current

density (mA/cm2), (b) drive voltage (V), (c) luminance

(cd/m2), (d) CIEx coordinate, (e) CIEy coordinate, and (f)

EQE or p/e (photon/electron). 35

Figure 3.1 Molecular structures for materials discussed in Chapter 3. 48

Figure 3.2 FIrpic concentration series for TCTA:UGH3 mixed host

OLED devices. 49

Figure 3.3 OLED performance data at 5 mA/cm2 for varying FIrpic

concentration in mixed host devices; Anode/35 nm

TAPC/40 nm TCTA+UGH3 (1:1)+x% FIrpic/30 nm

Page 19: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xix

TmPyPB/20 nm BPhen/Cathode, x=9–26%; (a) EL spectra

at 5 mA/cm2, (b) EQE vs. current density, (c) current

density vs. voltage. 51

Figure 3.4 Varying the mixed host doping ratio using FIrpic as

emitting dopant. 52

Figure 3.5 OLED performance data at 5 mA/cm2 for varying mixed

host composition while keeping FIrpic at 14%; Anode/35

nm TAPC/40 nm TCTA+UGH3 (x:y)+14% FIrpic/30 nm

TmPyPB/20 nm BPhen/1 nm Cathode, x+y=86%; (a) EL

spectra at 5 mA/cm2, (b) EQE vs. current density, (c)

current density vs. voltage. 53

Figure 3.6 Removing TmPyPB and using BAlq or BPhen next to the

LEL for TCTA:UGH3 mixed host devices. 54

Figure 3.7 OLED performance data at 5 mA/cm2 for mixed host

devices with alternate ETL’s; Anode/35 nm TAPC/40 nm

TCTA+UGH3 (1:1)+12% FIrpic/ETL/Cathode. ETL is 50

nm BPhen or 10 nm BAlq/50 nm BPhen; (a) EL spectra at

5 mA/cm2, (b) EQE vs. current density, (c) current density

vs. voltage. 55

Page 20: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xx

Figure 3.8 HOMO, LUMO, Triplet Energy levels for materials used to

fabricate OLED devices in Chapter 2. Triplet Energies are

in parenthesis. Shaded boxes indicate typical electron

transport materials. All values in eV. 57

Figure 3.9 Device lifetime taken at 5 mA/cm2 for mixed-host devices;

Anode/35 nm TAPC/40 nm TCTA+UGH3 (1:1)+12%

FIrpic/ETL/Cathode. ETL is 50 nm BPhen, 10 nm BAlq/50

nm BPhen, or 30 nm TmPyPB/20 nm BPhen. 58

Figure 3.10 FIrpic concentration series for TCTA single host OLED

devices. 59

Figure 3.11 EL spectrum at 20 mA/cm2 showing exciplex of

TCTA:TmPyPB for undoped device. Anode/30 nm

TAPC/30 nm TCTA/30nm TmPyPB/Cathode. 61

Figure 3.12 Removing TmPyPB and using UGH3 or BPhen next to the

LEL for single host devices with FIrpic concentration at

1% or 12%. 62

Figure 3.13 EL spectra at 5 mA/cm2 for devices using UGH3 as the

ETL and 1% FIrpic in the LEL. Anode/35 nm TAPC/40 nm

TCTA+12% FIrpic/y nm UGH3/50 nm BPhen/Cathode,

y=0–20 nm. 66

Page 21: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xxi

Figure 3.14 BAlq or BPhen next to the LEL for TCTA single host

devices. 67

Figure 3.15 FIrpic concentration series for mCBP single host devices. 69

Figure 3.16 OLED performance for devices with mCBP as host and 2-

18% FIrpic, Anode/35 nm TAPC/mCBP+x% FIrpic/30 nm

TmPyPB/20 nm BPhen/Cathode. 71

Figure 3.17 EL spectrum at 20 mA/cm2 showing exciplex of

mCBP:TmPyPB. Anode/75 nm mCBP/75 nm

TmPyPB/Cathode. 72

Figure 3.18 Removing TmPyPB and using BAlq or BPhen next to the

LEL for mCBP single host devices with the FIrpic

concentration at 6% or 12%. 73

Figure 3.19 Using BAlq as ETL and NPB as HTL at FIrpic

concentrations of 6% and 12%. 76

Figure 4.1 Molecular structures and HOMO/LUMO energy levels of

materials used in devices. Units in electronvolts (eV). 87

Figure 4.2 EL spectra at 20 mA/cm2 for the following device

architecture: ITO/1.0 nm MoOx/75 nm HTL/75 nm Alq/1.0

nm LiF/100 nm Al, where the HTL’s are: Ref [75 nm

NPB], A1 [75 nm mCBP], B4 [37.5 nm mCBP+12%

Page 22: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xxii

FIrpic|37.5 nm NPB], C1 [37.5 nm mCBP|27.5 nm

mCBP+12% FIrpic|10 nm NPB], D1 [37.5 nm mCBP|37.5

nm mCBP+12% FIrpic], E1 [75 nm NPB+6% FIrpic]. 89

Figure 4.3 OLED performance for devices with an undoped HTL

adjacent to MoOx. ITO/MoOx/HTL/Alq/LiF/Al. Ref[NPB],

A1[mCBP], A2[mCBP/NPB]. (a) Initial current density-

voltage-luminance (J-V-L) characteristics, (b) Device

lifetime at 20 mA/cm2, (c) Operational voltage rise at 20

mA/cm2. 90

Figure 4.4 OLED performance for devices with FIrpic doped mCBP

adjacent to MoOx. ITO/MoOx/mCBP+x%

FIrpic/Alq/LiF/Al. B1[1%], B2[3%], B3[6%], B4[12%]. (a)

Initial current density-voltage-luminance (J-V-L)

characteristics, (b) Device lifetime at 20 mA/cm2, (c)

Operational voltage rise at 20 mA/cm2. 92

Figure 4.5 OLED performance for devices with FIrpic doped mCBP

spaced away from MoOx and for a device with FIrpic

doped in NPB. ITO/MoOx/mCBP/HTL/NPB/Alq/LiF/Al:

C1[mCBP+12%], ITO/MoOx/HTL/Alq/LiF/Al]: D1[mCBP/mCBP

+12% FIrpic] and E1[NPB+6% FIrpic]. (a) Initial current density-

voltage-luminance (J-V-L) characteristics, (b) Device

Page 23: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xxiii

lifetime at 20 mA/cm2, (c) Operational voltage rise at 20

mA/cm2. 95

Figure 5.1 Molecular structures and energy level diagram of the

materials (values in eV). 102

Figure 5.2 UV-Vis absorption and PL spectra of Ir(iprpmi)3 in

CH2Cl2. Inset: PL spectrum of Ir(iprpmi)3 in 2-MeTHF at

77 K. 103

Figure 5.3 Simulated frontier molecular orbitals of Ir(iprpmi)3. (a)

HOMO electron density, (b) ball and stick model, (c)

LUMO electron density. 104

Figure 5.4 OLED performance data for 6% and 12% concentrations of

Ir(iprpmi)3 with NPB as HTL and BAlq as ETL. Anode/40

nm TAPC/30 nm mCBP+ 6% or 12% Ir(iprpmi)3/40 nm

BAlq/Cathode. (a) EL spectra at 5 mA/cm2, (b) current

density vs. voltage, (c) EQE vs. current density. 106

Figure 5.5 OLED performance data for Ir(iprpmi)3 concentration

series with TAPC HTL and BAlq ETL. Anode/40 nm

TAPC/30 nm mCBP+ x% Ir(iprpmi)3/40 nm

BAlq/Cathode. x=0%–24%. (a) EL spectra at 5 mA/cm2,

(b) current density vs. voltage, (c) EQE vs. current density. 108

Page 24: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xxiv

Figure 5.6 OLED performance data when probing the recombination

zone with 6% Ir(iprpmi)3. Anode/40 nm TAPC/x nm

mCBP+6% Ir(iprpmi)3/y nm mCBP/40 nm BAlq/Cathode.

[x=25 nm, y=5 nm], [x=20 nm, y=10 nm],[x=10 nm, y=20

nm], and [x=5 nm, y=25 nm]: (a) current density vs.

voltage, (b) EQE vs. current density. 112

Figure 5.7 OLED performance data when probing the recombination

zone with 25% Ir(iprpmi)3. Anode/40 nm TAPC/x nm

mCBP+25% Ir(iprpmi)3/y nm mCBP/40 nm BAlq/Cathode.

[x=25 nm, y=5 nm],[x=20 nm, y=10 nm], [x=10 nm, y=20

nm], and [x=5 nm, y=25 nm]: (a) current density vs.

voltage [ inset: voltage vs mCBP layer thickness of devices

under current densities of 1, 5, and 10 mA/cm2] (b) EQE

vs. current density. 113

Figure 5.8 OLED performance data for Ir(iprpmi)3 concentration

series with TmPyPB ETL. Anode/40 nm TAPC/30 nm

mCBP + x% Ir(iprpmi)3/40 nm TmPyPB/Cathode. x=3%–

24% (a) current density vs. voltage, (b) EQE vs. current

density. 115

Figure 5.9 Lifetime testing at 5 mA/cm2 for device structure,

Anode/40 nm HTL/30 nm mCBP+x% Ir(iprpmi)3/40 nm

Page 25: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xxv

ETL/Cathode, [NPB-6% dopant-BAlq], [NPB-12% dopant-

BAlq], [TAPC-7% dopant-BAlq], [TAPC-13% dopant-

BAlq], [TAPC-6% dopant-TmPyPB]. 116

Figure 6.1 Different types of blue phosphorescent dopants. 129

Page 26: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xxvi

LIST OF TABLES

Table 2.1 Organic materials discussed throughout this thesis.

HOMO/LUMO/triplet energies were taken from literature.

HOMO energies are based on ionization potential, LUMO

energies based on difference between optical bandgap and

HOMO energy, and triplet energy based on

phosphorescence spectra taken at 77K. 38

Table 3.1 OLED performance data at 5 mA/cm2 for varying FIrpic

concentration in mixed host devices; Anode/35 nm

TAPC/40 nm TCTA+UGH3 (1:1)+x% FIrpic/30 nm

TmPyPB/20 nm BPhen/Cathode, x=9–26%. 51

Table 3.2 Device data taken at 5 mA/cm2 for single host TCTA and

FIrpic concentrations 3-6%. 60

Table 3.3 OLED performance data at 5 mA/cm2 for devices using

UGH3 as the ETL and 12% FIrpic in the LEL; Anode/35

nm TAPC/40 nm TCTA+12% FIrpic/x nm TmPyPB/y nm

UGH3/z nm BPhen/Cathode. 63

Table 3.4 OLED performance data at 5 mA/cm2 for devices using

UGH3 as the ETL and 1% FIrpic in the LEL. Anode/35 nm

Page 27: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xxvii

TAPC/40 nm TCTA+12% FIrpic/x nm TmPyPB/y nm

UGH3/z nm BPhen/Cathode. 65

Table 3.5 OLED performance data for devices using TCTA as single

host and BAlq as the ETL. Anode/35 nm TAPC/40 nm

TCTA+12% FIrpic/x nm BAlq/50 nm BPhen/Cathode. 67

Table 3.6 OLED performance at 5 mA/cm2 for devices using mCBP

as host and varying the FIrpic concentration from 2–18%. 70

Table 3.7 OLED performance data at 5 mA/cm2 for devices using

mCBP as the single host and BAlq as the ETL. Anode/35

nm TAPC/40 nm mCBP+12% FIrpic/x nm BAlq/50 nm

BPhen/Cathode. 74

Table 3.8 OLED performance data at 5 mA/cm2 for devices using

mCBP as the single host and BAlq as the ETL. Anode/35

nm TAPC/40 nm mCBP+6% FIrpic/x nm BAlq/50 nm

BPhen/Cathode. 75

Table 3.9 OLED performance data at 5 mA/cm2 for devices using

NPB as HTL, BAlq as ETL, and x% FIrpic. Anode/35 nm

NPB/40 nm mCBP+x% FIrpic/y nm BAlq/50 nm

BPhen/Cathode. 77

Page 28: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

xxviii

Table 4.1 OLED device structure and initial EQE and voltage at 20

mA/cm2. ITO/1 nm MoOx/HTL/75 nm Alq/1 nm LiF/100

nm Al. 88

Page 29: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

1

CHAPTER 1

BACKGROUND AND INTRODUCTION

1.1. ORGANIC ELECTROLUMINESCENCE

Light-emission resulting from electrical stimulation of an organic molecule is

known as organic electroluminescence (EL). In 1953, Bernanose et al. [1] adsorbed

acridine dyes on a sheet of cellophane and observed organic EL by applying a strong

alternating electric field that required a threshold voltage of 400-800 V. In 1963 Pope et

al. [2] observed organic EL from single crystal anthracene at ~400 volts. Additional

organic EL research focused on single crystal anthracene where high voltage continued to

be required [3-9]. Vincett et al. [10] was able to reduce the voltage to below 30 V by

exploring anthracene thin films formed by vacuum deposition. While reduced, the voltage

was still much too high for practical applications. In 1987, Tang and VanSlyke [11]

published their seminal work on organic light-emitting diodes (OLED), using thermal

evaporation to deposit an organic bi-layer sandwiched between two electrodes. This

structure achieved 1% external quantum efficiency (EQE) and decreased the voltage to

below 10 V. In 1990, Burrough et al. [12] introduced the first OLED using polymers that

were deposited by spin-coating instead of vapor deposition.

1.2. ORGANIC LIGHT-EMITTING DIODE OPERATING PRINCIPLE

The OLED device structure and materials utilized by Tang and VanSlyke are

shown in Figure 1.1. The distinguishing feature for this device was utilization of a bi-

Page 30: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

2

layer structure to split the charge transport functions between two materials. The bilayer

structure consisted of the di-amine 1,1-bis(di-4-tolylaminophenyl)cyclohexane (TAPC)

and the aluminum complex tris(8-hydroxyquinolinato) aluminium (Alq) sandwiched

between an indium-tin-oxide (ITO) anode and a magnesium:silver cathode. Upon

applying an electric field, electrons are injected from the cathode into the lowest

unoccupied molecular orbital (LUMO) of the electron transport layer (ETL) Alq and

holes are injected from ITO into the highest occupied molecular orbital (HOMO) of the

hole transport layer (HTL) TAPC. The charge carriers recombine at the TAPC/Alq

interface to produce an exciton (bound electron-hole pair) that decays either radiatively to

produce light or non-radiatively to produce heat. In this device, the exciton resides on

Alq, meaning that Alq functions as both the ETL and as the light emitting layer (LEL).

Figure 1.1: OLED configuration and molecular structures used by Tang and VanSlyke.

Since Tang and VanSlyke published their groundbreaking work, there have been intensive

efforts to improve both materials and device structures that have resulted in OLEDs

Glass Substrate

ITO

TAPC

Alq

+

- Mg:Ag

Alq

TAPC

Page 31: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

3

achieving commercial success for display and lighting applications [13]. In order to

achieve such improvements, it was necessary to modify the initial device architecture,

adding layers that have very specific functions. Figure 1.2 shows a state-of-the art OLED

that is made up of several layers. In this structure, the cathode is typically reflective so

light emission is observed through the glass substrate.

Figure 1.2: State-of-the-art OLED device structure.

High work-function transparent conducting oxides (TCO) such as indium-tin-

oxide (ITO) are typically used as the anode [14-20]. Low work-function metals such as

Mg, Ca or Ba or bi-layers such as LiF/Al or Li/Al are typically used for the cathode [21-

27]. In order to reduce resistance, various injection and transport layers are used to move

charge carriers from the electrodes to the emitting layer. Hole injection layers (HIL) and

electron injection layers (EIL) often utilize p-type (HIL) or n-type (EIL) doping, [28-31]

respectively, in an effort to better align the electrode work functions with those of the

organic material HOMO/LUMO levels. After injection into the HIL and EIL, charge

Glass Substrate

Anode

Hole Injection Layer

Hole Transport Layer

Emitting Layer

+

-

Electron Transport Layer

Electron Injection Layer

Cathode

Page 32: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

4

carriers are then injected into and through the HTL and ETL, after which they recombine

within the LEL. The LEL is typically made up of one or two host materials that are

designed to accept holes and/or electrons from the adjacent transport layers, with the goal

being to ensure that excitons form within the LEL.

1.3. EXCITON FORMATION

If the LEL contains only a single material, then the exciton will form on that

material. Tang et al. [32] introduced the concept of doping the LEL with a fluorescent

material to increase efficiency and to tune the color. Fluorescent dopants are typically

uniformly mixed within the LEL at a concentration of 0.5-2%. Higher concentrations

results in aggregation and self-quenching, resulting in a decrease in efficiency. The

dopant is of lower energy than the host so that excitons initially formed in the host

molecules upon excitation will result in the formation of excitons in the dopant molecules

through an energy transfer process such as Fӧrster energy transfer [33, 34]. Fӧrster

energy transfer is a long-range process that involves a dipole-dipole interaction between a

donor (host) emission transition dipole and an acceptor (dopant) absorption dipole. As a

result, it is important for the donor emission spectrum to overlap with the acceptor

absorption spectrum and also for the donor-acceptor pair to be within 10 nm of each

other. Excitons may also form through a short range (10-20 Å) electron transfer process

known as Dexter energy transfer [33, 34]. This process requires the host and dopant

molecules to have overlapping wave-functions. Excitons may also form directly on the

dopant if electron-hole recombinations occur at the dopant preceded by trapping of either

electron or hole at the dopant molecule.

Page 33: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

5

1.4. DEVICE EFFICIENCY

Fluorescent dopants were used in the early stages of OLED development;

however these types of dopants have limitations with respect to EQE. The EQE of an

OLED device is determined by the ratio of the number of photons exiting the device per

the number of electrons injected [35, 36] into the device and is described by equation

(1.1):

(1.1)

Holes and electrons are injected into OLED devices in equal quantities. With the use of

multi-layer OLEDs that can confine charge carriers within the organic stack, it is

generally assumed that every electron injected recombines with its counterpart hole,

resulting in a charge balance factor ( rec) of 100%. The photoluminescence efficiency

(ɳPL) is a property of the emitting material and describes the efficiency of fluorescence

for that material. For the purposes of determining theoretical maximum EQE, this

quantity is generally assumed to be 100%. For OLEDs fabricated on glass substrates and

using reflective cathodes, a large amount of the light is lost within the device and cannot

exit through the glass. This is due predominantly to loss through internal reflections and

wave-guiding, resulting in the optical out-coupling factor (ɳout) to be only ~20% [35].

There is much research devoted to finding ways to improve out-coupling efficiency,

however these techniques will not be discussed here. Spin statistics predicts that

electrically generated excitons produce singlet to triplet excitons in a ratio of 1:3 [37].

Therefore, fluorescent materials (emission from singlet state) have a spin statistic factor

Page 34: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

6

(ɳs) of 25%, assuming that only the singlet excitons can decay radiatively. The triplet

excitons are lost through radiationless decay. Taking into consideration an outcoupling

efficiency of 20%, OLEDs based on fluorescent materials are typically assumed to have

an upper limit of 5% EQE. Baldo et al. [38] first utilized a phosphorescent platinum

complex, 2,3,7,8,12,13,17,18-octaethyl-21H,23H-porphine platinum(II) (PtOEP), as the

emitting dopant in an OLED. Porphin complexes without a heavy metal as the central ion

are known to have long-lived triplet states [39] and low phosphorescence efficiencies.

Metalloporphine complexes with a heavy metal ion such as platinum exhibit spin-orbit

coupling effect, which effectively reduces the phosphorescence lifetime and increases the

efficiency of intersystem crossing from the singlet excited state to the triplet state,

resulting in efficient phosphorescent emission. In a highly efficient OLED device, 100%

of the electrically generated excitons are emissive through both singlet and triplet decays,

leading to an EQE upper limit of 20%. Kondakov et al. [40] recently found that triplet

annihilation processes in fluorescent OLEDs produces radiative singlet excited states,

raising the EQE upper limit to 12.5%. Even with this improvement, phosphorescent

materials continue to have higher limits for efficiency and for this reason, there is

continued intensive research aimed at developing phosphorescent materials.

1.5. IRIDIUM CYCLOMETALATED COMPLEXES

While various transition metals such as platinum [38, 41, 42], iridium [43-46],

ruthenium [47-49] and osmium [50-52] have been utilized to form phosphorescent OLED

dopants, iridium cyclometalated dopants have become the most utilized due to their short

phosphorescent lifetimes (microseconds) and high photoluminescence efficiencies.

Page 35: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

7

Figure 1.3 shows examples of three different Ir dopants. The emission color can be tuned

based on the choice of ligands. Fac-tris(2-phenylpyridine)iridium(III)Ir(ppy)3 [Ir(ppy)3] is

green emitting with an emission maximum (max) of 515 nm. This material has 2-phenyl-

pyridine as the coordinating ligand. The ligand possesses a donor-acceptor pair, with the

phenyl substituent being the donor and the pyridine substituent being the acceptor. A blue

shift can be produced in 3 different ways, 1) adding electron-donating groups to the

pyridine ring, 2) adding electron withdrawing groups to the phenyl ring, or 3) utilize a

Figure 1.3: Blue [Ir(F2ppy)3], green [Ir(ppy)3] and red [Ir(piq)3] Ir-based

phosphorescent dopants and maximum emission peaks (max).

heteroleptic complex that has a strong electron withdrawing ancillary ligand. As shown in

Figure 1.3, fac-tris-4,6-difluorophenylpyridine Ir(III) [Ir(F2ppy)3] has two electron

withdrawing fluorine substituent’s added to the phenyl ring of Ir(ppy)3, resulting in blue

shift of 45 nm from Ir(ppy)3. A red shift can be affected by extending the conjugation of

the ligand’s chromophore. This is shown using fac-tris(1-phenylisoquinolinato) iridium

(III) [Ir(piq)3] which undergoes a 110 nm red-shift from Ir(ppy)3 by changing the pyridine

to isoquinoline. There are numerous types and combinations of ligands that can be

Ir(piq)3

max = 625 nm

Ir(F2ppy)3

max = 470 nm

Ir(ppy)3

max = 515 nm

Page 36: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

8

coordinated to Ir to form highly efficient phosphorescent dopants and this continues to be

an area of intense research [53-58].

Similar to fluorescent dopants, phosphorescent dopants require a host material to

function in an OLED device. In addition to being used for charge transport and exciton

formation, the host materials must have a triplet energy (ET) that is higher than that of the

phosphorescent dopant. If the ET of the host is lower than that of the dopant, then the host

triplets (assuming non-phosphorescent) will act as quenchers for the phosphorescent

dopant triplets, resulting in an overall loss in electroluminescence efficiency.

1.6. DEVICE LIFETIME

There are typically two ways to test device lifetime: 1) Choose a starting

luminance (brightness) and then provide the current density needed to achieve that

brightness. Luminance is the luminous intensity of a surface in a given direction per unit

of projected area, measured in units of candela per square meter, (cd/m2). Current density

is the current per unit area, such as milliamps per square centimeter (mA/cm2). A

commonly used lifetime metric, t50, is the time it takes for the initial luminance to drop by

50% while keeping the current density constant. 2) Choose a specific current density

which will give an initial luminance (or EQE). While keeping the current density fixed,

the luminance will decrease with time and the time taken for the initial luminance to

decrease by 50% is the lifetime (t50).

Universal Display Corporation (UDC) is one of the leading phosphorescent

materials development companies in the world. Without disclosing specific materials or

device architectures, UDC has disclosed t50’s based on a starting luminance of 1,000

Page 37: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

9

cd/m2 for red, green and blue phosphorescent OLED’s (PHOLED) of 900,000 hours,

400,000 hours and 20,000 hours, respectively [59]. The blue PHOLED has an order of

Figure 1.4: Commission Internationale De L'Eclairage (CIE) 1931 color coordinates.

magnitude lower t50 which, contrary to red and green PHOLEDs, prevents the blue

PHOLED from being utilized in commercial applications.

Ideally, the red, green and blue (RGB) emitters should be able to produce

saturated red, green and blue colors with a sufficiently wide color gamut for display

applications. For lighting applications, these emitters should be able to produce white

light of various color temperatures. The Commission Internationale De L'Eclairage (CIE)

1931 color space [60] is shown in Figure 1.4 and provides a measure of color saturation

using an x,y coordinate system. The most saturated colors are those at the edge of the CIE

diagram. For the lifetimes mentioned above, the CIEx,y coordinates are 0.64, 0.36 (red),

Page 38: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

10

0.31, 0.63 (green) and 0.18, 0.42 (blue). While the red and green coordinates of the

phosphorescent materials are fine for commercial applications, the blue coordinates are

far from where they need to be, with both x and y coordinates needing to be 0.15 or

lower. Considering lifetime and color, the blue PHOLED performance lags far behind

that for red and green.

1.7. DEGRADATION

OLED lifetime is typically affected by two types of degradation, extrinsic and

intrinsic. Extrinsic degradation occurs without the device operating, such as the

electrodes and/or organic materials reacting with water and oxygen present in ambient,

resulting in formation of dark spots and/or luminance quenching species. Extrinsic

degradation has largely been eliminated by using well established encapsulation

techniques. Intrinsic degradation is continuous and results from device operation,

occurring across the device area and inversely dependent on the device current density or

luminance.

Intrinsic degradation is typically attributed to chemical or electrochemical

degradation of specific organic materials in the OLED stack, with the degradation

mechanisms being linked to OLED operational functions such as charge transport and

exciton formation. The identification of the degraded materials has been very difficult

due to the complexity of the OLED layer structure and various interfaces, and the

uncertainty of the material components that are primarily responsible for the device

degradation. As OLED degrades, it is certain that quenchers are formed that can affect

directly the emission efficiency. These quenchers can also affect the transport of charges

Page 39: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

11

by acting as traps for either electrons or holes or both, and thus can cause a shift of the

recombination zone as well as the balance of electron and hole densities, which indirectly

can result in a loss of emission efficiency. In most cases, the degraded OLED device will

also require a higher drive voltage to maintain a constant current density due to the

creation of charge traps.

It has been reported that hole and electron transport through Alq produces

unstable Alq radical cations [61] and anions [62], respectively, that form quenching

species in subsequent chemical reactions. It has been reported that electron transport

through 4,4’-bis(3-methylcarbazol-9-yl)-2,2’-biphenyl (BMB) results in luminance loss

and increasing drive voltage [63]. Kondakov et al. [64-67] have shown that charge

recombination at the HTL/LEL interface, where the HTL contains either an aryl-amine or

carbazole moiety, results in the formation of the HTL excited state and subsequent

homolytic C–N bond dissociation. Scholz et al. [68] have shown that the phosphorescent

host 4,4’,4’’-tris(carbazol-9-yl) triphenylamine (TCTA) undergoes dissociation in a

similar manner. Charge recombination at a the LEL/ETL interface with 4,7-diphenyl-

1,10-phenanthroline (BPhen) as the ETL has been shown to result in the formation of

BPhen dimers [68, 69], a process which introduces gap states that affect electron injection

into the LEL. Exciton-polaron interactions can lead to the formation of defects that act as

luminescent quenchers or non-radiative recombination centers [63, 70-74]. These

interactions can also cause molecular aggregation and luminance quenching [75].

Degradation products for iridium cyclometalated complexes have also been

identified. The green emitting dopant Ir(ppy)3 undergoes reversible dissociation of one of

Page 40: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

12

the phenyl-pyridine ligands [76], during which time it forms a radical cation that reacts

with the electron-transport material. Dopants such as bis(2-(4,6-difluorophenyl)pyridyl-

N,C2’)iridium(III) picolinate (FIrpic) and bis(2-methyldibenzo-[f,h]quinoxaline)

(acetylacetonate) (Ir(MDQ)2(acac)) also undergo dissociation of their ancillary oxygen

containing ligands, picolinate [77] and acetylacetonate [78, 79], respectively. It has also

been shown that one of the fluorine substituents on the phenyl ring within FIrpic also

dissociates from the molecule [80].

1.8. OBJECTIVES AND OUTLINE OF THESIS

In our power consuming digital world, the need to develop less power intensive

electronic devices is widely recognized. OLED devices are currently incorporated into

commercial products for lighting and displays. As these products compete with

alternative existing technologies such as compact fluorescent bulbs (lighting) and liquid

crystal displays (LCD), there is great motivation to continue on a path of research that

makes these devices more efficient than their counterparts.

Phosphorescent materials are used to make the most efficient OLED devices. Red

and green phosphorescent materials are currently utilized in commercial products. Highly

efficient blue phosphorescent OLEDs have been demonstrated, however short device

lifetime prohibits incorporation of these materials into commercial applications.

Therefore, it is imperative that the degradation mechanisms and processes are understood

so that new blue phosphorescent materials and device structures can be developed to

improve device lifetime to a point where OLED device products are comprised of R,G,B

phosphorescent emitters. The objective of this thesis is to investigate blue phosphorescent

Page 41: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

13

OLED instability from both a material and device perspective.

In Chapter 3, a series of OLED devices are fabricated with various host and

transport materials. In all devices, the well-known blue phosphorescent dopant FIrpic is

used as the light-emitter. Depending on the device structure, device lifetimes ranging

from minutes to 12 hours have been achieved. It is found that the formation of the FIrpic

excited state is extremely detrimental to device lifetime. It is also shown that

recombination and emission from the host or electron transport materials is damaging to

device lifetime. Fabrication of stable devices with lifetimes of thousands of hours is not

possible when FIrpic is used as the light emitting material.

In Chapter 4, we investigate how hole transport through FIrpic affects device

performance of model devices that have Alq as the light emitting material. FIrpic is

doped into hole transport materials at various concentrations. Hole transport through

FIrpic forms the FIrpic radical cation, an unstable species that degrades to form hole

trapping products. This results in large increases in the drive voltage. As the voltage rises,

the EQE decreases due to reduced electron-hole recombination occurring within the Alq

emitter. Additionally, the blue phosphorescent host mCBP is found to be stable to hole

transport. However this material contributes to device degradation when participating in

electron-hole recombination.

In Chapter 5, we utilize the blue phosphorescent dopant Ir(iprpmi)3. This dopant

has a very different molecular structure compared to FIrpic. It lacks fluorine substituents

and the picolinate ligand that are present in FIrpic. The photophysical properties of

Ir(iprpmi)3 are reported. A series of devices are fabricated where a range of Ir(iprpmi)3

Page 42: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

14

concentrations are studied while using various hole and electron transport materials. It is

found that EQE’s over 20% can be achieved using Ir(iprpmi)3 as the light-emitting

material. A high EQE is also demonstrated for devices using the low triplet energy BAlq

as the ETL. Device lifetime is significantly improved compared to devices using FIrpic as

the light-emitting material.

Page 43: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

15

REFERENCES

[1] A. Bernanose, M. Comte, P. Vouaux, J. Chim. Phys. Phys.-Chim. Biol., 50 (1953)

64.

[2] M. Pope, H.P. Kallmann, P. Magnante, J. Chem. Phys., 38 (1963) 2042.

[3] J. Dresner, RCA Rev., 30 (1969) 322.

[4] W. Helfrich, W. Schneider, Phys. Rev. Lett., 14 (1965) 229.

[5] M. Schadt, D.F. Williams, Phys. Status Solidi, 39 (1970) 223.

[6] D.F. Williams, M. Schadt, J. Chem. Phys., 53 (1970) 3480.

[7] D.F. Williams, M. Schadt, Proc. IEEE, 58 (1970) 476.

[8] D.F. Williams, M. Schadt, Proc. Int. Conf. Photocond., 3rd, (1971) 303.

[9] D.F. Williams, M. Schadt, U.S. Patent No. 3,624,632 (1971).

[10] P.S. Vincett, W.A. Barlow, R.A. Hann, G.G. Roberts, Thin Solid Films, 94 (1982)

171.

[11] C.W. Tang, S.A. VanSlyke, Appl. Phys. Lett., 51 (1987) 913.

[12] J.H. Burroughes, D.D.C. Bradley, A.R. Brown, R.N. Marks, K. Mackay, R.H.

Friend, P.L. Burns, A.B. Holmes, Nature (London), 347 (1990) 539.

[13] Mertens, R., www.oled-info.com/devices, 2014

[14] J.-H. Bae, J.-M. Moon, S.W. Jeong, J.-J. Kim, J.-W. Kang, D.-G. Kim, J.-K. Kim,

J.-W. Park, H.-K. Kim, J. Electrochem. Soc., 155 (2007) J1.

[15] J. Cui, A. Wang, N.L. Edleman, J. Ni, P. Lee, N.R. Armstrong, T.J. Marks, Adv.

Mater. (Weinheim, Ger.), 13 (2001) 1476.

[16] H. Kim, C.M. Gilmore, J.S. Horwitz, A. Pique, H. Murata, G.P. Kushto, R. Schlaf,

Page 44: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

16

Z.H. Kafafi, D.B. Chrisey, Appl. Phys. Lett., 76 (2000) 259.

[17] H. Kim, J.S. Horwitz, G.P. Kushto, S.B. Qadri, Z.H. Kafafi, D.B. Chrisey, Appl.

Phys. Lett., 78 (2001) 1050.

[18] T.W. Kim, D.C. Choo, Y.S. No, W.K. Choi, E.H. Choi, Appl. Surf. Sci., 253

(2006) 1917.

[19] T.J. Marks, J.G.C. Veinot, J. Cui, H. Yan, A. Wang, N.L. Edleman, J. Ni, Q.

Huang, P. Lee, N.R. Armstrong, Synth. Met., 127 (2002) 29.

[20] K.L. Purvis, G. Lu, J. Schwartz, S.L. Bernasek, J. Am. Chem. Soc., 122 (2000)

1808.

[21] H. Becker, A. Lux, A.B. Holmes, R.H. Friend, Synth. Met., 85 (1997) 1289.

[22] P. Broems, J. Birgersson, N. Johansson, M. Loegdlund, W.R. Salaneck, Synth.

Met., 74 (1995) 179.

[23] B. Chen, J. Hou, S. Xue, S. Liu, Semicond. Photonics Technol., 2 (1996) 293.

[24] P. Dannetun, M. Fahlman, C. Fauquet, K. Kaerijama, Y. Sonoda, R. Lazzaroni,

J.L. Bredas, W.R. Salaneck, Synth. Met., 67 (1994) 133.

[25] L. Duan, K. Xie, Y. Qiu, J. Soc. Inf. Disp., 19 (2011) 453.

[26] M. Stossel, J. Staudigel, F. Steuber, J. Blassing, J. Simmerer, A. Winnacker, H.

Neuner, D. Metzdorf, H.H. Johannes, W. Kowalsky, Synth. Met., 111-112 (2000)

19.

[27] M. Stossel, J. Staudigel, F. Steuber, J. Simmerer, A. Winnacker, Appl. Phys. A, 68

(1999) 387.

[28] R. Meerheim, B. Luessem, K. Leo, Proc. IEEE, 97 (2009) 1606.

Page 45: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

17

[29] B. Luessem, M. Riede, K. Leo, Phys. Status Solidi A, 210 (2013) 9.

[30] M. Pfeiffer, K. Leo, X. Zhou, J.S. Huang, M. Hofmann, A. Werner, J. Blochwitz-

Nimoth, Org. Electron., 4 (2003) 89.

[31] K. Walzer, B. Maennig, M. Pfeiffer, K. Leo, Chem. Rev. (Washington, DC, U. S.),

107 (2007) 1233.

[32] C.W. Tang, S.A. VanSlyke, C.H. Chen, J. Appl. Phys., 65 (1989) 3610.

[33] A.A. Shoustikov, Y. You, M.E. Thompson, IEEE J. Sel. Top. Quantum Electron., 4

(1998) 3.

[34] V. Bulovic, M.A. Baldo, S.R. Forrest, Springer Ser. Mater. Sci., 41 (2001) 391.

[35] C. Adachi, M.A. Baldo, M.E. Thompson, S.R. Forrest, J. Appl. Phys., 90 (2001)

5048.

[36] H. Yersin, A.F. Rausch, R. Czerwieniec, T. Hofbeck, T. Fischer, Coord. Chem.

Rev., 255 (2011) 2622.

[37] S.P. McGlynn, T. Azumi, M. Kinoshita, Molecular Spectroscopy of the Triplet

State (Prentice-Hall International Series in Chemistry), 1969.

[38] M.A. Baldo, D.F. O'Brien, Y. You, A. Shoustikov, S. Sibley, M.E. Thompson, S.R.

Forrest, Nature (London), 395 (1998) 151.

[39] M.S. Weaver, Opt. Sci. Eng., 111 (2007) 527.

[40] D.Y. Kondakov, T.D. Pawlik, T.K. Hatwar, J.P. Spindler, J. Appl. Phys., 106

(2009) 124510.

[41] J. Kalinowski, V. Fattori, M. Cocchi, J.A.G. Williams, Coord. Chem. Rev., 255

(2011) 2401.

Page 46: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

18

[42] H.-F. Xiang, S.-W. Lai, P.T. Lai, C.-M. Che, Highly Effic. OLEDs Phosphoresc.

Mater., (2008) 259.

[43] C. Adachi, M.A. Baldo, S.R. Forrest, S. Lamansky, M.E. Thompson, R.C.

Kwong, Appl. Phys. Lett., 78 (2001) 1622.

[44] M.A. Baldo, S. Lamansky, P.E. Burrows, M.E. Thompson, S.R. Forrest, Appl.

Phys. Lett., 75 (1999) 4.

[45] S. Lamansky, P.I. Djurovich, D. Murphy, F. Abdel-Razzaq, H.-E. Lee, C. Adachi,

P.E. Burrows, S.R. Forrest, M.E. Thompson, J. Am. Chem. Soc., 123 (2001) 4304.

[46] T. Tsutsui, M.-J. Yang, M. Yahiro, K. Nakamura, T. Watanabe, T. Tsuji, Y. Fukuda,

T. Wakimoto, S. Miyaguchi, Jpn. J. Appl. Phys., 38 (1999) L1502.

[47] W.K. Chan, P.K. Ng, X. Gong, S. Hou, Appl. Phys. Lett., 75 (1999) 3920.

[48] F.G. Gao, A.J. Bard, J. Am. Chem. Soc., 122 (2000) 7426.

[49] J. Yang, K.C. Gordon, Synth. Met., 152 (2005) 213.

[50] B. Carlson, X. Jiang, A.K.Y. Jen, L.R. Dalton, Proc. SPIE-Int. Soc. Opt. Eng.,

4809 (2002) 69.

[51] J. Lu, Y. Tao, Y. Chi, Y. Tung, Synth. Met., 155 (2005) 56.

[52] Y.-L. Tung, P.-C. Wu, C.-S. Liu, Y. Chi, J.-K. Yu, Y.-H. Hu, P.-T. Chou, S.-M.

Peng, G.-H. Lee, Y. Tao, A.J. Carty, C.-F. Shu, F.-I. Wu, Organometallics, 23

(2004) 3745.

[53] Y. Chi, P.-T. Chou, Chem. Soc. Rev., 39 (2010) 638.

[54] P.-T. Chou, Y. Chi, Chem. - Eur. J., 13 (2007) 380.

[55] B. Minaev, G. Baryshnikov, H. Agren, Phys. Chem. Chem. Phys., 16 (2014) 1719.

Page 47: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

19

[56] P.I. Djurovich, M.E. Thompson, Highly Effic. OLEDs Phosphoresc. Mater.,

(2008) 131.

[57] L.F. Gildea, J.A.G. Williams, Woodhead Publ. Ser. Electron. Opt. Mater., 36

(2013) 77.

[58] Y. Zhao, D. Ma, Trends Inorg. Chem., 13 (2012) 121.

[59] Universal Display Corporation, www.udcoled.com, 2014

[60] T. Smith, J. Guild, Trans. Optical Soc. (London), 33 (1932) 73.

[61] H. Aziz, Z.D. Popovic, N.-X. Hu, A.-M. Hor, G. Xu, Science (Washington, D. C.),

283 (1999) 1900.

[62] Y. Luo, H. Aziz, G. Xu, Z.D. Popovic, Chem. Mater., 19 (2007) 2079.

[63] N.C. Giebink, B.W. D'Andrade, M.S. Weaver, J.J. Brown, S.R. Forrest, J. Appl.

Phys., 105 (2009) 124514.

[64] D.Y. Kondakov, J. Appl. Phys., 104 (2008) 084520.

[65] D.Y. Kondakov, Proc. SPIE, 7051 (2008) 70510I.

[66] D.Y. Kondakov, T.D. Pawlik, W.F. Nichols, W.C. Lenhart, J. Soc. Inf. Disp., 16

(2008) 37.

[67] D.Y. Kondakov, organic electronics book, (2010) 211.

[68] S. Scholz, K. Walzer, K. Leo, Adv. Funct. Mater., 18 (2008) 2541.

[69] S. Scholz, C. Corten, K. Walzer, D. Kuckling, K. Leo, Org. Electron., 8 (2007)

709.

[70] S. Reineke, K. Walzer, K. Leo, Physical Review B, 75 (2007).

[71] N.C. Giebink, B.W. D'Andrade, M.S. Weaver, P.B. Mackenzie, J.J. Brown, M.E.

Page 48: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

20

Thompson, S.R. Forrest, J. Appl. Phys., 103 (2008) 044509.

[72] H. Zamani Siboni, H. Aziz, Appl. Phys. Lett., 101 (2012) 063502.

[73] Q. Wang, H. Aziz, ACS Appl. Mater. Interfaces, 5 (2013) 8733.

[74] Q. Wang, I.W.H. Oswald, M.R. Perez, H. Jia, B.E. Gnade, M.A. Omary, Adv.

Funct. Mater., 23 (2013) 5420.

[75] Q. Wang, B. Sun, H. Aziz, Adv. Funct. Mater., 24 (2014) 2975.

[76] I. Rabelo de Moraes, S. Scholz, B. Luessem, K. Leo, Appl. Phys. Lett., 99 (2011)

053302.

[77] I. Rabelo de Moraes, S. Scholz, B. Luessem, K. Leo, Org. Electron., 12 (2011)

341.

[78] R. Meerheim, S. Scholz, S. Olthof, G. Schwartz, S. Reineke, K. Walzer, K. Leo, J.

Appl. Phys., 104 (2008) 014510.

[79] S. Scholz, R. Meerheim, B. Luessem, K. Leo, Appl. Phys. Lett., 94 (2009)

043314.

[80] V. Sivasubramaniam, F. Brodkorb, S. Hanning, H.P. Loebl, V. van Elsbergen, H.

Boerner, U. Scherf, M. Kreyenschmidt, J. Fluorine Chem., 130 (2009) 640.

Page 49: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

21

CHAPTER 2

DEVICE FABRICATION EXPERIMENTAL DETAILS

2.1 THERMAL EVAPORATION COATER

All OLED devices described in this thesis were fabricated using materials

(organic, inorganic, and metal) that were deposited (coated) as thin films using the well

established thermal evaporation technique [1-5]. Figure 2.1 shows the basic design of a

vacuum coater for the deposition of thin films by thermal evaporation. The required

vacuum chamber pressure is typically 10-6

torr or lower.

A thermal evaporation source, often known as a boat, is used to vaporize solid

materials for thin film deposition. A typical boat consists of a resistive heating element,

which is typically made of refractive metals such as tungsten, tantalum or molybdenum in

the form of a sheet or coil. Often the metal sheet is shaped in the form of a boat to hold

the material (known as the source material) to be evaporated. A quartz or ceramic

crucible is also often used to hold the source material with the metal sheet or coil acting

the heating element. Low-voltage electrical power is generally the heating source. Upon

heating in the vacuum chamber to a sufficiently high temperature, the source solid

material will vaporize and transform into isolated molecules, which are ejected from the

boat into the vacuum space of the deposition chamber. Condensation of these molecules

on a substrate placed in a line-of-sight position from the boat source will form film. For

Page 50: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

22

the fabrication of OLED devices, the substrate is usually kept at or near room

temperature to promote the formation of uniform and generally amorphous thin films.

The deposition of thin films by thermal evaporation is usually controlled by a

commercially available thin-film deposition controller (DC), which is capable of

measuring the rate of thin film deposition (in Angstrom per sec) using a quartz crystal

microbalance (QCM). The QCM is placed in a line-of-sight position with respect to the

boat source. A feedback loop to the boat power supply is used to control the temperature

of the source material in the boat and thus its rate of evaporation. The DC can monitor

both the film thickness accumulated on the substrate and control the rate of accumulation

by adjusting the electric power supplied to the boat. Independently, the thickness of the

film deposited on the substrate can be measured using a profilometer or ellipsometer.

Figure 2.1: Basic design of a thermal evaporation coater.

Vacuum Pump

Deposition

Controller

Power Supply

Chamber

Substrate

Shutter

Quartz crystal microbalance

Resistively

heated boat

Page 51: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

23

A shutter is used to prevent any material from coating the substrate until a well-

controlled deposition rate is established. This enables accurate control of the thickness of

the thin film deposited on the substrate.

2.2 THE KOMET COATER

A custom-designed vacuum coater was designed and built by K. Klubek for the

fabrication of all OLED devices described in this thesis. This coater is known as the

KOMET coater, an acronym based on the names of those who contributed to this system

being built: Klubek (Kevin), Madathil (Joseph), and Tang (Ching). The last two letters of

the acronym are derived from our field of study, Organic Electronics. This coater was

designed to include several important features: 1) compact thermal source design, 2) low

power consumption boats for depositing organic materials, and 3) co-evaporation of

multiple materials.

Figure 2.2 shows a picture of the KOMET coater. This table-top system has a

14”x14” square base and a top plate that also measures 14”x14”. The chamber, placed

between these two plates, consists of a Pyrex collar (12.5” outside diameter) that sits on

top of a metal collar (13-7/8” outside diameter). The bottom plate is attached to 4 posts

(8-7/8” long) that sit on the bench-top. The posts raise the chamber and allow easy access

to various feedthroughs in the bottom plate. Between the top and bottom plates is the

working space of the coater. Mounted on these plates are all the key feedthroughs for

various electrical and mechanical components (electrical power, QCM signals,

turbopump, shutter controls, etc).

Page 52: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

24

Figure 2.2: KOMET coater.

The inside of the KOMET coater chamber is shown in Figure 2.3. Figure 2.3(a)

shows the top of the chamber which includes the substrate and mask platters, and the

shutter plate. The substrate platter can accommodate 6 square substrates (1.5”x1.5”). The

mask platter holds two different masks that are used for fabricating OLED devices, one

for depositing organic/inorganic materials and another for depositing metals. The shutter

plate is placed below the mask and substrate platters.

Figure 2.3(b) shows the bottom of the coater which is partitioned into two

separate sides, one for depositing organic materials and another for depositing

metal/inorganic materials. Partitioning the two sides is to prevent cross-contamination

from occurring. Figure 2.3(c) shows the side with the boat assembly for organic material

deposition. The key feature of this coater is the boat assembly (4” diameter) consisting of

Top plate

Base plate

Chamber

12.5”

6-1/16”

14”

14”

8-7/8”

Page 53: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

25

Figure 2.3: Inside the KOMET coater chamber. (a) Top of the chamber, (b) bottom of the chamber,

(c) area for depositing organic materials, and (d) area for depositing metals/inorganics.

12 custom-built, low-temperature, filament evaporation boats (organic boats) for

thermally depositing organic materials. The organic boats are based on a glass tube

design originally developed by S. Lee [6] and will be discussed in the following section,

section 2.3. The electrical sockets for the boats are mounted to a rotatable platter. There

are 4 water-cooled QCM’s (MCVAC Manufacturing Company part #’s 700-001, 100-011,

900-004, 900-002, 900-001) fixed in positions with respect to the rotatable platter. This

arrangement allows the deposition of any one material or the co-deposition of two to four

materials, all out of the 12 boat sources for any given thin-film deposition. The QCM’s

12-16-11-DOE site visit

Substrate platterMask platter

Shutter plate

4 Crystal sensors

Organic boats

Common electrode

Individual electrodes

Glass divider

(a) (b)

(c) (d)

Organic deposition

Metal/Inorganicdeposition

Organic source assembly

Metal boat

Page 54: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

26

are water cooled to minimize effects of radiant heat from the boats. The rotatable platter

design allows the selection of materials held in designated boats for co-evaporation.

Figure 2.3(d) shows the metal/inorganic side of the coater. The boats on this side

are made from tungsten (R.D. Mathis Company part # ME4-.005W) for the deposition of

aluminum (Al), and from tantalum (R.D. Mathis company part # SB3-AO-TA) lithium

fluoride (LiF) and molybdenum oxide (MoOx) deposition. The current required to

evaporate these materials are: 120-150 amps for Al (at a rate of 10 Å/s), 90-100 amps for

LiF and 55-65 amps for MoOx (at rates of 0.5 Å/s for each). There is a water cooled

QCM positioned above these boats.

Figure 2.4: KOMET coater top plate.

Figure 2.4 shows the top plate. This plate has several feedthroughs, including the

Top plate

Substrate port

Metal shutter

Organic shutter Substrate/

maskrotor

Page 55: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

27

substrate/mask rotor which is used to rotate the substrates into position for depositing

either organic or inorganic/metal materials. This plate also has the control for the organic

and metal shutters. Substrates are loaded into the substrate platter through the substrate

port.

Figure 2.5 shows a turbopump (Pfeiffer TPU 170) connected to the bottom plate.

This pump allows the KOMET coater to achieve a base pressure of 2 × 10-7

Torr.

Figure 2.5: Features below the KOMET bottom plate.

Also attached to the bottom plate is the rotor for the organic material evaporation source

assembly and the apparatus for indexing the organic boats. The indexed position relates

the boat positions to the QCM’s and determines which boats can be used to evaporate

materials.

Figure 2.6 shows the control cabinet for the KOMET coater. This cabinet contains

Organic boat rotor

Apparatus for indexing organic boats

Turbopump

Bottom plate

Page 56: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

28

Figure 2.6: Control cabinet.

a) the turbopump controller (Pfeiffer TCP 310), b) two deposition controllers, one for

organics (Kurt J. Lesker-FTC-2800) and one for metals/inorganics (Sigma Instruments

SQC-122c), c) organic switchboard connecting the power supplies to 4 out of the 12

organic boats, d) metal/inorganic switchboard for connecting a power supply to one of

the 4 available boats, and e) the power supplies. The deposition controller for the

organics is capable of monitoring the deposition of 1-4 materials at a time. There are 4

power supplies (Sorensen DCS8-125E) used for depositing the organic materials and

there is a single power supply (Sorensen DCS12-250) used for depositing inorganic/metal

materials. These particular power supplies were used out of convenience (they were

donated by Eastman Kodak Company) and not out of necessity as the power

Power supplies

Metal/inorganic

switchboard

Organic switchboard

Deposition controllers

Turbo-pump

controller

Page 57: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

29

requirements for the boats are all far lower than what these power supplies can provide.

2.3 ORGANIC BOATS

The components and design of the organic boat are shown in Figure 2.7. The steps

for assembling the boat are shown in Figure 2.8. A disposable glass test tube (VWR

catalog # 47729-570) is cut at both ends using a diamond wheel (SE DW13,

B000P49NCC) to form a cylinder and a side hole is made using a diamond drill bit

(McMaster-Carr part # 4376A12, 0.039” diameter). A nickel:chromium (nichrome) (22

BNC, 0.0253” nominal diameter) coil is used as the resistive heating element. This coil

was made using a 0.325” hollowed-out aluminum spindle attached to a lathe. There are

16 coils, with a space made between coils 6 and 7 (counting from top) to prevent the coil

from blocking the side hole. A ceramic rod made from Macor is machined to make the

base of the boat. The nichrome coil is inserted into the base and the glass cylinder then

surrounds the coil. Contact pins (Summit part# 030 1952 000) are attached to the

nichrome wire with a press. Aluminum foil is wrapped around the outside of the glass

tube, leaving an opening for the side hole. This foil is important because it helps to keep

heat within the tube. Without the foil, material condenses between the coils, resulting in

inconsistent film thicknesses. A high temperature epoxy (Cotronics Corporation, Resbond

907) is used to bind the contact pins to the macor base. This epoxy is also used to bind

the glass cylinder to the Macor base. A heat gun is used during application of the epoxy to

assist with drying. The entire assembly is then heated at 300º C in an oven (in air) to fully

cure the epoxy. There is a ~0.3” gap between the macor base and the bottom of the

nichrome coil. This space holds 100-300 mg of organic material.

Page 58: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

30

Figure 2.7: Picture of components used to build organic boat.

Figure 2.8: Components for assembling organic boat.

Test tube

Cylinder

Nichrome coil

Macor

base

Electrical

contact

pins

Assembled boats

Epoxy

Al

shield

0.157”

0.118”

0.407”0.512”

Macor base

Nichrome coilGlass cylinder with

0.039” side hole

Electrical

contact pin

0.41”

1.71”

Space for

material

0.47”

Al shield

0.3”

Page 59: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

31

As shown in Figure 2.3(c), the boats with the NiCr wires are fitted into the

ceramic electrical sockets with the side hole directly facing the QCM. As material is

heated within the boat, the vapor exits from the top opening and their relative deposition

rates are monitored by the top and side QCMs. The linear correlation between these rates

is used to establish the side QCM as the rate controller. The calibration of the side QCM

is done by actually measuring the thickness of the deposited film on the substrate position

using a KLA Tencor D100 profilometer. A tooling factor is established for the side QCM

so that its thickness reading can be equated to the measured thickness.

The organic boats require less than 10 W to deposit commonly used OLED

materials. The voltage and current requirements are ~2–4 V and ~1–2 A, respectively.

2.4 DEVICE FABRICATION

All devices were prepared on glass substrates (Tinwell Electronic Technology

Company) pre-patterned with indium-tin-oxide (ITO) having a thickness of 110 nm and a

sheet resistance of about 15 ohm/square. Figure 2.9 shows the ITO patterns. There are 4

test patterns on each substrate, all with an identical structure with etched ITO as

electrodes. The substrates were batch-wise cleaned by, sequentially, soaking and

mechanically scrubbing in detergent solution, washing in deionized water, acetone and

isopropanol using an ultrasonic bath, drying with nitrogen and then further treating with

O2 plasma. All films were prepared by thermal deposition (<5x10-6

Torr) without

breaking vacuum until the OLED device was completed with the top electrode. The

deposition rate for host and transport layers was 4 Å/s. Dopant rates were a percentage of

the host rate and therefore were varied based on the choice of concentration.

Page 60: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

32

Figure 2.9: ITO pattern on substrate.

Molybdenum trioxide (MoOx) of 1.2 nm was used as a hole-injection layer (HIL)

[7-11]. It was deposited on top of the ITO using an “organic” mask with a circular

opening that permits deposition over the 4 test patterns. The term “Anode” within this

thesis refers to the sequential thin film sequence consisting of the following bilayer

structure: Glass substrate/110 nm ITO/1.2 nm MoOx. The term “Cathode” within this

thesis refers to the sequential thin film sequence consisting of the following bilayer

structure: 1.0 nm LiF/100 nm Al [12-14]. A typical representation for devices that will be

discussed within this thesis is the following: Anode/organic layers/Cathode, where the

organic layers consist of a sequence of hole-transport, light-emitting, and electron-

Page 61: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

33

transport layers.

Figure 2.10(a) shows how the device active area is defined. On top of the ITO

pattern, through the organic mask (0.65” diameter) is deposited in order: MoOx, the

organic layers and then LiF. For the cathode, aluminum is then deposited through an L-

shape “metal” mask. The area where the ITO pattern and Al metal overlap (with organic

material in between) is the active area, which is 0.1 cm2 in our design. Figure 2.10(b)

shows a completed device with all 4 pixels being lit. After fabrication, the OLED devices

incurred a brief exposure to the ambient atmosphere before transferring to a vacuum

assembly. The vacuum assembly shown in Figure 2.10(b) and 2.10(c) consists of a valve

attached to a flange containing an o-ring. The device is placed on the o-ring, a mechanical

pump is attached to the end of the valve, and then the assembly was evacuated to below

50 mTorr using a mechanical pump. This valve assembly provides a a simple

encapsulated environment for the OLED device and is suitable for short-term lifetime

tests.

Figure 2.10: Pictures of OLED active area and encapsulation technique (a) ITO pattern, organic

deposited through circular mask, and metal deposited through L-shaped mask, (b) An OLED device

with 4 lit pixels, and (c) Vacuum assembly for encapsulating devices.

(a) (b) (c)

ValveConnection to vacuum pump

SubstrateO-ringFlange

Flange with O-ring0.65” diameter

Page 62: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

34

2.5 OLED PERFORMANCE MEASUREMENTS

The luminance-current-voltage (LIV) characteristics were measured using a

Keithley 2400 source meter and a Photoresearch PR650 SpectraScan Colorimeter. The

source meter and PR650 were computer controlled using a software program developed

by John Burtis from Eastman Kodak Company. To obtain an LIV trace, the source meter

supplies a constant current to the OLED device under test and the corresponding voltage

from the source meter and the luminance value from the PR650 were logged. These

measurements were repeated for a range of currents from 0.1 to 100 mA/cm2. The output

data from the software program is shown in Figure 2.11. They include both radiometric

and photometric data of the electroluminescence as a function of current density and

voltage. The spectral data, as well as the luminous efficiency data in terms of cd/A

(candela per ampere) were also plotted.

The devices were viewed perpendicular to the device plane, and the angular

distribution of the EL was assumed to be Lambertian.

2.6 DEVICE LIFETIME

Device lifetimes for phosphorescent OLEDs in this thesis were measured at a

current density of 5 mA/cm2. EL data cited in the literature are often reported at 5

mA/cm2. This makes it simple to compare data collected within this thesis to those

reported in the literature. Furthermore, this is a practical current density that typically

provides a brightness (500-2000 cd/m2) that is comparable to what would be required for

commercial device applications. For a typical OLED device lifetime measurement, the

initial brightness (cd/m2) was first measured using PR650 at 5 mA/cm

2. The device was

Page 63: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

35

Figure 2.11: Output from John Burtis software program (a) current density (mA/cm2), (b) drive voltage (V), (c) luminance (cd/m2), (d) CIEx

coordinate, (e) CIEy coordinate, and (f) EQE or p/e (photon/electron).

(a) (b) (c) (d) (e) (f)

Page 64: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

36

then moved to a lifetime test station where it was driven at 5 mA/cm2 and the brightness

was continuously monitored by a silicon photodiode (Hamamatsu S1787-08) placed in

front of the OLED device. The lifetime of the OLED is the time it takes for the brightness

to decrease by 50% (t50). The luminance output was again recorded by the PR650 at the

end of the lifetime.

Page 65: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

37

2.7 MATERIALS

The acronym, chemical name, molecular structure, and relevant energy levels for

all organic materials described in this thesis are shown in Table 2.1. The following

materials were purchased from Nichem Fine Technology Company and used as received:

FIrpic (Ni-D306), TCTA (Ni-HB03), and TmPyPB (Ni-E206). The following materials

were purchased from Luminescence Technology Corporation and used as received:

Ir(iprpmi)3 (LT-N643), and UGH3 (LT-N449). The following materials were obtained

from Eastman Kodak Company and used as received: Alq, BAlq, BPhen, NPB, TAPC.

mCBP (ALD-F007) was purchased from Jilin OLED Material Technology Company and

used as received. Aluminum (stock# 43357), LiF (stock# 14056) and molybdenum(VI)

oxide (stock# 11837) were all purchased from Alfa-Aesar and used as is.

Molybdenum(VI) oxide is referred to as MoOx throughout this thesis because the oxygen

oxidation state is variable [15] depending on the deposition conditions.

With the exception of Ir(iprpmi)3, the values of the highest occupied molecular

orbital (HOMO), the lowest unoccupied molecular orbital (LUMO), and the triplet energy

are from the literature. Unless noted otherwise, HOMO energies were determined using

ultraviolet photoelectron spectroscopy and LUMO energies were based on the difference

of the HOMO level and the optical band gap as determined by the onset of absorption.

Triplet energies are estimated experimentally from phosphorescence spectra at 77 K.

Page 66: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

38

Table 2.1: Organic materials discussed throughout this thesis. HOMO/LUMO/triplet energies were taken from literature. HOMO energies are based

on ionization potential, LUMO energies based on difference between optical bandgap and HOMO energy, and triplet energy based on

phosphorescence spectra taken at 77K.

aData determined using solution based cyclic voltammetry-HOMO energy from oxidation potential and LUMO energy from reduction potential [16]. bData based on work in this thesis, discussed further in Chapter 5.

Acronym Name Structure HOMO

(eV)

LUMO

(eV)

Triplet

(ev)

Alq tris(8-hydroxyquinolinato) aluminum

5.7[17]a 2.5[17]

a 2.2[17]

BAlq bis-(2-methyl-8-quinolinolate)-4-

(phenylphenolate)aluminum (III)

5.9[18] 2.9[18] 2.2[18]

BMB 4,4'-bis(3-methylcarbazol-9-yl)-2,2'-

biphenyl

BPhen 4,7-diphenyl-1,10-phenanthroline

6.4[19] 3.0[19] 2.6[20]

Page 67: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

39

Acronym Name Structure HOMO

(eV)

LUMO

(eV)

Triplet

(ev)

FIrpic bis(4,6-difluorophenyl-pyridinato-

N,C2) picolinate iridium (III)

5.6[21]a

5.9[22]

2.5[21]a

3.0[22] 2.6[22]

Ir(F2ppy)3 fac-tris-4,6-difluorophenylpyridine

Ir(III)

5.7[21] 3.0[21] 2.71[17]

Ir(iprpmi)3

fac-tris[(2,6-diisopropylphenyl)-2-

phenyl-1Himidazol[e]iridium(III)

4.8b 2.2

b 2.66

b

Ir(MDQ)2(acac) bis(2-methyldibenzo-[f,h]quinoxaline)

(acetylacetonate)

5.4[23] 2.8[23] 2.0[24]

Ir(piq)3 fac-tris(1-phenylisoquinolinato)

iridium(III)

5.1[25] 3.1[25] 2.0[26]

Page 68: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

40

Acronym Name Structure HOMO

(eV)

LUMO

(eV)

Triplet

(ev)

Ir(ppy)3 fac-tris(2-phenylpyridine)iridium(III)

5.6[27] 3.0[27] 2.49[17]

Ir(tdmp)3

fac-tris[3-(2,6-dimethylphenyl)-7-

methylimidazo[1,2-f] phenanthridine]

iridium (III)

mCBP 3,3'-bis(N-carbazolyl)biphenyl

5.6[28]a

6.0[29]

2.1[28]a

2.4[29] 2.8[28]

NPB 4,4'-bis[N-(1naphthyl)-N-phenyl-

amino]-biphenyl

5.3[17]a

5.4[18]

2.1[17]a

2.3[18] 2.3[18]

PtOEP 2,3,7,8,12,13,17,18-octaethyl-

21H,23H-porphine platinum(II)

Page 69: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

41

Acronym Name Structure HOMO

(eV)

LUMO

(eV)

Triplet

(ev)

TAPC 1,1-bis(di-4-tolylaminophenyl)

cyclohexane

5.5[30] 2.0[30] 2.9[30]

TCTA 4,4',4''-Tris(N-

carbazolyl)triphenylamine

5.7[18] 2.3[18] 2.9[18]

TmPyPB 1,3,5-Tris(3-pyridyl-3-phenyl)benzene

6.7[22] 2.8[22] 2.8[22]

UGH3 m-bis(triphenylsilyl)benzene

7.2[31] 2.8[31] 3.5[31]

Page 70: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

42

REFERENCES

[1] H.L. Caswell, Semicond. Prod. Solid State Technol., 6 (1963) 23.

[2] M.T. Thomas, Methods Exp. Phys., 14 (1979) 521.

[3] D.R. Biswas, J. Mater. Sci., 21 (1986) 2217.

[4] S. Bosch, Trends Vac. Sci. Technol., 2 (1997) 131.

[5] S.M. Rossnagel, J. Vac. Sci. Technol., A, 21 (2003) S74.

[6] S.M. Lee, PhD Thesis, Department of Chemistry, University of Rochester, 2012

[7] T. Matsushima, Y. Kinoshita, H. Murata, Appl. Phys. Lett., 91 (2007) 253504.

[8] H. You, Y. Dai, Z. Zhang, D. Ma, J. Appl. Phys., 101 (2007) 026105.

[9] T. Matsushima, G.-H. Jin, H. Murata, J. Appl. Phys., 104 (2008) 054501.

[10] G. Xie, Y. Meng, F. Wu, C. Tao, D. Zhang, M. Liu, Q. Xue, W. Chen, Y. Zhao,

Appl. Phys. Lett., 92 (2008) 093305/1.

[11] K.S. Yook, J.Y. Lee, Synth. Met., 159 (2009) 69.

[12] H. Heil, J. Steiger, S. Karg, M. Gastel, H. Ortner, H. von Seggern, M. Stossel, J.

Appl. Phys., 89 (2001) 420.

[13] G. Parthasarathy, C. Shen, A. Kahn, S.R. Forrest, J. Appl. Phys., 89 (2001) 4986.

[14] D.Y. Kondakov, J. Appl. Phys., 99 (2006) 024901.

[15] Irfan, H. Ding, Y. Gao, C. Small, D.Y. Kim, J. Subbiah, F. So, Appl. Phys. Lett.,

96 (2010) 243307.

[16] B.W. D'Andrade, S. Datta, S.R. Forrest, P. Djurovich, E. Polikarpov, M.E.

Thompson, Org. Electron., 6 (2005) 11.

[17] M.E. Kondakova, T.D. Pawlik, R.H. Young, D.J. Giesen, D.Y. Kondakov, C.T.

Page 71: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

43

Brown, J.C. Deaton, J.R. Lenhard, K.P. Klubek, J. Appl. Phys., 104 (2008)

094501.

[18] D.P.-K. Tsang, M.-Y. Chan, A.Y.-Y. Tam, V.W.-W. Yam, Org. Electron., 12 (2011)

1114.

[19] Y. Sun, S.R. Forrest, Appl. Phys. Lett., 91 (2008) 263503.

[20] J. Zhao, J. Yu, X. Hu, M. Hou, Y. Jiang, Thin Solid Films, 520 (2012) 4003.

[21] J.J. Brooks, R.C. Kwong, Y.-J. Tung, M.S. Weaver, B.W. D'Andrade, V.

Adamovich, M.E. Thompson, S.R. Forrest, J.J. Brown, Proc. SPIE, 5519 (2004)

35.

[22] J. Lee, J.-I. Lee, J.Y. Lee, H.Y. Chu, Appl. Phys. Lett., 94 (2009) 193305.

[23] R. Meerheim, S. Scholz, S. Olthof, G. Schwartz, S. Reineke, K. Walzer, K. Leo, J.

Appl. Phys., 104 (2008) 014510.

[24] G. Schwartz, S. Reineke, T.C. Rosenow, K. Walzer, K. Leo, Adv. Funct. Mater.,

19 (2009) 1319.

[25] T. Zhang, Y. Liang, J. Cheng, J. Li, J. Mater. Chem. C, 1 (2013) 757.

[26] C. Cai, S.-J. Su, T. Chiba, H. Sasabe, Y.-J. Pu, K. Nakayama, J. Kido, Org.

Electron., 12 (2011) 843.

[27] J.-H. Jou, W.-B. Wang, M.-F. Hsu, J.-J. Shyue, C.-H. Chiu, I.M. Lai, S.-Z. Chen,

P.-H. Wu, C.-C. Chen, C.-P. Liu, S.-M. Shen, ACS Nano, 4 (2010) 4054.

[28] S. Gong, X. He, Y. Chen, Z. Jiang, C. Zhong, D. Ma, J. Qin, C. Yang, J. Mater.

Chem., 22 (2012) 2894.

[29] H. Nakanotani, K. Masui, J. Nishide, T. Shibata, C. Adachi, Sci Rep, 3 (2013)

Page 72: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

44

2127.

[30] J. Lee, J.-I. Lee, J.Y. Lee, H.Y. Chu, Appl. Phys. Lett., 95 (2009) 253304.

[31] X. Ren, J. Li, R.J. Holmes, P.I. Djurovich, S.R. Forrest, M.E. Thompson, Chem.

Mater., 16 (2004) 4743.

Page 73: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

45

CHAPTER 3

FIRPIC AS THE LIGHT-EMITTING DOPANT FOR

PHOSPHORESCENT ORGANIC LIGHT-EMITTING

DIODES

3.1 INTRODUCTION

Organic light emitting diodes (OLED) have recently achieved commercial success

as a display and lighting technology [1]. The impact of using phosphorescent materials

has been discussed in Chapter 1. While recently introduced commercial products employ

phosphorescent red and green emitting materials, the poor performance of blue

phosphorescent emitters, notably the lifetime, has prevented adaptation into commercial

products. Universal Display Corporation (UDC), a leading PHOLED materials

development company, has reported light blue PHOLEDs that have 1931 Commission

Internationale de l’Eclairage (CIE) color coordinates of 0.18, 0.42, luminous efficiency of

50 cd/A and operating lifetime (t50; time for initial luminance to drop by 50%) of 20,000

hours, all at an initial luminance of 1,000 cd/m2 [2]. This lifetime is far inferior to those

of red and green devices reported by UDC, which have lifetimes of 900,000 and 400,000

hours, respectively. The specific materials and device architectures used by UDC have

not been disclosed.

Giebink et al. [3] achieved ~9% EQE and t50 of 700 hours at a current density of

Page 74: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

46

~7 mA/cm2

using the following device architecture, Indium tin oxide (ITO)/10 nm

undisclosed material/30 nm NPB/30 nm BMB + 9% Ir(tdmp)3/5 nm BMB/40 nm

Alq/LiF/Al. In this architecture, ITO is the anode, the undisclosed material is the hole

injecting layer (HIL), NPB is the hole transport layer (HTL), BMB is the light-emitting

layer (LEL) host with Ir(tdmp)3 as the dopant, undoped BMB is the electron transport

layer (ETL), Alq is used as both the electron injection layer (EIL) and ETL, and LiF/Al is

the cathode. The focus of this research revolved around exciton-polaron annihilation

reactions and the effect on the device voltage and lifetime. Yamamoto et al. [4] reported

similar device architecture and performance as Giebink et al., however they focused on

how lifetime is improved by fabricating devices under ultra high vacuum (6.5 × 10-7

torr).

In a U.S. patent, D’Andrade et al. [5] disclosed a similar device structure using the same

materials as Giebink et al. and achieved an EQE of ~7% and a t50 of 10,000 hours at ~4

mA/cm2. While device performance was briefly discussed in each of these references,

none related device lifetime and/or EQE to specific materials or the device’s layer

structure.

FIrpic is a phosphorescent blue emitter that has been widely used in blue

PHOLED devices. Recently, the stability of FIrpic has been examined using laser-

desorption/ionization time-of-flight mass spectrometry (LDI-TOF-MS) [6]. The results

indicated that FIrpic undergoes chemical dissociation, including loss of the picolinate

ligand and loss of CO2. Fragments of FIrpic were shown to react with other materials,

which could contribute to device instability. In another study with LDI-MS, the F atoms

on FIrpic were found to dissociate during device operation [7]. Depending on device

Page 75: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

47

architecture and the choice of host and transport materials, lifetimes ranging from

minutes to 110 hours have been reported with FIrpic as the phosphorescent blue emitter

[6, 8-10]. While there have been few reports on PHOLED device lifetime, there have

been numerous reports on FIrpic achieving an external quantum efficiency (EQE) greater

than 20% using a variety of host and transport materials [11-16]. Relatively few

publications report device lifetime for blue PHOLEDs, especially for devices optimized

specifically for achieving the highest EQE.

In this chapter, device performance is evaluated, including device lifetime, with

FIrpic as the phosphorescent dopant. The goal is to correlate the device performance with

variations in device architecture and material composition. Three different hosts are

compared, TCTA:UGH3 mixed host, TCTA single host, and mCBP single host. For each

of these hosts, a FIrpic concentration series is examined. The affect of varying the ETL

material is also explored. Two different HTL materials for devices with mCBP as the host

are compared.

3.2 EXPERIMENTAL

Device fabrication was discussed in Chapter 2. Molecular structures for all

materials are listed in Figure 3.1.

3.3 MIXED HOST

Mixed host technology has been used to improve charge balance, device

efficiency and lifetime for both fluorescent [17-20] and phosphorescent [21-27] OLEDs.

Instead of using a single host – a host comprising a pure material, a mixed host typically

Page 76: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

48

comprises two material components, usually mixed uniformly in a solid film, one for

Figure 3.1: Molecular structures of materials used in Chapter 3.

transporting electrons and the other for holes. In this way, the electron and hole transport

functions are supported separately in the two host components. Essentially, the mixed

host spatially extends the recombination region by allowing holes to inject into the LEL

through the hole-transport host while electrons are injected into the LEL through the

electron-transport host.

Mixed-host LEL’s have been shown to improve the device lifetime by extending

the recombination zone within the LEL to reduce both charge build-up and recombination

HTL Materials

Host Materials Blue Dopant

ETL Materials

NPB TAPC

TCTA UGH3 mCBP FIrpic

BPhen TmPyPB UGH3 BAlq

Page 77: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

49

cycles per molecule at the LEL/transport layer interface. Recently, an EQE of 21.6% was

achieved using a mixed-host of TCTA and UGH3 with FIrpic as the blue dopant [14].

However, similar to other publications, this article only highlighted the extraordinarily

high EQE achievable in a blue PHOLED without any reference to the device lifetime.

Previously, Lee et al. [27] of our laboratory studied the same mixed-host using a graded

structure, however the lifetime was not studied. In this section, device performance is

evaluated using TCTA:UGH3 as a mixed host. TCTA has a strong electron donating

character with the triphenylamine core substituted with carbazole substituents. This

material is the hole transporting component of the mixed host. UGH3 has

tetraphenylsilane as the core moiety. It has been described as the electron transporting

component for this mixed host system. Both materials have triplet energies higher than

that of FIrpic.

3.3.1 TCTA:UGH3 (1:1) mixed host with varied FIrpic concentration

Figure 3.2: FIrpic concentration series for TCTA:UGH3 mixed host OLED devices.

A series of PHOLED devices was fabricated with a 1:1 ratio of TCTA:UGH3 as

35 nm HTL

TAPC

30 nm ETL

TmPyPB

20 nm EIL

BPhen

40 nm Mixed host

TCTA:UGH3 (1:1)

Emitting dopant

FIrpic

Concentration

9-26%

Anode Cathode

Page 78: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

50

the mixed host where the concentration of FIrpic was varied from 9%–26%. The device

layer structure is as follows: Anode/35 nm TAPC/40 nm TCTA+UGH3 (1:1)+x%

FIrpic/30 nm TmPyPB/20 nm BPhen/Cathode (Figure 3.2). TAPC and TmPyPB have

triplet energies of 2.90 eV and 2.78 eV, respectively, and both are higher than that of

FIrpic (2.62 eV) to prevent exciton quenching. BPhen is known to be a good EIL. The

emission spectra, EQE, and current-voltage characteristics are shown in Figure 3.3 and

the device data at 5 mA/cm2 are shown in Table 3.1.

The electroluminescence (EL) spectra for devices having the lowest (9%) and

highest (26%) FIrpic concentrations as shown in Figure 3.3(a) indicate that the EL

emission originates from FIrpic. Essentially the EL spectra are independent of FIrpic

concentration. The slightly broader emission from the 26% FIrpic device can be

attributed to the solid-state solvation effect [28, 29]. Figure 3.3(b) shows that for all the

devices relatively high EQE’s of ~15-20% are obtained in the low current density range

of 0.1-5 mA/cm2, with the EQE then decreasing steadily with increasing current density.

At 20 mA/cm2, an EQE above 15% is maintained for all the devices except the device

with 26% FIrpic. Even this device, which presumably has a lower EQE due to

concentration quenching [30, 31] or increased triplet-triplet annihilation [32], attains a

high EQE in the range of 13-15% throughout the entire current range. The J-V

characteristics shown in Figure 3.3(c) indicate negligible differences for FIrpic

concentrations between 9%–17%. The shift to slightly higher voltage for the device with

26% FIrpic is likely due FIrpic interfering with the hole transport through the TCTA host.

Page 79: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

51

Table 3.1: OLED performance data at 5 mA/cm2 for varying FIrpic concentration in mixed host

devices; Anode/35 nm TAPC/40 nm TCTA+UGH3 (1:1)+x% FIrpic/30 nm TmPyPB/20 nm

BPhen/Cathode, x=9–26%.

FIrpic (%)

V cd/m2 EQE

(%)

9 6.4 1824 17.2

12 6.4 1987 18.3

17 6.5 1879 17.0

26 6.8 1598 13.8

Figure 3.3: OLED performance data at 5 mA/cm2 for varying FIrpic concentration in mixed host

devices; Anode/35 nm TAPC/40 nm TCTA+UGH3 (1:1)+x% FIrpic/30 nm TmPyPB/20 nm

BPhen/Cathode, x=9–26%; (a) EL spectra at 5 mA/cm2, (b) EQE vs. current density, (c) current

density vs. voltage.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

400 500 600 700

Rela

tiv

e I

nte

nsi

ty (a

.u.)

Wavelength (nm)

9%

26%

FIrpic

(a)

0

5

10

15

20

0.1 1 10 100

EQ

E (

%)

Current Density (mA/cm2)

9%

12%

17%

26%

FIrpic

(b)

0.1

1

10

100

4 6 8 10

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

9%

12%

17%

26%

FIrpic

(c)

Page 80: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

52

3.3.2 Mixed TCTA:UGH3 (x:y) hosts with fixed Flrpic concentration

Figure 3.4: Varying the mixed host doping ratio using FIrpic as emitting dopant.

The previous devices had a fixed host ratio while the FIrpic concentration was

varied. In the following set of devices, the host ratio was varied while fixing the FIrpic

concentration at 14%. The device layer structure is as follows: Anode/35 nm TAPC/40

nm TCTA+UGH3 (x:y)+14% FIrpic/30 nm TmPyPB/20 nm BPhen/Cathode (Figure 3.4).

The sum of x and y is equal to 86%. The TCTA and UGH3 concentrations were varied in

the range of 16%–70%. Figure 3.5(a) shows that the device EL spectra are invariant with

host compositions. Similar to what was observed by Lee et al., the EQE vs J curves

shown in Figure 3.5(b) indicate only minor differences with host composition. From 0.1–

5 mA/cm2, the EQE is 15–18% before decreasing to ~14–15% at 20 mA/cm

2. The J-V

data shown in Figure 3.5(c) show that the voltage is generally lower with increasing

TCTA concentration. The decrease in drive voltage is more than 1 V (at 5 mA/cm2) going

from low to high TCTA concentration. Devices using a mixed host of TCTA:UGH3 are

dominated by hole transport through the TCTA component with the UGH3 behaving

essentially as an inert component responsible for lowering the hole mobility in TCTA.

35 nm HTL

TAPC

30 nm ETL

TmPyPB

20 nm EIL

BPhen

40 nm Mixed host

TCTA:UGH3 (x:y)

Emitting dopant

14% FIrpic

Host Ratio (x:y)

70:16 – 16:70

Anode Cathode

Page 81: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

53

Figure 3.5: OLED performance data at 5 mA/cm2 for varying mixed host composition while

keeping FIrpic at 14%; Anode/35 nm TAPC/40 nm TCTA+UGH3 (x:y)+14% FIrpic/30 nm

TmPyPB/20 nm BPhen/1 nm Cathode, x+y=86%; (a) EL spectra at 5 mA/cm2, (b) EQE vs. current

density, (c) current density vs. voltage.

The role of UGH3 will be discussed later within this chapter. As the TCTA concentration

becomes too low, i.e. 31% and 16%, hole transport is effectively impeded and the drive

voltage increases.

0.1

1

10

100

4 6 8 10

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

70%

55%

31%

16%

TCTA (x)

0

5

10

15

20

0.1 1 10 100

EQ

E (

%)

Current Density (mA/cm2)

70%

55%

31%

16%

TCTA (x)

0.0

0.2

0.4

0.6

0.8

1.0

1.2

400 500 600 700R

ela

tiv

e I

nte

nsi

ty (a

.u.)

Wavelength (nm)

70%

16%

TCTA (x)

(a)

Page 82: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

54

3.3.3 Effect of electron transport materials

Figure 3.6: Removing TmPyPB and using BAlq or BPhen next to the LEL for TCTA:UGH3 mixed

host devices.

It has been reported that the recombination zone for devices utilizing an LEL of

TCTA:UGH3:FIrpic is at or near the LEL/ETL interface [27]. Aza-aromatic compounds

such as TmPyPB that are commonly used as ETL materials have long been suspected of

contributing significantly to OLED device degradation [33]. To assess the degradation

due to the ETL materials, devices with alternate ETL materials were fabricated with the

following layer structure: Anode/35 nm TAPC/40 nm TCTA+UGH3 (1:1)+12%

FIrpic/ETL/Cathode (Figure 3.6). The ETL is either a single layer of 50 nm BPhen or a

bilayer of 10 nm BAlq/50 nm BPhen with BAlq adjacent to the LEL. The EL spectra for

both devices as shown in Figure 3.7(a) indicate that the EL emissions from FIrpic are

slightly modified. The FIrpic EL spectrum typically has a peak at 472 nm with a shoulder

at 496 nm. The EL spectrum for the device with BAlq/BPhen as the ETL has peaks at 472

nm and 496 nm of almost equal intensity. This may be due to 2 reasons; 1) an optical

interference effect [22, 34], and/or 2) emission from BAlq [35]. Light emission from the

LEL is isotropic. For device structures studied in this thesis, light reflects from the Al

35 nm HTL

TAPC

50 nm EIL

BPhen

40 nm Mixed host

TCTA:UGH3 (1:1)

Emitting dopant

12% FIrpic

BAlq

0 nm or 10 nm

ETLAnode Cathode

Page 83: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

55

Figure 3.7: OLED performance data at 5 mA/cm2 for mixed host devices with alternate ETL’s;

Anode/35 nm TAPC/40 nm TCTA+UGH3 (1:1)+12% FIrpic/ETL/Cathode. ETL is 50 nm BPhen or

10 nm BAlq/50 nm BPhen; (a) EL spectra at 5 mA/cm2, (b) EQE vs. current density, (c) current

density vs. voltage.

cathode. There is an optimal distance from the LEL to the cathode that prevents optical

interference as light reflects from the Al. As the thickness between the LEL and cathode

changes, the EQE can also change due to this optical interference. Since recombination is

expected to occur nearest the LEL/ETL interface, it is also possible that BAlq can

contribute to the EL spectra, increasing the peak at 496 nm and cause a slight broadening

of the spectra. The EQE-J data for these two devices with modified ETL’s is shown in

Figure 3.7(b). At 0.1 mA/cm2, the EQE’s are 3-6% before rising to a maximum of 8-9%

at 20 mA/cm2. These EQE values are far less that what was observed for previous devices

0.0

0.2

0.4

0.6

0.8

1.0

1.2

400 500 600 700R

ela

tiv

e I

nte

nsi

ty (a

.u.)

Wavelength (nm)

BPhen

BAlq/BPhen

(a)

0.0

5.0

10.0

0.1 1 10 100

EQ

E (

%)

Current Density (mA/cm2)

BPhen

BAlq/BPhen

0.1

1

10

100

2 6 10

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

BPhen

BAlq/BPhen

(b) (c)

Page 84: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

56

using TmPyPB/BPhen as the ETL. As discussed in Chapter 1, the transport layers should

have higher triplet energy (ET) than the dopant so as to prevent the triplet excitons from

being quenched. Unlike TmPyPB (ET~2.8 eV), neither BPhen (ET~2.6 eV) nor BAlq (ET

~ 2.2 eV) have triplet energies exceeding that of FIrpic (ET~2.6 eV), so these materials

are able to quench triplet excitons, thereby reducing the EQE. These devices are

dominated by hole current at low current densities, therefore the EQE is suppressed due

to recombination occurring predominantly at the LEL/ETL interface. As the current

density increases, the electron injection from the ETL into the LEL is increased, resulting

in a shift of the recombination away from the LEL/ETL interface and consequently

reducing exciton quenching. The device with BPhen as the ETL has a lower EQE

compared to the device with BAlq/BPhen. This is somewhat surprising considering that

BPhen has a higher ET compared to BAlq. One possible explanation is that the host TCTA

forms an exciplex with BPhen, which is a stronger acceptor compared to BAlq, resulting

in the formation of a lower energy complex that can result in additional exciton

quenching. The J-V data shown in Figure 3.7(c) indicates that the devices using BPhen

and BAlq/BPhen as the ETL have lower voltages than those using TmPyPB/BPhen. At 5

mA/cm2, BPhen, BAlq/BPhen, and TmPyPB/BPhen devices have voltages of 5.0 V, 5.8

V, and 6.4 V, respectively. The reason for this trend can be explained by the relative

alignment of the LUMO levels of these materials as shown in Figure 3.8. For all three

devices, electrons are injected into and transported through BPhen. Without the BAlq or

TmPyPB layer, electrons can inject directly from BPhen into the LEL layer and

recombine directly on FIrpic. With BAlq and TmPyPB, there are barriers of 0.1 eV and

Page 85: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

57

0.2 eV for electron injection from BPhen into BAlq and TmPyPB, respectively, resulting

in higher drive voltages.

Figure 3.8: HOMO, LUMO, Triplet Energy levels for materials used to fabricate OLED devices in

Chapter 2. Triplet Energies are in parenthesis. Shaded boxes indicate typical electron transport

materials. All values in eV.

3.3.4 Device lifetime

Device lifetime, t50, measured at a fixed current density of 5 mA/cm2, refers to the

time duration for the initial EQE to decrease by 50%. The lifetime data for the mixed host

devices are shown in Figure 3.9. Regardless of the composition of the mixed hosts and

Flrpic concentrations, all devices using TmPyPB/BPhen as the ETL have a lifetime of

less than 20 minutes as represented by a single curve in Figure 3.9. The devices with

BPhen and BAlq/BPhen as the ETL have longer lifetimes of ~4 hours and ~12 hours,

respectively. The very short lifetime for the TmPyPB/BPhen devices is likely a result of

two factors: 1) these devices have high EQE’s, suggesting that the probability of FIrpic

excited states undergoing dissociation is increased; and 2) electron-hole recombination in

TmPyPB could result in bond breaking and the formation of quencher species. The

2.0

5.5

TCTA

(2.9)

TAPC

(2.9) TmPyPB

(2.8)FIrpic

(2.6)

2.3

5.7

3.0

5.9

6.7

2.8

7.2

2.8

UGH3

(3.5)

2.9

5.9

BAlq

(2.2) BPhen

(2.6)

3.0

6.4

HOMO

LUMO2.4

6.0

mCBP

(2.8)

2.3

5.4

NPB

(2.3)

Page 86: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

58

device with the BPhen ETL has a slightly longer lifetime. The device with BAlq has the

longest lifetime in agreement with previous reports with BAlq used as the ETL. The

stability of BAlq may be attributed to its lower excited state energy as well as the stability

of its radical cation compared to TmPyPB.

Figure 3.9: Device lifetime taken at 5 mA/cm2 for mixed-host devices; Anode/35 nm TAPC/40 nm

TCTA+UGH3 (1:1)+12% FIrpic/ETL/Cathode. ETL is 50 nm BPhen, 10 nm BAlq/50 nm BPhen,

or 30 nm TmPyPB/20 nm BPhen.

3.4 SINGLE HOST

While a mixed host is known to improve device performance, it is often beneficial

to simplify the device structure for the ease of fabrication and for a better understanding

of the device operation. To this end, there has been significant research effort aimed at

designing and synthesizing bipolar host materials that can perform both hole transport

and electron transport functions [36-41]. The HOMO/LUMO levels can be tuned based

on the functional groups within the molecule, and typically include both hole transport

and electron transport moieties that are electronically isolated from each other using

various spacer groups or substitution patterns. In this way, only a single host material is

0.5

0.6

0.7

0.8

0.9

1

1.1

0 5 10

Rela

tiv

e E

QE

Time (hours)

TmPyPB/BPhen

BAlq/BPhen

BPhen

ETL

Page 87: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

59

required within the LEL.

Carbazole-based derivatives have been used extensively as host materials for

phosphorescent device [21, 42-44]. CBP is one such derivative that reportedly has bipolar

characteristics [45, 46]. While this material does not have an electron withdrawing

moiety, the central biphenyl group linking the two carbazole units allows for -orbital

overlap that is conducive for electron transport. CBP has a triplet energy of 2.6 eV; which

is low enough that it can quench FIrpic triplet excitons. We fabricate a series of devices to

compare the performance of two different single host carbazole-based materials, TCTA

and mCBP. Both materials have triplet energies higher than FIrpic. TCTA was used as the

hole transporting component in the mixed-host system in section 3.3. This material lacks

a biphenyl moiety and is not considered a bipolar host material. In contrast, mCBP is

similar to CBP and has been used to transport both holes and electrons [5, 47]. Similar to

section 3.3, the ETL materials are also varied for these two host materials. The HTL is

varied for devices with mCBP as the host.

3.4.1 TCTA single host and FIrpic concentration series

Figure 3.10: FIrpic concentration series for TCTA single host OLED devices.

35 nm HTL

TAPC

30 nm ETL

TmPyPB

20 nm EIL

BPhen

40 nm Single host

TCTA

Emitting dopant

FIrpic

Concentration

3-26%

Anode Cathode

Page 88: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

60

In this series TCTA is used as the single host component in a layer structure as

follows: Anode/35 nm TAPC/40 nm TCTA+x% FIrpic/30 nm TmPyPB/20 nm

BPhen/Cathode (Figure 3.10). The range of Flrpic concentration in TCTA is 3–26%.

Device data taken at 5 mA/cm2 is shown in Table 3.2.

Table 3.2: Device data taken at 5 mA/cm2 for single host TCTA and FIrpic concentrations 3-6%.

FIrpic (%)

V cd/m2

EQE (%)

t50

(min.)

3 6.6 1176 10.9 3

6 6.7 1137 10.6 4

11 6.5 1523 13.8 7

26 6.4 1264 10.8 25

Similar to the mixed host devices, devices utilizing TCTA as a single host are

dominated by hole transport, with recombination occurring at the LEL/ETL interface.

The drive voltage is relatively independent of the FIrpic concentration. The EQE ranges

from 11–14%, with the highest efficiency occurring at a concentration of 12% FIrpic.

These EQE values are considerably lower than what was observed for mixed host devices

(17-18%). A possible explanation for the lower EQE is the formation of exciplex between

TCTA and TmPyPB. Exciplexes are known to form between hole transporting materials

and electron transporting materials [22, 48, 49]. With TCTA as the single host,

recombination is expected to occur at TCTA/TmPyBP interface and therefore exciplex

formation at this interface will likely affect the EQE. We examined the possibility of

exciplex formation using the structure: Anode/30 nm TAPC/30 nm TCTA/30 nm

TmPyPB/Cathode, where recombination is expected to occur at the TCTA/TmPyPB

Page 89: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

61

interface. The EL spectrum shown in Figure 3.11 shows a broad peak centered at 500 nm

which can be attributed to the TCTA/TmPyPB exciplex emission. The 415nm peak is

from TCTA [48]. The TCTA/TmPyPB exciplex, which has a lower energy than FIrpic, is

expected to quench FIrpic excitons, resulting in a lower EQE. An alternative mechanism

for the lower EQE is quenching due to TCTA cations [50]. A large build-up of TCTA

cations at the TCTA/TmPyPB interface could result in FIrpic exciton quenching due to

polaron-exciton interactions.

Figure 3.11: EL spectrum at 20 mA/cm2 showing exciplex of TCTA:TmPyPB for undoped device.

Anode/30 nm TAPC/30 nm TCTA/30nm TmPyPB/Cathode.

As shown in Table 3.2, the lifetime of the single-host devices is very short,

ranging from 3 to 25 minutes. There is a weak dependence on FIrpic concentration with

lifetime increasing as FIrpic concentration increases. This trend suggests that FIrpic is not

the material limiting the lifetime. As mentioned earlier, it is quite likely that TmPyPB is

the key culprit contributing to the short device lifetime.

0

0.2

0.4

0.6

0.8

1

1.2

400 500 600 700

Rela

tiv

e I

nte

nsi

ty (a

.u.)

Wavelength (nm)

TCTA:TmPyPB

Exciplex emission peak

Anode/TAPC/TCTA/TmPyPB/Cathode

Page 90: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

62

3.4.2 TCTA single host and UGH3 as electron transport material

Figure 3.12: Removing TmPyPB and using UGH3 or BPhen next to the LEL for single host

devices with FIrpic concentration at 1% or 12%.

In this series of devices, TmPyPB was replaced by UGH3 as the ETL. UGH3 has

been used in mixed host systems [14, 51] as well as for the ETL [51-53]. There have been

no reports of device lifetime specifically related to this material. UGH3 lacks

heterocyclic moieties that have been linked to instability. The device structure is as

follows: Anode/35 nm TAPC/40 nm TCTA+12% FIrpic/x nm UGH3/50 nm

BPhen/Cathode (Figure 3.12). The UGH3 thickness is varied from 0–30 nm. Device data

at 5 mA/cm2 is shown in Table 3.3.

Without the UGH3 layer, BPhen is in direct contact with the LEL, supporting both

electron injection and transport. The drive voltage is only 4.7 V, apparently due to the

absence of barriers to electron transport from BPhen to the LEL. The EQE is 4.7%, which

is low due to triplet exciton quenching by BPhen. Inserting a 2.5 nm UGH3 layer raised

the voltage by 0.4 V to 5.1 V. The drive voltage rises to 8.8 V with a 20 nm UGH3 layer.

35 nm HTL

TAPC

50 nm EIL

BPhen

UGH3

0–30 nm

ETL40 nm Single host

TCTA

Emitting dopant

FIrpic

Concentration

1% or 12%

Anode Cathode

Page 91: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

63

Table 3.3: OLED performance data at 5 mA/cm2 for devices using UGH3 as the ETL and 12%

FIrpic in the LEL; Anode/35 nm TAPC/40 nm TCTA+12% FIrpic/x nm TmPyPB/y nm UGH3/z

nm BPhen/Cathode.

TmPyPB (x nm)

UGH3 (y nm)

BPhen (z nm)

V cd/m2

EQE (%)

t50 (min.)

― 0 50 4.7 491 4.7 200

― 2.5 50 5.1 1203 11.5 35

― 5.0 50 5.3 1336 13.1 10

― 10.0 50 6.6 281 2.7 11

― 20.0 50 8.8 352 3.0 120

30 ― 20 6.5 1523 13.8 7

― 30 20 10.2 318 3.2 200

In order to understand the voltage dependence, the following two devices were

fabricated: Anode/35 nm TAPC/40 nm TCTA+12% FIrpic/30 nm UGH3 or TmPyPB/20

nm BPhen/Cathode. Table 3.3 shows the device performance for these two devices at 5

mA/cm2. The voltages are 6.5 V and 10.2 V for the TmPyPB and UGH3 devices,

respectively. Despite the total ETL/BPhen thicknesses being the same, the UGH3 device

voltage is almost 4 V higher than the TmPyPB device. There is a barrier of 0.2 eV (Figure

3.8) for electron injection from BPhen into either UGH3 or TmPyPB. Considering that

the electron barrier heights are the same and the UGH3 device has a higher voltage

compared to the TmPyPB device, this would indicate that UGH3 has a much lower

electron mobility compared to TmPyPB.

With 2.5 nm and 5.0 nm layers of UGH3, the EQE’s are 9.2% and 11.7%, more

than double that of the device without UGH3. Further increasing the UGH3 thickness to

10.0–20.0 nm results in a significant decrease in EQE (to 2‒3%). It would seem that a

thin UGH3 layer, i.e. 2.5 nm or 5.0 nm, is effective for reducing triplet exciton quenching

by BPhen while still allowing for electrons to transit the UGH3 layer and to recombine

Page 92: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

64

with the holes at the LEL/ETL interface. It is inferred that a minimum thickness of 5 nm

of UGH3 is required to separate the LEL and BPhen layers. In fact, it is possible that as

the UGH3 thickness increases from 10.0 to 20.0 nm, electron injection from the BPhen

layer is so impeded that the holes are drawn into the UGH3 to cause electron-hole

recombination within the UGH3 layer or at the UGH3/BPhen interface.

Lifetimes for all devices with UGH3 as the ETL are improved compared to

devices with TmPyPB as the ETL. For the device using BPhen as the ETL and EIL, the

lifetime is highest, reaching 200 minutes. The lifetime decreases to 35 min. for the device

with 2.5 nm of UGH3 and levels out at 10 minutes for UGH3 thicknesses between 5.0–

10.0 nm. With UGH3 thicknesses of 20 nm and 30 nm, the lifetime increases to 120

minutes and 200 minutes, respectively. While lifetime can be improved by using UGH3

instead of TmPyPB, the EQE is generally lower due to UGH3 being a poor electron

transporting material.

The data shown here suggests that UGH3 does not support electron transport in a

mixed host with TCTA. It is more likely that UGH3 impedes hole transport in the mixed

host, and thus causes a shift of recombination away from the LEL interface. UGH3 could

also reduce the formation of exciplex between TCTA and TmPyPB. For these reasons, the

EQE’s for the mixed TCTA/UGH3 host systems are higher than those of the TCTA single

host systems.

To further evaluate UGH3 as the ETL, the following structure was fabricated:

Anode/35 nm TAPC/40 nm TCTA+1% FIrpic/x nm UGH3/50 nm BPhen/Cathode, where

the UGH3 thickness is varied from 0–20 nm. Device data is shown in Table 3.4. Flrpic

Page 93: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

65

concentration was kept at 1%. By using a low FIrpic concentration, recombination

occurring directly on FIrpic is reduced. Device lifetime will become more dependent on

recombination occurring on other materials. If the device lifetime increases at low FIrpic

concentrations, FIrpic is limiting the lifetime. However, if device lifetime decreases at

low FIrpic concentrations, device lifetime is being limited by other materials within the

device.

Table 3.4: OLED performance data at 5 mA/cm2 for devices using UGH3 as the ETL and 1%

FIrpic in the LEL. Anode/35 nm TAPC/40 nm TCTA+12% FIrpic/x nm TmPyPB/y nm UGH3/z nm

BPhen/Cathode.

UGH3

(nm) V cd/m

2

EQE (%)

t50

(min.)

0 4.7 41 0.5 12

2.5 5.2 290 2.9 8

5.0 5.4 419 4.2 4

10.0 6.4 121 1.2 1

20.0 9.0 299 2.6 56

The EL spectra shown in Figure 3.13 indicate that the emission occurs

predominantly from FIrpic for all devices regardless of the UGH3 thicknesses. All the

devices with UGH3 show additional emission occurring around 410 nm that is attributed

to TCTA (Figure 3.13, inset). The device without UGH3 has a broad shoulder from 400-

450 nm that is possibly due to an exciplex formed between TCTA and BPhen. It is not

possible for all excitons to form on FIrpic with the concentration being so low. As a

result, the excitons are formed on the materials at the LEL/ETL interface.

Since TCTA dominates hole transport and recombination occurs at the TCTA/ETL

interface, the voltages for these low FIrpic concentration devices range from 4.7–9.0 V,

Page 94: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

66

Figure 3.13: EL spectra at 5 mA/cm2 for devices using UGH3 as the ETL and 1% FIrpic in the

LEL. Anode/35 nm TAPC/40 nm TCTA+12% FIrpic/y nm UGH3/50 nm BPhen/Cathode, y=0–20

nm. Inset: Enhanced region from 400-450 nm showing TCTA emission.

which is comparable to the devices that had 12% FIrpic (4.7–8.8 V). The EQE’s for these

1% devices are significantly reduced compared to the 12% devices, reaching a maximum

of 4.2% for the device with 5 nm of UGH3. As shown by the EL spectra, the lower EQE

is due to a reduction in the concentration of highly efficient FIrpic recombination centers,

resulting in incomplete capture of the triplet excitons by the FIrpic molecules. Emission

is observed from the inefficient transport layers. The device lifetimes for the 1% devices

range from 1–56 minutes, much shorter than what was observed for the 12% devices (7–

200 minutes). With device lifetime decreasing as the FIrpic concentration is reduced,

FIrpic is presumably not the key factor in limiting the device lifetime. This implicates the

host and transport layers as culprits for causing device degradation if recombination

occurs on these materials.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

400 500 600 700

Rel

ati

ve

Inte

nsi

ty (a.u

.)

Wavelength (nm)

0 nm

5 nm

20 nm

UGH3

0.00

0.02

0.04

0.06

0.08

400 425 450

Rela

tiv

e I

nte

nsi

ty

Wavelength (nm)

Page 95: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

67

3.4.3 TCTA host and BAlq as electron transport material

Figure 3.14: BAlq or BPhen next to the LEL for TCTA single host devices.

BAlq was used as the ETL for the mixed host devices (section 3.3.3) to improve

device lifetime. Similar improvements were observed in single host devices with the

following structure: Anode/35 nm TAPC/40 nm TCTA+12% FIrpic/x nm BAlq/50 nm

BPhen/Cathode (Figure 3.14). This structure is similar to the devices using UGH3 as the

ETL. The BAlq thickness is varied from 0–20 nm. Device data is shown in Table 3.5.

Electroluminescence originates from FIrpic.

Table 3.5: OLED performance data at 5mA/cm2 for devices using TCTA as single host and BAlq as

the ETL. Anode/35 nm TAPC/40 nm TCTA+12% FIrpic/x nm BAlq/50 nm BPhen/Cathode.

BAlq (nm)

V cd/m2

EQE (%)

t50

(min.)

0 4.6 539 5.0 200

2.5 4.9 705 5.6 450

5.0 5.1 843 7.3 390

10.0 5.6 844 7.2 350

20.0 6.6 821 6.5 390

The voltage increases with increasing BAlq thickness, which is the same behavior

35 nm HTL

TAPC

50 nm EIL

BPhen

40 nm Mixed host

TCTA:UGH3 (1:1)

Emitting dopant

12% FIrpic

BAlq

0 nm or 10 nm

ETLAnode Cathode

Page 96: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

68

that was observed for the UGH3 series of devices. This is most likely due to the fact that

with a sufficiently thin layer, both UGH3 and BAlq behave as tunneling layers separating

the BPhen layer from the LEL layer, and as such the drive voltages are similar. For

thicker layers of 10.0 nm and 20.0 nm, the voltages for BAlq devices are 5.6 V and 6.6 V

respectively compared to 6.7 V and 8.8 V for the UGH3 devices. The lower voltages for

BAlq based devices are likely due to higher electron mobility in BAlq compared to

UGH3. With thinner films (2.5 to 5.0 nm), the BAlq devices have lower EQE’s compared

to those with UGH3 because BAlq, with a triplet energy lower than that of FIrpic, is

capable of quenching FIrpic excitons. For BAlq or UGH3 thicknesses of 10.0 nm or 20.0

nm, the BAlq devices have higher EQE’s than the UGH3 devices. The field across BAlq

is lower than that of UGH3, resulting in improved electron transport and

recombination/emission occurring from the LEL/ETL interface. The larger field across

UGH3 results in reduced recombination at the LEL/ETL interface, with the possibility

that some percentage of recombination occurs within the UGH3 layer or possibly at the

UGH3/BPhen interface. While BAlq (ET=2.2 eV) has a lower triplet energy than BPhen

(ET=2.6 eV) and would be expected to at least quench FIrpic excitons to the same degree,

the devices with BAlq have higher EQE’s than BPhen only devices. It is possible that due

to a lower electron injection barrier into the LEL, recombination for the BAlq devices is

moved slightly away from the quenching interface.

The lifetimes for devices with BAlq reach 6–8 hours, which is a significant

improvement over devices with UGH3 (10 minutes to 3 hours). This lifetime

improvement is due to two reasons. 1) For thin layers, i.e. 2.5 nm and 5.0 nm, the BAlq

Page 97: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

69

devices have lower EQE’s which is due to fewer FIrpic molecules undergoing radiative

emission. As a result, fewer FIrpic molecules degrade and lifetime increases. 2) For

thicker layers, i.e. 10.0 nm and 20.0 nm, the EQE for the BAlq devices doesn’t shift

much, so lifetime is relatively constant. For the UGH3 devices, the EQE decreases

dramatically due to recombination likely occurring within UGH3. Recombination outside

the LEL is more detrimental to device lifetime than recombination occurring within the

LEL.

3.4.4 mCBP host and FIrpic concentration series

Figure 3.15: FIrpic concentration series for mCBP single host devices.

As mentioned earlier, stable PHOLEDs using mCBP as the host material with

Ir(tdmp)3 as the blue phosphorescent dopant have been demonstrated. At an initial

luminance of 500 cd/m2, device lifetimes up to 10,000 hours have been achieved. The

HOMO and LUMO energy values for mCBP are 6.0 eV and 2.4 eV, respectively, both

deeper than TCTA (Figure 3.8). The triplet energy of 2.8 eV is higher than FIrpic. In the

following structure, TCTA was replaced by mCBP in the LEL: Anode/35 nm

TAPC/mCBP+x% FIrpic/30 nm TmPyPB/20 nm BPhen/Cathode (Figure 3.15), where the

35 nm HTL

TAPC

30 nm ETL

TmPyPB

20 nm EIL

BPhen

40 nm Single host

mCBP

Emitting dopant

FIrpic

Concentration

2-18%

Anode Cathode

Page 98: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

70

FIrpic concentration is varied from 2–18%. Device performance is shown in Figure 3.16

and data taken at 5 mA/cm2 is shown in Table 3.6.

Table 3.6: OLED performance at 5 mA/cm2 for devices using mCBP as host and varying the FIrpic

concentration from 2–18%.

FIrpic (%)

V cd/m2 EQE

(%)

t50

(min.)

2 9.2 1661 15.5 -

6 9.0 1762 16.3 4

12 8.7 1619 14.4 6

18 8.4 1231 10.4 9

The EL spectra shown in Figure 3.16(a) indicate that emission occurs from FIrpic

for all devices and is independent of FIrpic concentration. Figure 3.16(b) show the J-V

data. As the concentration of FIrpic increases from 2 to18%, the voltage decreases

monotonically from 9.2 V to 8.4 V at 5 mA/cm2, which is different from the TCTA-based

devices. In TCTA devices, there was essentially no change in voltage as FIrpic

concentration increased. Furthermore, the mCBP-based devices require significantly

higher voltages regardless of the dopant concentration. TCTA-based devices are

dominated by hole transport, with holes building up at the TCTA/ETL interface where

recombination occurs. As a result, there was little change in voltage even if the dopant

concentration was over 20%. The higher voltage for mCBP devices results from two

factors, 1) with mCBP as the host, there is a 0.5 eV hole injection barrier from TAPC into

mCBP, creating a field across the LEL and 2) mCBP lacks the triarylamine moiety that is

present in TCTA, making it likely that mCBP has a lower hole mobility. These two

Page 99: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

71

factors lead to fewer holes building up at the LEL/ETL interface, allowing electrons to

more easily inject from TmPyPB directly onto the FIrpic molecules. As the FIrpic

concentration increases, electron injection is also easier and the voltage decreases.

Figure 3.16(c) shows the EQE-J data. At 0.1 mA/cm2 the EQE’s for the 2–12%

FIrpic devices range from 8%–14%. The EQE’s rise to ~14–17% at 2 mA/cm2 before

decreasing again at higher current densities. These EQE’s are generally higher than what

was observed for TCTA single-host devices (~14% maximum EQE) and lower than the

mixed-host devices (~20% maximum EQE). The TCTA single-host devices had triplet

quenching due to exciplex formation between TCTA and TmPyPB whereas the UGH3

Figure 3.16: OLED performance for devices with mCBP as host and 2-18% FIrpic, Anode/35 nm

TAPC/mCBP+x% FIrpic/30 nm TmPyPB/20 nm BPhen/Cathode.

0.1

1

10

100

4 8 12

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

2%

6%

12%

18%

0

4

8

12

16

0.1 1 10 100

EQ

E (

%)

Current Density (mA/cm2)

2%

6%

12%

18%

0.0

0.2

0.4

0.6

0.8

1.0

1.2

400 500 600 700

Rela

tiv

e I

nte

nsi

ty (a

.u.)

Wavelength (nm)

2%

6%

12%

18%

(a)

(b) (c)

Page 100: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

72

component of the mixed host helps to prevent the TCTA:TmPyPB exciplex from

forming. The following simplified device was fabricated to investigate whether mCBP

also forms an exciplex with TmPyPB: Anode/75 nm mCBP/75 nm TmPyPB/Cathode.

Figure 3.17 shows the EL spectrum for this device taken at 20 mA/cm2. The peak

centered at ~550 nm is not due to either mCBP [54] or TmPyPB [55]. This indicates that

this is due to an exciplex formed between these two materials. The formation of this

exciplex is likely the reason why the EQE’s are lower than what was observed for mixed-

host devices. Since fewer holes arrive at the mCBP/TmPyPB interface compared to the

TCTA/TmPyPB interface, the mCBP:TmPyPB exciplex may not be as strong as the

TCTA:TmPyPB exciplex, resulting in higher EQE’s for the mCBP devices compared to

Figure 3.17: EL spectrum at 20 mA/cm2 showing exciplex of mCBP:TmPyPB. Anode/75 nm

mCBP/75 nm TmPyPB/Cathode.

TCTA single-host devices. This exciplex may also explain the EQE trend observed with

current density. At low current densities, hole transport is expected to dominate over

electron transport, with recombination occurring at the mCBP/TmPyPB interface. As the

current density increases, more electrons arrive at the mCBP/TmPyPB interface which

0.0

0.2

0.4

0.6

0.8

1.0

1.2

400 500 600 700

Rela

tiv

e I

nte

nsi

ty (a

.u.)

Wavelength (nm)

Anode/mCBP/TmPyPB/Cathode

mCBP:TmPyPB

Exciplex emission peak

Page 101: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

73

decreases the formation of exciplex and increases the EQE. The EQE for the device with

18% FIrpic stays between 8–10% regardless of the current density. This low EQE is

attributed to either dopant self-quenching or triplet-triplet annihilation.

The device lifetimes using mCBP as host are all very short, less than 10 minutes

at 5 mA/cm2. This is similar to what was observed for other devices utilizing FIrpic as the

emitting dopant with TmPyPB as the electron transport layer.

3.4.5 mCBP host and BAlq as electron transport material

Figure 3.18: Removing TmPyPB and using BAlq or BPhen next to the LEL for mCBP single host

devices with the FIrpic concentration at 6% or 12%.

BAlq was used as the electron transport for both mixed-host and TCTA single-

host devices to improve lifetime. It was also used to evaluate lifetime with mCBP as the

host in the following device structure: Anode/35 nm TAPC/40 nm mCBP+12% FIrpic/x

nm BAlq/50 nm BPhen/Cathode (Figure 3.18). The BAlq ETL is varied from 0‒20 nm.

The FIrpic concentration was fixed at 12%, the same as what was used in the TCTA

35 nm HTL

TAPC

50 nm EIL

BPhen

BAlq

0–20 nm

ETL40 nm Single host

mCBP

Emitting dopant

FIrpic

Concentration

6% or 12%

Anode Cathode

Page 102: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

74

device structure with BAlq. Device data at 5 mA/cm2 are shown in Table 3.7.

Electroluminescence originates from FIrpic.

As the BAlq thickness increases from 0–10 nm, the voltage increases from 6.7–

8.5 V. This trend, which was also observed for the devices with TCTA as the host and

BAlq as the ETL, arises due to the generation of an electric field across BAlq based on

the electron injection barrier of 0.2 eV from BPhen into BAlq (Figure 3.8). With BAlq

quenching triplet excitons, the EQE ranges from 4.8–6.7%, which is about half of what

was achieved compared to devices with TmPyPB as the ETL. The lifetime ranges from

75–220 minutes, an improvement due to the lower EQE and the removal of TmPyPB

from the device.

Table 3.7: OLED performance data at 5 mA/cm2 for devices using mCBP as the single host and

BAlq as the ETL. Anode/35 nm TAPC/40 nm mCBP+12% FIrpic/x nm BAlq/50 nm

BPhen/Cathode.

BAlq (nm)

V cd/m2 EQE

(%)

t50

(min.)

0 6.7 736 6.6 210

2.5 7.0 771 6.7 90

5.0 7.1 770 6.6 220

10.0 7.4 762 6.4 75

20.0 8.5 620 4.8 110

While the previous devices that replaced TmPyPB with BAlq had FIrpic doped at

12% to compare with the TCTA devices, doping 6% FIrpic yields the optimum EQE for

mCBP devices. In the following device structure, BAlq was varied from 0–20 nm while

FIrpic was fixed at 6%: Anode/35 nm TAPC/40 nm mCBP+6% FIrpic/x nm BAlq/50 nm

BPhen/Cathode. Device performance is shown in Figure 3.13.

Page 103: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

75

As the BAlq thickness increases from 0–10 nm, the voltage increases from 7.8–

8.9 V, a trend that has been observed and described for other devices using BAlq. BAlq

also quenches the EQE, resulting in a range from ~6.6-8%. These values are

approximately half those achieved compared to devices using TmPyPB as the ETL. The

lifetime ranges from 250-600 minutes, which is similar to the TCTA devices that use

BAlq, and an improvement over the mCBP devices that 12% FIrpic and have BAlq as the

ETL. By decreasing the FIrpic concentration, the lifetime improves.

Table 3.8: OLED performance data at 5 mA/cm2 for devices using mCBP as the single host and

BAlq as the ETL. Anode/35 nm TAPC/40 nm mCBP+6% FIrpic/x nm BAlq/50 nm

BPhen/Cathode.

BAlq (nm)

V cd/m2

EQE t50

(min.)

0 7.8 848 8.0 250

2.5 8.0 714 6.6 600

5.0 8.0 830 7.4 425

10.0 8.3 856 7.5 400

20.0 8.9 793 6.6 500

Page 104: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

76

3.4.6 mCBP host and NPB as hole transport layer

Figure 3.19: Using BAlq as ETL and NPB as HTL at FIrpic concentrations of 6% and 12%.

In the prior experiments, it was found that modification of the ETL significantly

impacts device performance. In order to understand how the choice of HTL impacts

device performance, TAPC was replaced with NPB in the following device structure:

Anode/35 nm NPB/40 nm mCBP+12% FIrpic/x nm BAlq/50 nm BPhen/Cathode,

whereas the BAlq thickness varies from 5–20 nm. The FIrpic concentration was fixed at

12%. NPB has a lower triplet energy (2.3 eV) compared to TAPC (2.9 eV), however

TAPC has been found to be less stable than NPB as a hole transporting material. Device

data taken at 5 mA/cm2 is shown in Figure 3.16.

Electroluminescence originates from FIrpic. Compared to the devices in Table 3.7

that had TAPC as the HTL, these devices have a voltage that is approximately 1 V higher

while the EQE drops by ~50%. The higher voltage may be due to a higher injection

barrier from NPB into the mCBP host while the lower EQE indicates that NPB quenches

50 nm EIL

BPhen

BAlq

5–20 nm

ETL40 nm Single host

mCBP

Emitting dopant

FIrpic

Concentration

6% or 12%

35 nm HTL

NPB

Anode Cathode

Page 105: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

77

the triplet excitons.

The devices with NPB and BAlq have a 4-6x lifetime improvement compared to

those with TAPC and BAlq shown in Table 3.7. This is likely due to a decrease in

radiative recombination (from NPB quenching excitons) and a subsequent decrease in the

resulting excited state reactions. It also may be due to NPB being more stable compared

to TAPC.

In order to understand how NPB affects device performance with a FIrpic

concentration of 6%, devices were fabricated using the following structure: Anode/35 nm

NPB/40 nm mCBP+6% FIrpic/x nm BAlq/50 nm BPhen/Cathode. The BAlq thickness

was either 5 nm or 10 nm. Device performance at 5 mA/cm2 is shown in Table 3.9.

Table 3.9: OLED performance data at 5 mA/cm2 for devices using NPB as HTL, BAlq as ETL, and

x% FIrpic. Anode/35 nm NPB/40 nm mCBP+x% FIrpic/y nm BAlq/50 nm BPhen/Cathode.

FIrpic (%)

BAlq (nm)

V cd/m2 EQE

(%)

t50

(min.)

12 5.0 8.1 341 2.9 450

12 10.0 8.5 338 2.8 500

12 20.0 9.5 286 2.3 550

6 5.0 8.7 677 6.5 380

6 10.0 9.0 755 7.1 380

The 5 nm and 10 nm BAlq devices have voltages of 8.7 V and 9.0 V, respectively.

The EQE’s are ~7% while the lifetimes are around 400 minutes. With a FIrpic

concentration of 12%, replacing TAPC with NPB resulted in a voltage increase of 1.0 V

and an EQE decrease of 50%. By adjusting the FIrpic concentration to 6%, replacing

TAPC with NPB resulted in the voltage increasing by 0.7 V while the EQE decrease was

Page 106: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

78

marginal. As mentioned earlier, the voltage increase is attributed to the hole injection

barrier into the LEL being larger for NPB. It is expected that the devices using mCBP as

the HTL are still dominated by hole transport, with recombination occurring primarily at

the LEL/ETL interface. Switching to NPB didn’t significantly affect the EQE for the 6%

FIrpic devices, unlike what was observed for the 12% devices, indicating a shorter

exciton diffusion length for decreased FIrpic concentrations. The reduced FIrpic

concentration results in fewer excitons diffusing to and getting quenched at the NPB/LEL

interface. The lifetime for the 6% FIrpic devices are also very similar, regardless if the

HTL is NPB or TAPC.

3.5 CONCLUSIONS

With the use of high triplet charge transport materials, TAPC as the HTL and

TmPyPB as the ETL, high EQE’s over 15% were achieved with FIrpic as the light-

emitting dopant. While the efficiency is high the device lifetimes are all very short (less

than 20 minutes at 5 mA/cm2) regardless of the host configuration for the emitting layer.

Device lifetime can be improved to 2–11 hours by replacing TmPyPB with

alternate ETL materials such as BAlq, BPhen, and UGH3. However the EQE is reduced

significantly by utilizing these materials. BAlq and BPhen have lower triplet energies

than FIrpic and quench triplet excitons. UGH3 has a high triplet energy of 3.5 eV and

does not quench FIrpic excitons. However, due to a large electron injection barrier into

UGH3 (from BPhen) accompanied by low electron mobility, UGH3 is a poor electron

transport material and prevents efficient recombination from occurring in the LEL.

However, a thin layer of UGH3 between the LEL and BPhen can be used effectively as a

Page 107: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

79

buffer to reduce quenching from the lower triplet energy BPhen (2.6 eV).

Use of a mixed host LEL (TCTA:UGH3) in combination with the low triplet

energy BAlq ETL resulted in the best combination for EQE (~9%) and lifetime (~12

hours). This lifetime is far greater compared to the lifetime of a few minutes which is

observed by using TmPyPB as the ETL. With UGH3 being a poor electron transport

material, its role within the mixed host LEL is that of a spacer group, slowing down hole

transport to mitigate recombination occurring at the LEL/ETL interface.

A study of undoped devices indicates that TmPyPB forms charge transfer

complexes (exciplex) with both TCTA and mCBP. This explains the enhanced efficiency

in mixed host devices utilizing UGH3 and TCTA. With UGH3 being used primarily as a

spacer component, the direct contact between TCTA and TmPyPB is decreased, which

limits or prevents exciplex formation.

It was found that devices with 1% FIrpic doped into TCTA have shorter lifetimes

than devices with 12% FIrpic. At 1% doping, there are not enough dopant molecules to

capture all excitons. Instead, excitons form on the host and/or transport materials and

emission occurs from these materials. This indicates that recombination and emission

occurring on the host/transport materials is detrimental to device lifetime.

Using mCBP as a single host does not improve lifetime. The voltage is higher if

mCBP is used as the host compared to TCTA, indicating that the hole injection barrier

into mCBP is higher and/or the hole mobility in mCBP is decreased. The EQE’s for

mCBP devices are higher than TCTA devices, which is explained by 1) a reduction in

charge quenching by accumulated holes at the LEL/ETL interface; and 2) a decrease in

Page 108: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

80

the formation of the mCBP:TmPyPB exciplex (compared to TCTA:TmPyPB exciplex)

that can quench triplet excitons.

The effect of replacing the high triplet energy HTL TAPC (2.9 eV) with low

triplet energy NPB (2.3 eV) was dependent on the concentration of FIrpic within the

LEL. For devices with 12% of FIrpic doped into mCBP, replacing TAPC with NPB

results in the EQE decreasing by ~50%. Reducing the FIrpic concentration to 6%, there is

minimal loss in EQE after replacing TAPC with NPB. This difference in the quenching

between the two different FIrpic concentrations is attributed to a reduction of triplet

excitons (which need to diffuse through FIrpic molecules) at the HTL/LEL interface.

With higher concentrations of FIrpic, the triplet exciton formed at the LEL/ETL interface

is more likely to diffuse to the HTL/LEL interface to get quenched. There is also ~1 V

increase in drive voltage after replacing TAPC with NPB, indicating increased hole

accumulation at the NPB/LEL interface that can also adversely affect the emission

efficiency. Device lifetime is approximately the same regardless if NPB or TAPC is used.

It is concluded that the blue phosphorescent dopant FIrpic along with the transport

and host materials all contribute to device instability. FIrpic is extremely unstable as a

light-emitting material, regardless of the host or transport materials. Low triplet energy

ETL’s such as BAlq or BPhen quench triplet excitons, effectively slowing down excited

state reactions involving FIrpic. Recombination occurring on the host or aza-aromatic

electron transport materials is also detrimental to device lifetime, indicating that these

materials are also unstable in the excited state.

Page 109: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

81

REFERENCES

[1] Mertens, R., www.oled-info.com/devices, 2014

[2] Universal Display Corporation, www.udcoled.com, 2014

[3] N.C. Giebink, B.W. D'Andrade, M.S. Weaver, P.B. Mackenzie, J.J. Brown, M.E.

Thompson, S.R. Forrest, J. Appl. Phys., 103 (2008) 044509.

[4] H. Yamamoto, J. Brooks, M.S. Weaver, J.J. Brown, T. Murakami, H. Murata,

Appl. Phys. Lett., 99 (2011) 033301.

[5] B. D'Andrade, P.B. MacKenzie, M.S. Weaver, J.J. Brown, U.S. Patent No.

8,557,399 (2013).

[6] I. Rabelo de Moraes, S. Scholz, B. Luessem, K. Leo, Org. Electron., 12 (2011)

341.

[7] V. Sivasubramaniam, F. Brodkorb, S. Hanning, H.P. Loebl, V. van Elsbergen, H.

Boerner, U. Scherf, M. Kreyenschmidt, J. Fluorine Chem., 130 (2009) 640.

[8] W. Sotoyama, T. Satoh, M. Kinoshita, M. Tobise, K. Kawato, T. Ise, H. Takizawa,

S. Yamashita, Dig. Tech. Pap. - Soc. Inf. Disp. Int. Symp., 41 (2010) 556.

[9] H. Lee, H. Ahn, C. Lee, J. Inf. Disp., 12 (2011) 219.

[10] K.S. Yook, O.Y. Kim, J.Y. Lee, Synth. Met., 161 (2011) 2677.

[11] N. Chopra, J. Lee, Y. Zheng, S.-H. Eom, J. Xue, F. So, Appl. Phys. Lett., 93

(2008) 143307.

[12] H. Sasabe, E. Gonmori, T. Chiba, Y.-J. Li, D. Tanaka, S.-J. Su, T. Takeda, Y.-J. Pu,

K.-i. Nakayama, J. Kido, Chem. Mater., 20 (2008) 5951.

[13] S.-J. Su, E. Gonmori, H. Sasabe, J. Kido, Adv. Mater., 20 (2008) 4189.

Page 110: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

82

[14] J. Lee, J.-I. Lee, J.Y. Lee, H.Y. Chu, Appl. Phys. Lett., 94 (2009) 193305.

[15] L. Xiao, S.-J. Su, Y. Agata, H. Lan, J. Kido, Adv. Mater., 21 (2009) 1271.

[16] C.W. Lee, J.Y. Lee, Adv. Mater., 25 (2013) 1.

[17] V.-E. Choong, S. Shi, J. Curless, C.-L. Shieh, H.-C. Lee, F. So, J. Shen, J. Yang,

Appl. Phys. Lett., 75 (1999) 172.

[18] A.B. Chwang, R.C. Kwong, J.J. Brown, Appl. Phys. Lett., 80 (2002) 725.

[19] D. Ma, C.S. Lee, S.T. Lee, L.S. Hung, Appl. Phys. Lett., 80 (2002) 3641.

[20] Y. Shao, Y. Yang, Appl. Phys. Lett., 83 (2003) 2453.

[21] S.H. Kim, J. Jang, K.S. Yook, J.Y. Lee, M.-S. Gong, S. Ryu, G.-k. Chang, H.J.

Chang, J. Appl. Phys., 103 (2008) 054502/1.

[22] M.E. Kondakova, T.D. Pawlik, R.H. Young, D.J. Giesen, D.Y. Kondakov, C.T.

Brown, J.C. Deaton, J.R. Lenhard, K.P. Klubek, J. Appl. Phys., 104 (2008)

094501.

[23] J. Lee, J.-I. Lee, J.Y. Lee, H.Y. Chu, Org. Electron., 10 (2009) 1529.

[24] K.H. Kim, J.Y. Lee, T.J. Park, W.S. Jeon, G.P. Kennedy, J.H. Kwon, Synth. Met.,

160 (2010) 631.

[25] D.P.-K. Tsang, M.-Y. Chan, A.Y.-Y. Tam, V.W.-W. Yam, Org. Electron., 12 (2011)

1114.

[26] C.W. Seo, J.Y. Lee, Thin Solid Films, 520 (2012) 7022.

[27] S. Lee, C.W. Tang, L.J. Rothberg, Appl. Phys. Lett., 101 (2012) 043303.

[28] V. Bulovic, A. Shoustikov, M.A. Baldo, E. Bose, V.G. Kozlov, M.E. Thompson,

S.R. Forrest, Chem. Phys. Lett., 287 (1998) 455.

Page 111: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

83

[29] V. Bulovic, R. Deshpande, M.E. Thompson, S.R. Forrest, Chem. Phys. Lett., 308

(1999) 317.

[30] Y. Kawamura, J. Brooks, J.J. Brown, H. Sasabe, C. Adachi, Phys. Rev. Lett., 96

(2006) 017404.

[31] G. Ramos-Ortiz, Y. Oki, B. Domercq, B. Kippelen, Phys. Chem. Chem. Phys., 4

(2002) 4109.

[32] M.A. Baldo, M.E. Thompson, S.R. Forrest, Nature (London), 403 (2000) 750.

[33] V.V. Jarikov, K.P. Klubek, L.-S. Liao, C.T. Brown, J. Appl. Phys., 104 (2008)

074914.

[34] C.-J. Lee, D.-Y. Yoon, D.-G. Moon, D.-K. Choi, J.-I. Han, ECS Trans., 16 (2009)

17.

[35] N.H. Kim, Y.-H. Kim, J.-A. Yoon, S.Y. Lee, H.H. Yu, A. Turak, W.Y. Kim, MRS

Online Proc. Libr., 1567 (2013) opl 2013 734.

[36] C.-H. Chen, W.-S. Huang, M.-Y. Lai, W.-C. Tsao, J.T. Lin, Y.-H. Wu, T.-H. Ke,

L.-Y. Chen, C.-C. Wu, Adv. Funct. Mater., 19 (2009) 2661.

[37] S.-y. Takizawa, V.A. Montes, P. Anzenbacher, Jr., Chem. Mater., 21 (2009) 2452.

[38] Y.-M. Chen, W.-Y. Hung, H.-W. You, A. Chaskar, H.-C. Ting, H.-F. Chen, K.-T.

Wong, Y.-H. Liu, J. Mater. Chem., 21 (2011) 14971.

[39] S. Gong, Y. Chen, J. Luo, C. Yang, C. Zhong, J. Qin, D. Ma, Adv. Funct. Mater.,

21 (2011) 1168.

[40] W. Jiang, L. Duan, J. Qiao, G. Dong, L. Wang, Y. Qiu, Org. Lett., 13 (2011) 3146.

[41] C. Fan, F. Zhao, P. Gan, S. Yang, T. Liu, C. Zhong, D. Ma, J. Qin, C. Yang, Chem.

Page 112: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

84

- Eur. J., 18 (2012) 5510.

[42] K. Brunner, A. Van Dijken, H. Boerner, J.J.A.M. Bastiaansen, N.M.M. Kiggen,

B.M.W. Langeveld, J. Am. Chem. Soc., 126 (2004) 6035.

[43] P.-I. Shih, C.-L. Chiang, A.K. Dixit, C.-K. Chen, M.-C. Yuan, R.-Y. Lee, C.-T.

Chen, E.W.-G. Diau, C.-F. Shu, Org. Lett., 8 (2006) 2799.

[44] L.-S. Cui, S.-C. Dong, Y. Liu, Q. Li, Z.-Q. Jiang, L.-S. Liao, J. Mater. Chem. C, 1

(2013) 3967.

[45] T. Yamada, F. Suzuki, A. Goto, T. Sato, K. Tanaka, H. Kaji, Org. Electron., 12

(2011) 169.

[46] T. Zhang, Y. Liang, J. Cheng, J. Li, J. Mater. Chem. C, 1 (2013) 757.

[47] D.B. Knowles, C. Lin, P.B. Mackenzie, J.-Y. Tsai, R.W. Walters, S.A. Beers, C.S.

Brown, W.H. Yeager, U.S. Patent No. 7,915,415 (2011).

[48] Y.-S. Park, K.-H. Kim, J.-J. Kim, Appl. Phys. Lett., 102 (2013) 153306.

[49] Y.-S. Park, S. Lee, K.-H. Kim, S.-Y. Kim, J.-H. Lee, J.-J. Kim, Adv. Funct. Mater.,

23 (2013) 4914.

[50] R.H. Young, J.R. Lenhard, D.Y. Kondakov, T.K. Hatwar, SID Int. Symp. Digest

Tech. Papers, 39 (2008) 705.

[51] J. Lee, J.-I. Lee, H.Y. Chu, Synth. Met., 159 (2009) 991.

[52] J. Lee, J.-I. Lee, K.-I. Song, S.J. Lee, H.Y. Chu, Appl. Phys. Lett., 92 (2008).

[53] J. Lee, J.-I. Lee, J.Y. Lee, H.Y. Chu, Synth. Met., 159 (2009) 1956.

[54] P. Schroegel, N. Langer, C. Schildknecht, G. Wagenblast, C. Lennartz, P.

Strohriegl, Org. Electron., 12 (2011) 2047.

Page 113: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

85

[55] S.-J. Su, T. Chiba, T. Takeda, J. Kido, Adv. Mater., 20 (2008) 2125.

Page 114: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

86

CHAPTER 4

CHARGE TRANSPORT THROUGH FIRPIC

4.1 INTRODUCTION

As discussed in Chapter 1 and Chapter 3, FIrpic is one of the most widely used Ir

cyclometalated dopants used for testing new blue emitting PHOLED device architectures

and materials. High external quantum efficiencies (EQE) exceeding 20% can be achieved

with this material used as an emitter doped into a host matrix. However, regardless of the

choice of transport and host materials, use of this material as an emitter results in OLED

devices that have lifetimes well below 50 hours, which is a lifetime far removed from

what is required for display and solid-state lighting applications.

A recent study [3] showed that FIrpic undergoes irreversible chemical dissociation in

an operating device, producing fragments that can quench excitons and reduce the OLED

efficiency. Several possible pathways for FIrpic degradation have been proposed,

suggesting mainly the involvement of FIrpic excited states formed as a result of electron-

hole recombination in the dopant/host matrix. In this chapter, the effect of charge

transport in a FIrpic doped mCBP matrix on the stability of the OLED device is

investigated. We conclude that hole transport on FIrpic and charge recombination on

mCBP both lead to rapid device degradation and derive insight into materials selection

criteria.

Page 115: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

87

4.2 EXPERIMENTAL

Molecular structures along with highest occupied molecular orbital (HOMO) and

lowest unoccupied molecular orbital (LUMO) energy levels for all materials in this

chapter are shown in Figure 4.1. Additional information for these materials are shown in

Table 2.1 in Chapter 2. Device fabrication procedures were discussed in Chapter 2. A 1.0

nm thin film of MoOx was deposited on ITO as a hole-injecting layer (HIL). Neat NPB,

FIrpic doped NPB, neat mCBP and FIrpic doped mCBP were used as the hole transport

layers (HTL). Alq served as the electron transport layer (ETL) as well as the light emitter.

Figure 4.1: Molecular structures and HOMO/LUMO energy levels of materials used in devices.

Units in electronvolts (eV).

2.1 2.1

5.3

NPB

+

FIrpic

5.6

ITO

4.7

LiF/Al

4.1

MoOx

2.5

Alq

5.65.7

2.5

mCBP

+

FIrpic

2.5

AlqNPB

FIrpic

mCBP

Page 116: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

88

4.3 RESULTS AND DISCUSSION

The layer structures of a series of OLEDs designed to probe the effect of HTL

formulation on OLED performance are listed in Table 4.1. All devices have in common

an HTL/ETL device structure with Alq as the ETL and emitter. The thicknesses of the

HTL and ETL are 75 nm each.

Table 4.1: OLED device structure and initial EQE and voltage at 20 mA/cm2. ITO/1 nm

MoOx/HTL/75 nm Alq/1 nm LiF/100 nm Al.

Device HTL: Voltage

(V)

EQE

(%)

Ref. 75 nm NPB 8.3 1.2

A1 75 nm mCBP 8.3 1.5

A2 37.5 nm mCBP|37.5 nm NPB 7.5 1.1

B1 37.5 nm mCBP+1% FIrpic|37.5 nm NPB 9.0 1.1

B2 37.5 nm mCBP+3% FIrpic|37.5 nm NPB 9.3 1.0

B3 37.5 nm mCBP+6% FIrpic|37.5 nm NPB 9.8 1.1

B4 37.5 nm mCBP+12% FIrpic|37.5 nm NPB 9.2 1.3

C1 37.5 nm mCBP|27.5 nm mCBP+12% FIrpic|10 nm NPB 8.4 1.3

D1 37.5 nm mCBP|37.5 nm mCBP+12% FIrpic 7.9 1.5

E1 75 nm NPB+6% FIrpic 7.9 1.1

Electroluminescence originates from Alq for all devices (Figure 4.2). Alq has an

EL maximum around 530 nm. Each of the HTL materials used in this study emits at a

shorter wavelength. The EL maximum is 470 nm for NPB [6] and ~468-472 nm for

FIrpic [7, 8]. mCBP has a PL maximum at 340 nm in solution [9]. It is clear from Figure

4.2 that the EL spectra are solely characteristic of Alq without any contribution from the

HTL materials. While it is generally possible for electrons to inject into the HTL from the

emitter layer in OLED devices, the devices in this study have a large energy barrier of 0.4

eV for electron injection from Alq into either mCBP or NPB (Figure 4.1). The electron

concentration in the HTL should be small, if any. If electrons were to penetrate deep into

Page 117: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

89

the HTL, then there would be emission from the HTL material. However, as already

noted, there is no such emission. Recombination at the HTL/Alq interface yields emission

only from Alq due to its smaller energy gap.

Figure 4.2: EL spectra at 20 mA/cm2 for the following device architecture: ITO/1.0 nm MoOx/75

nm HTL/75 nm Alq/1.0 nm LiF/100 nm Al, where the HTL’s are: Ref [75 nm NPB], A1 [75 nm

mCBP], B4 [37.5 nm mCBP+12% FIrpic|37.5 nm NPB], C1 [37.5 nm mCBP|27.5 nm mCBP+12%

FIrpic|10 nm NPB], D1 [37.5 nm mCBP|37.5 nm mCBP+12% FIrpic], E1 [75 nm NPB+6%

FIrpic].

4.3.1 Undoped hole transport layers

The HTL of the reference (Ref), A1 and A2 devices are NPB, mCBP, and

mCBP/NPB, respectively. The Ref and A1 have a single-layer HTL sandwiched between

Alq and the ITO/MoOx anode. A2 has a bi-layer HTL with NPB adjacent to Alq and

mCBP in contact with MoOx. As shown in Figure 4.3(a), the J-V-L characteristics are

very similar for all three devices, independent of the HTL formulation. However, as

shown in Figure 4.3(b), the EQE decay (lifetime) is quite different with A1 degrading

much faster than Ref and A2. The t50 (time to 50% decay in EQE) for A1 is only about 90

hours, whereas the t80 (time to 80% decay in EQE) for Ref and A2 is greater than 200

0

0.2

0.4

0.6

0.8

1

1.2

350 450 550 650 750

Rel

ati

ve

Inte

nsi

ty (a.u

.)

Wavelength (nm)

Ref.

A1

B4

C1

D1

E1

Page 118: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

90

Figure 4.3: OLED performance for devices with an undoped HTL adjacent to MoOx.

ITO/MoOx/HTL/Alq/LiF/Al. Ref[NPB], A1[mCBP], A2[mCBP/NPB]. (a) Initial current density-

voltage-luminance (J-V-L) characteristics, (b) Device lifetime at 20 mA/cm2, (c) Operational

voltage rise at 20 mA/cm2.

hours. These lifetime results support previous findings by Kondakov et al. [10, 11] that

electron-hole recombination at the HTL/Alq interface produces HTL excited states which

can lead to the dissociation of the HTL molecules. The probability for dissociation is

1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

0

20

40

60

80

100

120

2.0 6.0 10.0 14.0 18.0

Lu

min

ance

(cd

/m2 )

Cu

rren

t D

en

sity

(m

A/c

m2 )

Voltage (V)

Ref.

A1

A2

(a)

0.5

0.6

0.7

0.8

0.9

1.0

1.1

0 50 100 150 200 250 300

EQ

E (

a.u

.)

Time (hrs.)

Ref.

A1

A2

(b)

4

6

8

10

12

14

0 50 100 150 200 250 300

Vo

ltag

e (V

)

Time (hrs.)

Ref.

A1

A2

(c)

Page 119: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

91

higher if the excited state energy is larger than the bond dissociation energy, which is the

case for carbazole-based HTL’s such as mCBP, where C-N or C-C homolytic bond

cleavage can occur. As shown in Figure 4.3(c), there is essentially no difference between

A1 and A2 in voltage rise, which is in distinct contrast to the vast difference in EQE

decay. This indicates that mCBP and NPB are equally stable with respect to hole

transport, but mCBP is unstable with respect to charge recombination at the mCBP/Alq

interface. Evidently, luminescence quenchers are more likely to form at the mCBP/Alq

interface.

4.3.2 FIrpic doped into mCBP adjacent to the anode

The next set of devices, B1–B4 of Table 1, were designed for the purpose of

evaluating the stability of FIrpic with respect to hole transport only. Similar to A2, the

HTL of B1–B4 is a bi-layer mCBP/NPB stack. Unlike A2, which has an undoped mCBP

layer, the mCBP layer of B1–B4 is doped with FIrpic ranging in concentration from 1 to

12%. Since this doped layer is spaced away from the Alq emitter layer by a layer of NPB,

it presumes to have little effect on charge recombination at the NPB/Alq interface. As

shown in Figure 4.4(a) and Table 1, the FIrpic doped devices generally require a higher

drive voltage (9.0–9.8 V compared to 7.9 V for the undoped device A2). This increase in

drive voltage may indicate: 1) reduced hole injection efficiency from the MoOx in the

presence of FIrpic, and/or 2) increased hole trapping in the Flrpic doped mCPB layer. It

has been reported that hole injection efficiency from ITO/MoOx into the HTL can be

Page 120: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

92

Figure 4.4: OLED performance for devices with FIrpic doped mCBP adjacent to MoOx.

ITO/MoOx/mCBP+x% FIrpic/Alq/LiF/Al. B1[1%], B2[3%], B3[6%], B4[12%]. (a) Initial

current density-voltage-luminance (J-V-L) characteristics, (b) Device lifetime at 20 mA/cm2, (c)

Operational voltage rise at 20 mA/cm2.

affected by electron transfer from the HTL to MoOx [12]. mCBP and FIrpic

coincidentally have a similar HOMO of 5.6 eV, and thus FIrpic can accept direct hole

injection from MoOx and consequently can affect the hole injection efficiency as well as

1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

0

20

40

60

80

100

120

2.0 6.0 10.0 14.0 18.0

Lu

min

ance

(cd

/m2 )

Cu

rren

t Den

sity

(m

A/c

m2 )

Voltage (V)

Ref.

B1

B2

B3

B4

(a)

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

0 25 50 75 100 125 150

EQ

E (

a.u

.)

Time (hrs.)

Ref.

B1

B2

B3

B4

(b)

5.0

7.0

9.0

11.0

13.0

15.0

17.0

19.0

0 25 50 75 100 125 150

Vo

ltag

e (V

)

Time (hrs.)

Ref.

B1

B2

B3

B4

(c)

Page 121: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

93

the hole transport in mCPB.

As shown in Figure 4.4(b) the EQE degradation is only modestly dependent on

the concentration of FIrpic in mCBP. At low FIrpic concentrations, devices B1 (1%), B2

(3%) and B3 (6%) show a monotonic decrease in EQE at a rate similar to the Ref device,

indicating that the FIpic doped mCBP layer, which is remote from the NPB/Alq interface,

has little effect on EQE. With a higher FIrpic concentration, device B4 (12%) exhibits a

slight rise in EQE before undergoing a steeper decay compared to devices B1–B3. In any

case, the degradation in EQE is relatively minor compared to the large rise in drive

voltage. Figure 4.4(c) shows that the drive voltage increases for all devices at a rate that

increases with FIrpic concentration. The voltage increase (after 200 h of operation at 20

mA/cm2) is as high as 88% for B4 with 12% FIrpic compared to 6.7% for the Ref device

without FIrpic, and 12%, 16.7% and 23.1% for B1, B2, and B3, respectively. This

suggests that FIrpic radical cations are unstable in the mCBP matrix and deep hole traps

are formed as a result of current stress. Since FIrpic can effectively compete with mCBP

for hole transport, more hole traps are likely to form in the HTL with a higher

concentration of FIrpic, which in turn will give rise to a larger increase in the drive

voltage in the OLED devices. On the other hand, the change in EQE reflects the

degradation at the NPB/Alq interface where electron-hole recombination is confined and

should be independent of the degradation in the FIrpic doped mCBP layer, which is

remote from the NPB/Alq interface. This is the case for B1–B3 devices over the course of

200 h. In the B4 device, the initial rise in EQE with drive voltage may be attributed to a

reduction of hole accumulation at the NPB/Alq interface which affects favorably the

Page 122: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

94

charge balance factor and thus a reduction in quenching by NPB radical cations [13]. The

subsequent steep decrease in EQE coincides with a faster rise in drive voltage, indicating

that hole traps are formed at a faster rate in the FIrpic doped mCBP layer, which in turn

could cause electric field induced degradation or even electron-hole recombination in this

layer with decreased emission efficiency.

4.3.3 FIrpic doped mCBP sandwiched between undoped hole transport materials

In order to further understand the cause of the voltage rise, we fabricated device

C1 where the HTL consists of a tri-layer with a 12% FIrpic doped mCBP layer (27.5nm)

sandwiched between an undoped mCBP layer (37.5 nm) and an NPB layer (10 nm). As

shown in Figure 4.5(a), the J-V-L characteristics of C1 are similar to the Ref. device. The

fact that there is little voltage shift provides further support that the presence of FIrpic

doped mCBP near MoOx in the HTL affects hole injection as observed in the B1–4 series

of devices. Figure 4.5(b) shows that the EQE for C1 increases modestly to ~1.05 within

the first 130 hours then undergoes a steep decrease. This behavior is similar to B4 which

also has a high concentration of FIrpic (12%) in mCBP.

Also shown in Figure 4.5(a) and Figure 4.5(b) are the J-V-L characteristics and

EQE decay, respectively, for device D1 where, unlike C1, the FIrpic doped mCBP layer is

in direct contact with Alq. Whereas the J-V-L characteristics of C1 and D1 are similar,

their EQE decay behaviors are dramatically different. D1 suffers a severe loss of EQE

and reaches t50 at 48 h as compared to a 10% increase in EQE for C1 up to 115 h, after

which EQE starts to decrease. Figure 4.5(c) shows that the voltage rise during operation

Page 123: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

95

Figure 4.5: OLED performance for devices with FIrpic doped mCBP spaced away from MoOx and

for a device with FIrpic doped in NPB. ITO/MoOx/mCBP/HTL/NPB/Alq/LiF/Al:

C1[mCBP+12%], ITO/MoOx/HTL/Alq/LiF/Al]: D1[mCBP/mCBP+12% FIrpic] and

E1[NPB+6% FIrpic]. (a) Initial current density-voltage-luminance (J-V-L) characteristics, (b)

Device lifetime at 20 mA/cm2, (c) Operational voltage rise at 20 mA/cm2.

correlates with the EQE decay for both C1 and D1. The voltage for D1 increases by 90%

at the t50 for EQE decay. The voltage change in C1 is fairly modest during the first 115

hours before undergoing a steep increase that correlates with the EQE loss, similar to

1.E-02

1.E-01

1.E+00

1.E+01

1.E+02

1.E+03

1.E+04

0

20

40

60

80

100

120

2.0 6.0 10.0 14.0 18.0

Lu

min

ance

(cd

/m2 )

Cu

rren

t Den

sity

(m

A/c

m2 )

Voltage (V)

Ref.

C1

D1

E1

(a)

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

0 50 100 150 200 250 300

EQ

E (

a.u

.)

Time (hrs.)

Ref.

C1

D1

E1

(b)

5.0

7.0

9.0

11.0

13.0

15.0

17.0

19.0

0 50 100 150 200 250 300

Vo

ltage (

V)

Time (hrs.)

Ref.

C1

D1

E1

(c)

Page 124: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

96

what was observed for B4. It is also noted that the initial drive voltages for both C1 and

D1 are lower compared to B1–B4. This is due to the fact the FIpic doped layer in these

two devices is not in contact with MoOx, providing further evidence that direct contact

between MoOx and FIrpic doped mCBP impedes hole injection.

For device D1, because the FIrpic doped mCBP layer is in direct contact with Alq,

it is likely that recombination at this interface involves FIrpic, resulting in the formation

of quencher species originating from FIrpic radical cations. This would lead to a loss in

EQE and a rise in drive voltage that is more severe compared to C1 where FIrpic is

spaced away from the recombination interface by a layer of NPB and thus not involved in

direct recombination. The fact that the t50 for D1 is even less than that of A1 which has a

neat mCBP layer as the HTL also indicates that recombination directly on FIrpic is a

channel for EQE degradation.

4.3.4 FIrpic doped into NPB

To further validate that FIrpic is unstable with respect to hole transport, device E1

was fabricated where the HTL is a layer of NPB doped with 6% Flrpic. Figure 4.5(a)

shows that the J-V-L of E1 is comparable to the Ref device. Comparing the EQE losses in

D1 and E1 in Figure 4.5(b), it can be seen that E1 is far more stable and closely

resembles the Ref device. Similarly the drive voltage rise for E1, as shown in Figure

4.5(c), is relatively modest with very little difference between E1 and the Ref device. As

previously mentioned, FIrpic and mCBP have similar HOMO’s, so they both can

contribute to hole transport in an HTL containing FIrpic doped in mCBP. In contrast, in

an HTL containing Flrpic doped NPB, holes can only transport in NPB because FIrpic,

Page 125: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

97

with a much deeper HOMO than NPB, should be incapable of transporting or trapping

holes. In fact, this also explains why the initial drive voltage for E1 is low compared to

B1–B4 devices even though the FIrpic doped NPB layer is in direct contact with MoOx.

FIrpic is precluded to compete with NBP for holes injected from MoOx because of its

deeper HOMO.

4.4 CONCLUSIONS

In summary, we have shown that hole transport and electron-hole recombination

involving FIrpic both result in OLED degradation. Using FIrpic doped mCBP as an HTL

in conjunction with Alq as the emitter/ETL layer, we found that hole transport in FIrpic

primarily results in voltage rise caused by hole trap formation in the HTL. In contrast,

electron-hole recombination in FIrpic results in a large EQE loss as well as significant

voltage rise. We attribute the loss in device performance to the instability of FIrpic radical

cations. One possible consequence of FIrpic oxidation is reportedly the breaking of the Ir-

O bond, followed by loss of CO2 and formation of FIrpic+ [3]. This product would

contribute to the voltage rise we observed in our devices. The detailed chemical pathway

for FIrpic degradation requires further investigation. These results indicate that

FIrpic/mCBP is unsuitable for practical use as an emitter in OLED devices. A possible

solution is to avoid Flrpic radical cation formation by pairing Flrpic with a host with

which emission is through energy transfer from the host to Flrpic rather than through

hole-trapping or transporting in Flrpic.

Page 126: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

98

REFERENCES

[1] R.J. Holmes, S.R. Forrest, Y.J. Tung, R.C. Kwong, J.J. Brown, S. Garon, M.E.

Thompson, Appl. Phys. Lett., 82 (2003) 2422.

[2] J.J. Brooks, R.C. Kwong, Y.-J. Tung, M.S. Weaver, B.W. D'Andrade, V.

Adamovich, M.E. Thompson, S.R. Forrest, J.J. Brown, Proc. SPIE, 5519 (2004)

35.

[3] I. Rabelo de Moraes, S. Scholz, B. Luessem, K. Leo, Org. Electron., 12 (2011)

341.

[4] M.E. Kondakova, T.D. Pawlik, R.H. Young, D.J. Giesen, D.Y. Kondakov, C.T.

Brown, J.C. Deaton, J.R. Lenhard, K.P. Klubek, J. Appl. Phys., 104 (2008)

094501.

[5] S. Gong, X. He, Y. Chen, Z. Jiang, C. Zhong, D. Ma, J. Qin, C. Yang, J. Mater.

Chem., 22 (2012) 2894.

[6] S.C. Tse, K.C. Kwok, S.K. So, Appl. Phys. Lett., 89 (2006) 262102.

[7] C.W. Lee, K.S. Yook, J.Y. Lee, Org. Electron., 14 (2013) 1009.

[8] J. Zhuang, W. Li, W. Su, Y. Liu, Q. Shen, L. Liao, M. Zhou, Org. Electron., 14

(2013) 2596.

[9] P. Schroegel, N. Langer, C. Schildknecht, G. Wagenblast, C. Lennartz, P.

Strohriegl, Org. Electron., 12 (2011) 2047.

[10] D.Y. Kondakov, J. Appl. Phys., 104 (2008) 084520.

[11] D.Y. Kondakov, T.D. Pawlik, W.F. Nichols, W.C. Lenhart, J. Soc. Inf. Disp., 16

(2008) 37.

Page 127: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

99

[12] T. Matsushima, Y. Kinoshita, H. Murata, Appl. Phys. Lett., 91 (2007) 253504.

[13] R.H. Young, J.R. Lenhard, D.Y. Kondakov, T.K. Hatwar, SID Int. Symp. Digest

Tech. Papers, 39 (2008) 705.

Page 128: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

100

CHAPTER 5

INVESTIGATING BLUE PHOSPHORESCENT IRIDIUM

DOPANT WITH PHENYL-IMIDAZOLE LIGANDS

5.1 INTRODUCTION

As discussed in Chapters 1 and 3, excellent device lifetime (>100,000 h) has been

achieved for red and green PHOLEDs while the lifetime for blue PHOLEDs remains

relatively short (typically less than 10,000 h) [1] and inadequate for most applications,

such as display and lighting. Therefore, considerable research efforts have focused on

developing efficient and stable blue phosphorescent PHOLEDs [2]. However, there are

few reports on the structure-property relationships for phosphorescent molecules as they

pertain to device lifetime. FIrpic is among the most studied phosphorescent molecules.

As discussed in Chapter 3, when FIrpic is used as a dopant in blue PHOLEDs, high

external quantum efficiencies (EQE) can be achieved and there are also reports that

EQE’s greater than 20% have been achieved [3-8]. However, FIrpic doped PHOLEDs are

very unstable, with reported lifetimes in the range of 0.1-110 hours depending on the

device architecture and test conditions [9-12]. Based on lifetime testing at a current

density of 5 mA/cm2 that was outlined in Chapter 3, we have observed FIrpic-based

devices having lifetimes ranging from minutes up to 12 hours depending on the choice of

host and transport materials. The instability of FIrpic doped PHOLEDs has been largely

Page 129: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

101

attributed to the intrinsic instability of Flrpic molecules upon excitation. Both the

ancillary picolinate ligand and the fluorine substituted phenyl-pyridyl ligands are

susceptible to photo-induced dissociation [11, 13, 14]. As discussed in Chapter 4, FIrpic

is also unstable to hole transport [15]. To avoid the detrimental structural features present

in FIrpic, an Ir complex with phenyl-imidizole ligands has been examined, one which

reportedly can provide PHOLEDs with lifetimes as long as 10,000 hours [16-22].

In this chapter, a detailed study of Ir(iprpmi)3 is performed. The results presented

here will show that its performance as a blue phosphorescent dopant is not only related to

its intrinsic photophysical properties, but also highly dependent on the compositions of

the host matrix and the adjacent transport layers.

5.2 MATERIAL AND METHODS

Figure 5.1 shows the molecular structures and the energy level diagram for all

materials used in this work. Additional details regarding these materials were discussed in

Chapter 2. The device fabrication procedures were also outlined in Chapter 2.

UV-Vis absorption spectra were obtained using a Perkin Elmer Lambda 750

spectrophotometer. PL and phosphorescent spectra were recorded on a Hitachi F-4600

fluorescence spectrophotometer. The quantum efficiency was obtained by using fac-

Ir(ppy)3 as a standard [23]. The transient spectrum was acquired with an Edinburgh

FLS920 spectrometer using the time-correlated single-photon counting option and data

was analyzed by F900 software (Edinburgh Instruments). The transient spectrum and

quantum efficiency were conducted at room temperature in degassed dichloromethane.

Degassed solutions were prepared by purging with argon for 30 minutes. Cyclic

Page 130: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

102

voltammetry (CV) was carried out on a CHI600 voltammetric analyzer at room

temperature with a conventional three-electrode configuration consisting of a platinum

disk working electrode, a platinum wire auxiliary electrode, and an Ag wire pseudo-

reference electrode with ferrocenium-ferrocene (Fc+/Fc) as the internal standard.

Deaerated dichloromethane was used as solvent. UPS analyses were carried out with an

unfiltered HeI (21.2 eV) gas discharge lamp and a hemispherical analyzer. DFT

calculations were performed at the B3LYP level. The 6-31g(d) basis set was employed for

H, C, N atoms and a “double-ζ” quality basis set, LANL2DZ, was employed for the

Ir(III) metal atom. All calculations were carried out using Gaussian 03 [24].

Figure 5.1 Molecular structures and energy level diagram of the materials (values in eV).

Hole Transport Materials Electron Transport Materials

LEL Host LEL Dopant

TmPyPB

6.7

2.8

2.3

5.4

mCBP

+

dopant

NPB

2.4

6.0

ITO/MoOx

LiF/Al2.9

BAlq

5.9

2.2

4.8

TAPC

5.5

2.0

NPB TAPC TmPyPB BAlq

Ir(iprpmi)3mCBP

Page 131: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

103

5.3 RESULTS AND DISCUSSION

5.3.1 Photophysical and electrochemical measurements

Figure 5.2 shows the UV-Vis absorption and photoluminescence (PL) spectra of

Ir(iprpmi)3. The absorption bands below 330 nm are due to the spin-allowed ligand-

centered (LC) 1(-

*) transitions of the phenyl-imidizole moiety [18, 25-27]. The band at

359 nm is attributed to a spin-allowed metal-to-ligand charge transfer (MLCT) transition

[27] while bands longer than 359 nm are due to mixtures of spin-forbidden LC and

MLCT transitions. The PL spectrum has maximum peak intensity at 474 nm which is

similar to what has been reported for FIrpic (470-472 nm) [18, 28-30]. There is a

Figure 5.2: UV-Vis absorption and PL spectra of Ir(iprpmi)3 in CH2Cl2. Inset: PL spectrum of

Ir(iprpmi)3 in 2-MeTHF at 77 K.

secondary peak at 504 nm and a shoulder at ~550 nm. The solution quantum efficiency

for Ir(iprpmi)3 in CH2Cl2 is 0.57. The triplet energy (ET) is 2.66 eV as determined by the

0

0.2

0.4

0.6

0.8

1

1.2

0

0.1

0.2

0.3

0.4

0.5

250 350 450 550 650 750 850

Rel

ativ

e P

L I

nte

nsi

ty (

a.u

.)

Ex

tin

ctio

n C

oef

fici

ent (1

05M

-1cm

-1)

Wavelength (nm)

Absorption

Photoluminescence

0

0.2

0.4

0.6

0.8

1

1.2

400 500 600

Rela

tiv

e P

L I

nte

nsi

ty (a

.u.)

Wavelength (nm)

Page 132: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

104

peak intensity of the shortest wavelength peak in the PL spectrum at 77 K (inset of Figure

5.2). The transient PL taken in degassed CH2Cl2 indicates a lifetime of 1.19 s

(Appendix, Figure A1). The highest occupied molecular orbital (HOMO) is 4.8 eV based

on ultraviolet photoelectron spectroscopy (UPS) while the lowest unoccupied molecular

orbital (LUMO) is 2.2 eV considering a HOMO-LUMO gap of 2.6 eV as determined by

the 0-0 absorption band (469 nm). Cyclic voltammetry (CV) indicates that Ir(iprpmi)3

undergoes a reversible one-electron oxidation process with an onset of -0.15 V

(Appendix, Figure A2). The HOMO based on CV is 4.7 eV as calculated according to

E E Eonseto wherein EFC is the HOMO level of ferrocene (4.8 eV).

Figure 5.3 depicts the HOMO and LUMO obtained from density functional theory

(DFT) calculations. The HOMO is localized in the Ir d-orbital and the -orbital of the

phenyl moiety on the ligand whereas the LUMO is distributed on the phenyl-imidizole

ligand. These electron densities are similar to what has been reported for other

phenylimidazolinato Ir (III) complexes [27]. The calculated HOMO level is 4.3 eV.

Figure 5.3: Simulated frontier molecular orbitals of Ir(iprpmi)3. (a) HOMO electron density, (b)

ball and stick model, (c) LUMO electron density.

(a) (b) (c)

Page 133: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

105

5.3.2 Low EQE OLEDs

Ir(iprpmi)3 as a blue phosphorescent dopant in PHOLEDs was first reported in the

patent literature by Lin et al.[19]. They showed that Ir(iprpmi)3 and similar dopants with

an imidazole as the ligand generally provide a longer lifetime than other blue

phosphorescent dopants, such as lrpic. ollowing Lin’s work, we e amined in more

detail not only the effect of the dopant and host composition of the emitter layer, but also

the composition of the adjacent transport layers on the device performance. Two devices

were fabricated with a layer structure as follows: Anode/40 nm NPB/30 nm mCBP+x%

Ir(iprpmi)3/40 nm BAlq/Cathode. In this structure, which is similar to that used by Lin,

NPB and BAlq are the hole-transport and electron-transport layers, respectively, and

mCBP is the host matrix for the Ir(iprpmi)3 in the emitter layer. The concentration of

Ir(iprpmi)3 in mCBP is 6% and 12%. Figure 5.4 compares the electroluminescence

characteristics of these two devices. The EL spectra as shown in Figure 5.4(a) are

characteristic of Ir(iprpmi)3 emission and relatively independent of concentration. The

current-voltage curves of Figure 5.4(b) indicates a shift to lower drive voltage with

increasing Ir(iprpmi)3 concentration in the emitter layer, by about 1.3 V at 5 mA/cm2. As

shown in Figure 5.4(c), the EQE dependence on drive current density is different for the

two devices. Although both devices reach an EQE maximum of about 14%, they achieve

their maxima at different current densities. At 5 mA/cm2, the EQE is 8.4% for the device

with 6% Ir(iprpmi)3 compared to 13.7% for the device with 12% Ir(iprpmi)3. Similar to

what has been reported by Lin et al, the maximum EQE values (~14%) are substantially

lower than what can be expected from similar Ir-based dopants. The triplet energy levels

Page 134: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

106

Figure 5.4: OLED performance data for 6% and 12% concentrations of Ir(iprpmi)3 with NPB as

HTL and BAlq as ETL. Anode/40 nm TAPC/30 nm mCBP+ 6% or 12% Ir(iprpmi)3/40 nm

BAlq/Cathode. (a) EL spectra at 5 mA/cm2, (b) current density vs. voltage, (c) EQE vs. current

density.

of NPB and BAlq are 2.41eV and 2.18 eV, respectively, which are lower than that of

Ir(iprpmi)3 (2.66 eV). Thus both NPB and BAlq can cause quenching of the Ir(iprpmi)3

triplet excitons generated by electron-hole recombination in the emitter layer and the

extent of quenching will depend on the proximity of the recombination zone relative to

the interfaces with NPB and BAlq. The effect of shifting the recombination zone on

device performance is explored further in the following sections.

0.0

0.2

0.4

0.6

0.8

1.0

1.2

400 500 600 700R

ela

tiv

e I

nte

nsi

tyWavelength (nm)

6%

12%

0.1

1

10

100

0 4 8 12 16

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

6%

12%

0

4

8

12

16

0.1 1 10 100

EQ

E (

%)

Current Density (mA/cm2)

6%

12%

(a)

(b)(c)

Ir(iprpmi)3

Ir(iprpmi)3

Ir(iprpmi)3

Page 135: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

107

5.3.3 High EQE OLEDs

In the layer structure described above, the Ir(iprpmi)3 doped emitter layer was

sandwiched between the hole-transport layer NPB and the electron-transport layer BAlq.

To examine exciton quenching by the hole-transport layer, we fabricated devices similar

to those described above, except that NPB is substituted by TAPC for the hole-transport

layer. TAPC was chosen to eliminate the possibility of exciton quenching since the triplet

energy of TAPC (2.96 eV) is significantly higher than that of Ir(iprpmi)3 (2.66 eV). In

this series, the Ir(iprpmi)3 concentration range was broadened to cover from 0%

(undoped) to 24%.

The EL spectrum for the undoped device is shown in Figure 5.5(a). The spectrum

is invariant with current density and shows a broad peak centered at 484 nm, which can

be attributed to BAlq [31]. A weak shoulder appears around 400 nm, which can be due to

emissions from the host mCBP [32, 33], or the hole-transport layer TAPC [34], or both.

Qualitatively, the EL spectrum indicates that electron-hole recombination occurs

throughout the emitter layer and at both TAPC and BAlq interfaces adjacent to the

emitting layer. The relative intensities of the broad peak and the weak shoulder may

indicate that the location of the recombination zone is at or near the BAlq interface.

However, this is not necessarily the case as shown in experiments involving doped

devices where the recombination zone can be better located. In any case, the EQE of the

undoped device is very low compared with the Ir(iprpmi)3 doped devices.

The EL spectra for Ir(iprpmi)3 doped devices are also shown in Figure 5.5(a).

Regardless of the dopant concentration, the EL spectra are similar and characteristic of

Page 136: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

108

Figure 5.5: OLED performance data for Ir(iprpmi)3 concentration series with TAPC HTL and

BAlq ETL. Anode/40 nm TAPC/30 nm mCBP+ x% Ir(iprpmi)3/40 nm BAlq/Cathode. x=0%–24%.

(a) EL spectra at 5 mA/cm2, (b) current density vs. voltage, (c) EQE vs. current density.

Ir(iprpmi)3 emission. The corresponding CIEx,y coordinates are 0.17, 0.37.

The J-V plots in Figure 5.5(b) show that the drive voltage is dependent on the

Ir(iprpmi)3 dopant concentration in the emitting layer. At low concentrations (3% and

7%), the drive voltage (at 5 mA/cm2) is 10.9 V, which is as much as 2 volts higher than

the undoped device (9.0 V) at the same current density. With higher concentrations, the

drive voltage is reduced. For the device with 24% Ir(iprpmi)3, the drive voltage is about 1

volt less than the undoped cell. These J-V characteristics indicate that Ir(iprpmi)3 may act

as a charge trap or as a conductive channel. It will be shown that it is the hole-transport in

host mCBP that is primarily affected by Ir(iprpmi)3. As shown in Figure 5.5(c), the EQE

0

4

8

12

16

20

24

0.1 1 10 100 1000

EQ

E (

%)

Current Density (mA/cm2)

0%

3%

7%

13%

17%

24%

0.0

0.2

0.4

0.6

0.8

1.0

1.2

380 480 580 680

Rela

tiv

e I

nte

nsi

ty (a

.u.)

Wavelength (nm)

0%

3%

7%

13%

17%

24%

0.1

1

10

100

0 5 10 15

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

0%

3%

7%

13%

17%

24%

(c)

(a)

Ir(iprpmi)3

(b)

Ir(iprpmi)3 Ir(iprpmi)3

Page 137: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

109

is highly sensitive to the Ir(iprpmi)3 concentration in the emitter layer. At low

concentrations (3% and 7%), the EQE is high and relatively constant, ranging from 18 to

22% over a wide range of current densities from 0.1 to 20 mA/cm2 and decreasing only

slightly with increasing current density. At high concentrations (13%, 17%, and 24%),

the EQE is much reduced, but rises rapidly with increasing current density. For the device

with 13% Ir(iprpmi)3, the EQE increases from 8% at 0.1 mA/cm2 to a plateau of nearly

18% beyond 5 mA/cm2. For devices with 17% and 24% of Ir(iprpmi)3, the EQE is further

reduced and the fall-off at low current densities is much worse. The EQE at 0.1 mA/cm2

is only 0.8% and 1.7% for Ir(iprpmi)3 concentrations of 17% and 24%, respectively.

The fact that the devices with a lower Ir(iprpmi)3 dopant concentration (3% and

7%) exhibit a higher EQE and require a higher drive voltage indicates that electron-hole

recombination and subsequent generation and emission of Ir(iprpmi)3 excitons occurs

mainly near the interface with the hole-transport TAPC layer. Because BAlq is capable of

quenching Ir(iprpmi)3 triplet excitons, a lower EQE would have been observed if

electron-hole recombination was to occur at the BAlq interface. Ir(iprpmi)3 can affect

carrier transport in the emitter layer. Due to its HOMO (4.8 eV) and LUMO (2.2 eV)

energies, Ir(iprpmi)3 mainly affects the transport of holes in an mCBP matrix, but not

electrons. At low Ir(iprpmi)3 concentrations, holes can be effectively trapped by

Ir(iprpmi)3, causing recombination to be more localized at the TAPC interface. As a

result, the EQE is enhanced as observed in devices with 3% and 7% of Ir(iprpmi)3, since

TAPC, unlike BAlq, is incapable of quenching Ir(iprpmi)3 triplet excitons. As the

Ir(iprpmi)3 concentration increases, both the drive voltage and EQE are decreased. The

Page 138: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

110

voltage is decreased because Ir(iprpmi)3 molecules at a high concentration can form a

continuous phase to accept hole injection from TAPC and to assist the hole transport

through the mCBP host matrix. The decrease in EQE can be attributed to two reasons: 1)

Ir(iprpmi)3 triplet excitons can diffuse more easily through the Ir(iprpmi)3 phase towards

the BAlq interface, where they suffer quenching by BAlq; and 2) increased hole transport

through the emitter layer results in increased recombination and exciton quenching at the

BAlq interface. As will be shown in section 5.3.5, self-quenching and/or triplet-triplet

annihilation play only a minor role in reducing EQE in Ir(iprpmi)3 doped emitters. .

The EQE dependence on current density, Figure 5.5(c), may also be correlated to

the hole trapping and transport characteristics of Ir(iprpmi)3. For devices with a low

concentration of Ir(iprpmi)3 (such as 3% and 7%), the total current through the emitter

layer is presumably dominated by the electron current as holes are trapped by Ir(iprpmi)3

molecules and the recombination zone is effectively localized at TAPC interface. High

EQE’s are maintained in these devices, regardless of the current density, as long as the

electron current dominates. With increasing concentration of Ir(iprpmi)3 in the emitter

layer (7%–24%), the charge transport becomes more bi-polar as more holes are

transported through Ir(iprpmi)3 in addition to electron transport through mCBP. Since

there is no barrier for hole injection from TAPC to Ir(iprpmi)3 in contrast to a large

barrier (0.5 eV) for electron injection from BAlq into mCBP, it is likely that the hole

current is transport limited whereas the electron current is injection limited, leading to a

space distribution of electrons and holes across the emitter layer that is dependent on the

bias voltage. At a low bias voltage, the hole current dominates as the voltage is not high

Page 139: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

111

enough to cause electron injection from BAlq to mCBP, causing the recombination zone

to shift towards the BAlq interface, resulting in a decrease in EQE. With increasing bias,

more electron injection takes place, causing the recombination zone to shift away from

the BAlq interface towards the TAPC interface, resulting in increasing EQE. The extent

of this shift in the recombination zone is dependent on the Ir(iprpmi)3 concentration, with

the highest concentration resulting in the smallest shift away from the BAlq interface and

thus resulting in the lowest EQE.

5.3.4 Probing the recombination zone

In section 5.3.3 the dopant concentration for each device was fixed throughout the

entire mCBP layer. To gain a better understanding of where recombination and emission

occurs in the emitter layer, devices were designed and fabricated where the emitter layer

was split into two adjoining layers consisting of a layer of Ir(iprpmi)3 doped mCBP and a

layer of undoped mCBP. The general device structure is as follows: Anode/40 nm

TAPC/x nm mCBP+6% Ir(iprpmi)3/y nm mCBP/40 nm BAlq/Cathode, where x and y

refer to the thicknesses of the adjoined emitting layer. The Ir(iprpmi)3 was kept at 6%,

representing the low dopant concentration and optimal EQE derived from devices

discussed in section 5.3.3. The total thickness (x+y) of the doped (x) and undoped (y)

layers was kept at 30 nm with x and y varied from 5 to 25 nm. Each of these devices

exhibit identical EL spectra (not shown), similar to that of the device with 7% Ir(iprpmi)3

shown in Figure 5.5(a), indicating that emission is from Ir(iprpmi)3 and the

recombination zone is adjacent to the TAPC interface. Figure 5.6(a) and Figure 5.6(b)

show the device J-V and EQE-J data respectively. Essentially there are only minor

Page 140: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

112

differences between the devices and their performance is similar to the 7% Ir(iprpmi)3

device in section 5.3.3; there is little difference in drive voltage and the EQE varied from

15 to 20% among the devices with a mild dependence on current density over a current

density range from 0.1 to 20 mA/cm2. This indicates that the recombination zone is

located within the first 5 nm of the TAPC interface. Furthermore, this also provides

evidence that triplet Ir(iprpmi)3 excitons are unable to diffuse to the BAlq interface and

explains why BAlq can be utilized as the ETL without affecting the device efficiency as

long as the Ir(iprpmi)3 dopant concentration in mCBP is relatively low (~6%).

Figure 5.6: OLED performance data when probing the recombination zone with 6% Ir(iprpmi)3.

Anode/40 nm TAPC/x nm mCBP+6% Ir(iprpmi)3/y nm mCBP/40 nm BAlq/Cathode. [x=25 nm,

y=5 nm], [x=20 nm, y=10 nm],[x=10 nm, y=20 nm], and [x=5 nm, y=25 nm]: (a) current density

vs. voltage, (b) EQE vs. current density.

To further illustrate the effect of Ir(iprpmi)3 on the recombination zone shift in the

emitter layer, devices were again fabricated with the emitter layer being split into two

adjoining layers consisting of a layer of Ir(iprpmi)3 doped mCBP and a layer of undoped

mCBP, except that the Ir(iprpmi)3 dopant concentration in the emitter layer was increased

to 25%. The EL spectra (not shown) for these devices are identical and are the same as

0.1

1

10

100

0 5 10 15

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

5 nm

10 nm

20 nm

25 nm

0

4

8

12

16

20

24

0.1 1 10 100

EQ

E (

%)

Current Density (mA/cm2)

5 nm

10 nm

20 nm

25 nm(a) (b)

Undoped mCBP

Undoped mCBP

Page 141: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

113

what was observed for the 6% doped devices, indicating the emission is from Ir(iprpmi)3.

Also similar to the 6% doped devices, the EQE, Figure 5.7(b), for these 25% doped

devices is uniformly high (15-18%) and relatively independent of the current density.

This is in direct contrast with the devices having an emitter configuration where the

highly doped layer is in direct contact with the BAlq electron-transport layer. As shown

in the device where a 25% Ir(iprpmi) doped layer is in contact with BAlq (section 5.3.3),

the EQE is low (< 7%) and highly dependent on current density. Evidently, it is critical to

Figure 5.7: OLED performance data when probing the recombination zone with 25% Ir(iprpmi)3.

Anode/40 nm TAPC/x nm mCBP+25% Ir(iprpmi)3/y nm mCBP/40 nm BAlq/Cathode. [x=25 nm,

y=5 nm],[x=20 nm, y=10 nm], [x=10 nm, y=20 nm], and [x=5 nm, y=25 nm]: (a) current density

vs. voltage [ inset: voltage vs mCBP layer thickness of devices under current densities of 1, 5, and

10 mA/cm2] (b) EQE vs. current density.

separate the doped mCBP layer from the BAlq interface at high Ir(iprpmi)3

concentrations in order to avoid exciton quenching by the BAlq and achieve high EQE

values. However, the presence of an undoped mCBP layer can significantly increase the

drive voltage depending on the layer thickness. As shown in Figure 5.7(a), the increase in

0.1

1

10

100

1000

0 5 10 15

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

5 nm

10 nm

20 nm

25 nm

0

4

8

12

16

20

24

0.1 1 10 100

EQ

E (

%)

Current Density (mA/cm2)

5 nm

10 nm

20 nm

25 nm(a) (b)

Undoped mCBP

Undoped mCBP

6

8

10

12

0 10 20 30 40

Undoped mCBP (nm)

J (mA/cm2)

10

5

1

Voltage

Page 142: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

114

voltage can be as large as 2 volts with a 25 nm undoped mCBP layer, which apparently

blocks hole injection from the mCBP doped layer. The inset of Figure 5.7(a) shows a

linear relationship between device voltage and the undoped mCBP layer thickness,

indicating the electric field across undoped mCBP is constant.These 25% doped devices

all have high EQE’s because the undoped mCBP layer prevents recombination from

occurring at the EML/BAlq interface while also preventing triplet excitons from diffusing

to this interface. There are noteworthy trends among the EQE-J curves displayed in

Figure 5.7(b). Depending on current density, the EQE ranges from 13-18% (device with

y=5 nm), 16-20% (devices with y=10 nm and y=20 nm), and 13-17% (device with y=25

nm). Devices with y=10 nm and y=20 nm have the highest EQE’s. Apparently an

undoped mCBP layer of 10-20 nm thick is needed to completely eliminate the influence

of BAlq. The device with only 5 nm of undoped mCBP has a lower efficiency

presumably from BAlq quenching. The device with 25 nm of undoped mCBP is

sufficiently thick to prevent BAlq quenching and yet the EQE is somewhat lower than

devices with 10 nm and 20 nm thick undoped mCBP. A possible explanation is that the

device with 25 nm of undoped mCBP has a thin doped layer which, being only 5 nm

thick, can lead to a higher density of triplet excitons and consequently a higher rate of

triplet-triplet annihilation and/or exciton-polaron interactions.

5.3.5 High EQE, low voltage OLEDs

Of all the devices described so far, those in section 5.3.3 with 3% and 7% of

Ir(iprpmi)3 have the highest EQE’s but also relatively high drive voltages due to

primarily the use of BAlq as the electron-transport layer, which has a low electron

Page 143: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

115

mobility of 10-5

cm2/Vs [35]. By replacing BAlq with TmPyPB, a known electron-

transport material of high mobility (10-3

cm2/Vs) [36], we were able to demonstrate both

high EQE and low voltage in devices that had the following structure: Anode/40 nm

TAPC/30 nm mCBP+x% Ir(iprpmi)3/40 nm TmPyPB/Cathode. The Ir(iprpmi)3

concentration ranged from 3%–24%.

Figure 5.8: OLED performance data for Ir(iprpmi)3 concentration series with TmPyPB ETL.

Anode/40 nm TAPC/30 nm mCBP + x% Ir(iprpmi)3/40 nm TmPyPB/Cathode. x=3%–24% (a)

current density vs. voltage, (b) EQE vs. current density.

The J-V characteristics are shown in Figure 5.8(a) and indicate a clear reduction

in drive voltage with TmPyPB as the ETL. Furthermore, similar to what was observed for

devices in section 5.3.3 with BAlq as the ETL, the drive voltage decreases as the dopant

concentration increases. A reduction of approximately 1 V is realized by increasing the

Ir(iprpmi)3 concentration from 3% to 24%. As shown in Figure 5.8(b), high EQE’s were

obtained for all devices, ranging from 23% at low current densities to 15% at high current

densities. The efficiency remains high regardless of dopant concentration because

TmPyPB has a triplet energy of 2.78 eV and is unable to quench the Ir(iprpmi)3 triplet

0.1

1

10

100

0 2 4 6 8

Cu

rren

t D

en

sity

(m

A/c

m2)

Voltage (V)

3%

6%

13%

18%

24%

0

5

10

15

20

25

0.1 1 10 100

EQ

E (

%)

Current Density (mA/cm2)

3%

6%

13%

18%

24%

Ir(iprpmi)3

(a) (b)

Ir(iprpmi)3

Page 144: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

116

excitons. This also indicates that self-quenching and/or triplet-triplet annihilation play

only a minor role in reducing the EQE.

5.3.6 Device lifetime

Lifetime testing was conducted on several representative Ir(iprpmi)3 doped

devices containing various hole and electron transport layers. For two devices, NPB was

used as the hole-transport layer, BAlq was used as the electron-tranport layer, and the

Ir(iprpmi)3 concentration was 6% and 12%. Two additional devices had the same device

structure except that TAPC replaced NPB as the hole-transport layer. For the last device,

TAPC and TmPyPB were the hole-transport layer and electron-transport layer,

respectively, while the Ir(iprpmi)3 concentration was 6%. It can be seen from Figure 5.9

Figure 5.9: Lifetime testing at 5 mA/cm2 for device structure, Anode/40 nm HTL/30 nm

mCBP+x% Ir(iprpmi)3/40 nm ETL/Cathode, [NPB-6% dopant-BAlq], [NPB-12% dopant-BAlq],

[TAPC-7% dopant-BAlq], [TAPC-13% dopant-BAlq], [TAPC-6% dopant-TmPyPB].

0.5

0.6

0.7

0.8

0.9

1

1.1

0 20 40 60 80 100

No

rmali

zed

Lu

min

ance

Time (hours)

NPB, 6% dopant, BAlq

NPB, 12% dopant, BAlq

TAPC, 7% dopant, BAlq

TAPC, 13% dopant, BAlq

0.5

0.6

0.7

0.8

0.9

1

0 1 2

No

rmal

ized

Lu

min

ance

Time (hours)

Page 145: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

117

that the half-life (at a constant current of 5 mA/cm2), while quite short for all devices –

less than 100 hours, is rather dependent on the transport layers. Among the devices, the

device with TAPC and TmPyPB as the hole and electron transport layers, respectively, is

the most short-lived with a half-life of less than ½ hour (see Figure 5.9 insert), whereas

devices with NPB and BAlq replacing TAPC and TmPyPb, respectively, are the most

stable with a half-life in the range of 50-80 hours. Devices with TAPC as the hole-

transport layer and BAlq as the electron-transport layer have an intermediate half-life in

the range of 2-10 hours, respectively. The variation in lifetimes with the transport layers

indicates that the instability in these phosphorescent blue devices may not necessarily be

due to the emitter layer. It is known that fluorescent OLED devices are more stable with

NPB instead of TAPC as the hole-transport layer [50, 51]. As an electron-transport

material, BAlq is known to improve device stability [37]. The better stability of A1 and

A2 are in large part due to the use of NPB and BAlq as the respective hole and electron

transport layers, suggesting that recombination in these layers or at their interfaces with

the emitter layer, be it a fluorescent or phosphorescent emitter, are most detrimental to the

operational stability. Comparing the devices with NPB as the HTL, the device with 6%

Ir(iprpmi)3 has a longer half-life (80 vs. 50 hours) but a lower EQE (7% vs. 13%) than

the device with 12% Ir(iprpmi)3. This correlation of higher stability with lower efficiency

also applies to the devices with TAPC as the HTL, implying that the formation of triplet

excitons invariably also leads to device degradation.

The observed lifetimes for the devices with NPB as the HTL are inferior to what

was reported by Lin et al. [19] where lifetimes reached ~400-700 hrs. using Ir(iprpmi)3 as

Page 146: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

118

the emitter. These disparities in lifetime can be attributed to different device architectures

and initial EQE’s.

The lifetime data presented here show that Ir(iprpmi)3-based devices with suitable

transport layers have longer lifetimes compared to FIrpic-based devices. In Flrpic-based

devices, the half-life is on the order of an hour or less regardless of the transport layer,

which has been attributed to the electrochemical degradation of Flrpic molecules related

to its fluorine substituents and the picolinate ligand. These structural features are not

present in Ir(iprpmi)3. Since recombination and emission can be engineered to take place

largely at the hole-transport interface using low-concentration phosphorescent dopants as

hole traps, the ETL material has no significant impact on device lifetime. Thus device

stability is more hinged on the selection of hole-transport materials as well as the host

materials for the emitter, assuming that the Ir-based dopant is relatively stable. We have

shown in this study how important a role the HTL material plays for both device EQE

and lifetime. Since amines and carbazoles are known to undergo homolytic C-N bond

cleavage in the excited state [38-40], it is likely that these common classes of hole-

transport materials are prompt to degrade in blue phosphorescent emitters because of

possible recombination in the hole-transport layer or interface resulting in excitons with

energy exceeding the dissociation energy of the C-N bonds. It follows that none of the

HTL materials used in this study are practical for achieving lifetimes that match those of

green and red phosphorescent devices, where the low-energy dopants can suppress

recombination in both hole and electron transport layers. To significantly improve the

stability of blue phosphorescent devices, new hole-transport materials will be needed.

Page 147: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

119

5.4 CONCLUSIONS

Ir(iprpmi)3 as a dopant in blue phosphorescent emitters has been investigated with

mCBP as the host and various materials as the hole and electron transport layers. We

have shown that Ir(iprpmi)3 as a dopant in host mCBP is capable of trapping (at low

concentration) and transporting (at high concentration) holes. Utilizing Ir(iprpmi)3 to trap

holes, the recombination zone can be confined at or near the hole-transport layer. As a

result, EQE values obtained are highly dependent on the HTL used. With TAPC as HTL,

which does not quench Ir(iprpmi)3 triplet e citons, EQE’s over 20% have been achieved.

Utilizing Ir(iprpmi)3 for hole transport, the recombination zone can be shifted towards the

ETL. With low-triplet BAlq as the ETL, much reduced EQE’s are obtained that vary not

only with dopant concentration but also electric field, indicating various extent of exciton

quenching by BAlq. Quenching by the ETL can be eliminated and high EQE’s retained

by inserting a thin undoped mCBP layer between the ETL and EML. We show that

device lifetime using Ir(iprpmi)3 as the emitting dopant is significantly improved

compared to when FIrpic is the emitting dopant. Furthermore, device lifetime is highly

dependent on the choice of host and hole/electron transport layers.

Page 148: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

120

REFERENCES

[1] M. Segal, C. Mulder, K. Celebi, M. Singh, K. Rivoire, S. Difley, T. Van Voorhis,

M.A. Baldo, Proc. SPIE, 6999 (2008) 699912.

[2] K.S. Yook, J.Y. Lee, Adv. Mater., 24 (2012) 3169.

[3] N. Chopra, J. Lee, Y. Zheng, S.-H. Eom, J. Xue, F. So, Appl. Phys. Lett., 93

(2008) 143307.

[4] H. Sasabe, E. Gonmori, T. Chiba, Y.-J. Li, D. Tanaka, S.-J. Su, T. Takeda, Y.-J. Pu,

K.-i. Nakayama, J. Kido, Chem. Mater., 20 (2008) 5951.

[5] S.-J. Su, E. Gonmori, H. Sasabe, J. Kido, Adv. Mater., 20 (2008) 4189.

[6] J. Lee, J.-I. Lee, J.Y. Lee, H.Y. Chu, Appl. Phys. Lett., 94 (2009) 193305.

[7] L. Xiao, S.-J. Su, Y. Agata, H. Lan, J. Kido, Adv. Mater., 21 (2009) 1271.

[8] C.W. Lee, J.Y. Lee, Adv. Mater., 25 (2013) 1.

[9] W. Sotoyama, T. Satoh, M. Kinoshita, M. Tobise, K. Kawato, T. Ise, H. Takizawa,

S. Yamashita, Dig. Tech. Pap. - Soc. Inf. Disp. Int. Symp., 41 (2010) 556.

[10] H. Lee, H. Ahn, C. Lee, J. Inf. Disp., 12 (2011) 219.

[11] I. Rabelo de Moraes, S. Scholz, B. Luessem, K. Leo, Org. Electron., 12 (2011)

341.

[12] K.S. Yook, O.Y. Kim, J.Y. Lee, Synth. Met., 161 (2011) 2677.

[13] V. Sivasubramaniam, F. Brodkorb, S. Hanning, H.P. Loebl, V. van Elsbergen, H.

Boerner, U. Scherf, M. Kreyenschmidt, J. Fluorine Chem., 130 (2009) 640.

[14] E. Baranoff, B.F.E. Curchod, J. Frey, R. Scopelliti, F. Kessler, I. Tavernelli, U.

Rothlisberger, M. Gratzel, M.K. Nazeeruddin, Inorg. Chem., 51 (2012) 215.

Page 149: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

121

[15] K.P. Klubek, C.W. Tang, L.J. Rothberg, Org. Electron., 15 (2014) 1312.

[16] N.C. Giebink, B.W. D'Andrade, M.S. Weaver, P.B. Mackenzie, J.J. Brown, M.E.

Thompson, S.R. Forrest, J. Appl. Phys., 103 (2008) 044509.

[17] H. Yamamoto, J. Brooks, M.S. Weaver, J.J. Brown, T. Murakami, H. Murata,

Appl. Phys. Lett., 99 (2011) 033301.

[18] J. Zhuang, W. Li, W. Su, Y. Liu, Q. Shen, L. Liao, M. Zhou, Org. Electron., 14

(2013) 2596.

[19] C. Lin, P.B. MacKenzie, R.W. Walters, J.-Y. Tsai, C.S. Brown, J. Deng, U.S.

Patent No. 7,902,374 (2011).

[20] D.B. Knowles, C. Lin, P.B. Mackenzie, J.-Y. Tsai, R.W. Walters, S.A. Beers, C.S.

Brown, W.H. Yeager, U.S. Patent No. 7,915,415 (2011).

[21] V. Adamovich, M.S. Weaver, B. D'Andrade, U.S. Patent No. 8,580,402 (2013).

[22] B. D'Andrade, P.B. MacKenzie, M.S. Weaver, J.J. Brown, U.S. Patent No.

8,557,399 (2013).

[23] A.B. Tamayo, B.D. Alleyne, P.I. Djurovich, S. Lamansky, I. Tsyba, N.N. Ho, R.

Bau, M.E. Thompson, J. Am. Chem. Soc., 125 (2003) 7377.

[24] M.J.T. Frisch, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman,

J. R.; Montgomery, Jr., J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J.

M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.;

Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.;

Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai,

H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.;

Page 150: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

122

Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A.

J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G.

A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A.

D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.;

Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.;

Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.;

Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.;

Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez,

C.; and Pople, J. A., Gaussian 03, Revision C.02, Gaussian, Inc., Wallingford CT,

2004.

[25] T. Fei, X. Gu, M. Zhang, C. Wang, M. Hanif, H. Zhang, Y. Ma, Synth. Met., 159

(2009) 113.

[26] S.J. Lee, K.-M. Park, K. Yang, Y. Kang, Inorg. Chem., 48 (2009) 1030.

[27] T. Karatsu, M. Takahashi, S. Yagai, A. Kitamura, Inorg. Chem., 52 (2013) 12338.

[28] J.J. Brooks, R.C. Kwong, Y.-J. Tung, M.S. Weaver, B.W. D'Andrade, V.

Adamovich, M.E. Thompson, S.R. Forrest, J.J. Brown, Proc. SPIE, 5519 (2004)

35.

[29] S.-J. Yeh, M.-F. Wu, C.-T. Chen, Y.-H. Song, Y. Chi, M.-H. Ho, S.-F. Hsu, C.H.

Chen, Adv. Mater., 17 (2005) 285.

[30] F.-C. Fang, Y.-L. Liao, K.-T. Wong, M.-H. Tsai, H.-W. Lin, H.-C. Su, C.-c. Wu,

C.-I. Wu, Proc. SPIE-Int. Soc. Opt. Eng., 6192 (2006) 61921R.

[31] J.-h. Niu, W.-l. Li, H.-z. Wei, M.-t. Li, W.-m. Su, Q. Xin, Z.-q. Zhang, Z.-z. Hu, J.

Page 151: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

123

Phys. D Appl. Phys., 38 (2005) 1136.

[32] P. Schroegel, N. Langer, C. Schildknecht, G. Wagenblast, C. Lennartz, P.

Strohriegl, Org. Electron., 12 (2011) 2047.

[33] S. Gong, X. He, Y. Chen, Z. Jiang, C. Zhong, D. Ma, J. Qin, C. Yang, J. Mater.

Chem., 22 (2012) 2894.

[34] K.B. Kahen, J.R. Vargas, D.Y. Kondakov, C.T. Brown, L. Cosimbescu, V. Jarikov,

U.S. Patent No. 6,876,684 (2005).

[35] J.H. Seo, K.H. Lee, B.M. Seo, J.R. Koo, S.J. Moon, J.K. Park, S.S. Yoon, Y.K.

Kim, Org. Electron., 11 (2010) 1605.

[36] S.-J. Su, T. Chiba, T. Takeda, J. Kido, Adv. Mater., 20 (2008) 2125.

[37] T. Watanabe, K. Nakamura, S. Kawami, Y. Fukuda, T. Tsuji, T. Wakimoto, S.

Miyaguchi, Proc. SPIE-Int. Soc. Opt. Eng., 4105 (2001) 175.

[38] D.Y. Kondakov, W.C. Lenhart, W.F. Nichols, J. Appl. Phys., 101 (2007) 024512.

[39] D.Y. Kondakov, J. Appl. Phys., 104 (2008) 084520.

[40] D.Y. Kondakov, Proc. SPIE, 7051 (2008) 70510I.

Page 152: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

124

CHAPTER 6

SUMMARY AND FUTURE RESEARCH

6.1 SUMMARY

Even as OLEDs are currently achieving commercial success, significant research

effort is focused on improving device performance metrics such as efficiency, voltage and

color. The use of iridium (Ir) cyclometalated phosphorescent emitters has been shown to

produce the highest efficiency in OLED devices. In fact, red and green Ir-based emitters

have already been adopted into consumer products due to their ability to produce devices

that have a combination of high efficiency and excellent lifetime. While high efficiency

has also been demonstrated for PHOLEDs using blue Ir-based emitters, the device

lifetimes lag far behind those of the red and green. This thesis focused on blue PHOLED

instability mechanisms from a materials and device perspective. In particular, two key

blue emitting iridium compounds were extensively studied, bis(4,6-difluorophenyl-

pyridinato-N,C2) picolinate iridium (FIrpic) and tris[1-(2,6-diisopropylphenyl)-2-phenyl-

1H-imidazole]iridium(III) (Ir(iprpmi)3).

FIrpic is the most commonly employed blue emitting phosphorescent dopant and

is often used to demonstrate high efficiencies for researchers developing new host and

transport materials. However device lifetime is typically not reported. In Chapter 3,

FIrpic was studied as a light-emitting dopant using both single and mixed host materials.

Devices incorporating high triplet energy host and transport materials yielded devices

Page 153: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

125

with EQE’s over 15%, however lifetimes for these devices were all very short (less than

20 minutes at 5 mA/cm2). Switching to charge transport materials with triplet energies

lower than FIrpic reduced the EQE by 50% or more, however the lifetime was improved.

A mixed host LEL (TCTA:UGH3) in combination with the low triplet energy BAlq as the

ETL was found to provide the best combination for EQE (~9%) and lifetime (~12 hours).

While this lifetime is significantly improved compared to devices having the highest

EQE’s, it is ultimately concluded that achieving stable blue phosphorescent devices for

commercial applications is not possible when using FIrpic as the light-emitting dopant.

It was also found that the supposedly electron-transporting co-host UGH3 is in

fact a poor electron transporting material. Its role within the LEL is determined to be that

of a spacer group, serving to slow down hole transport and minimizing exciplex

formation between the hole transport host and the adjacent electron transport material. It

was also found that emission occurring from the host or transport materials is detrimental

to device lifetime. The host mCBP was used in device structures that have been reported

in the patent literature to achieve blue PHOLED device lifetime of 10,000 hours.

However, we found that mCBP as a host for FIrpic does not improve device lifetimes

over those of single host TCTA or mixed host TCTA:UGH3 devices.

In Chapter 4, hole transport in FIrpic doped mCBP and the effect on the device

operational voltage and stability were investigated. The host mCBP was found to be

stable in supporting hole transport but unstable with respect to electron-hole

recombination. As a dopant, FIrpic was found to be unstable with respect to both hole

transport and charge recombination processes. The results from Chapter 3 and Chapter 4

Page 154: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

126

provide clear evidence that FIrpic is unsuitable to use for generating light and for

conducting hole transport within OLED devices.

In Chapter 5, the blue phosphorescent dopant Ir(iprpmi)3 was investigated.

Ir(iprpmi)3 lacks specific structural features present within FIrpic, namely fluorine

substituents and the ancillary picolinate ligand, that have been attributed to molecular

degradation. The photophysical properties of this material were reported. PHOLED

devices using Ir(iprpmi)3 as the light-emitting dopant and mCBP as the host have been

evaluated. By optimizing the dopant concentration and the materials for the electron and

hole-transport layers, external quantum efficiencies greater than 20% have been achieved.

Device lifetimes are significantly improved compared to FIrpic-based devices, with

lifetimes approaching 100 hours having been achieved. In addition to the differences in

molecular design, these improvements are attributed to the control of the electron-hole

recombination and emission regions within the emitter layer as well as the choice of

material for the transport layers.

As discussed in Chapters 3 and 5, high triplet energy transport materials are

necessary to confine triplet excitons within the light-emitting layer for devices. TmPyPB,

with a high ET of 2.8 eV, was used as ETL. Presumably, triplet excitons formed on FIrpic

(ET = 2.6 eV) or Ir(iprpmi)3 (ET = 2.66 eV) within the LEL near the LEL/ETL interface

will not be quenched by TmPyPB. However, it is possible that non-thermalized FIrpic or

Ir(iprpmi)3 triplet excitons with a triplet energy higher than 2.8 eV could produce

TmPyPB triplet excitons through energy transfer. Based on a Boltzmann distribution,

there is also a finite probability for the formation of TmPyPB excited states at the

Page 155: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

127

LEL/ETL interface. Regardless of how it is formed, the TmPyPB excited state has high

enough energy to promote bond breaking, resulting in the formation of quencher species

capable of impacting device lifetime. This concept can be further extended to include the

host and hole transport materials. The creation of high energy excited states in these

materials may also contribute significantly to the poor lifetimes that we observed for the

highest EQE devices.

As discussed in Chapter 2, the KOMET coater is a bench-top thermal evaporation

system that we designed and built for the fabrication of all OLED devices described in

this thesis. The key feature for this system is the compact thermal source design utilizing

multiple, low power consumption tube boats for depositing organic materials. In addition

to low-power consumption, these boats are highly efficient with respect to material

utilization. The KOMET coater produced OLED devices with excellent reproducibility.

With a compact design, the KOMET coater (<$40,000) is more affordable compared to

traditional evaporation coaters (>$100,000).

6.2 FUTURE RESEARCH

Based on the results presented in this thesis, it is clear that FIrpic should be

avoided as the quest to improve blue PHOLED device performance continues.

Researchers should instead opt for alternative phosphorescent blue dopants such as

Ir(iprpmi)3. While we have shown that using this dopant has produced devices having

significantly improved lifetime compared to FIrpic, the lifetimes we have achieved (~100

hours) are still far too low for commercial applications. Using Ir(iprpmi)3 as the emitter in

a different device architecture, Lin et al. [1] showed the device lifetime can approach

Page 156: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

128

400-700 hours. Additional studies need to be conducted with Ir(iprpmi)3 to investigate

whether or not this material is intrinsically unstable or if device stability is instead limited

by the choice of charge transport and host materials. Laser desorption ionization mass

spectrometry (LDI-MS) has been utilized to determine that FIrpic degradation results in

the loss of the picolinate ligand [2]. LDI-MS can similarly be used to investigate the

molecular stability of Ir(iprpmi)3 upon device degradation. Wtih Ir-based phosphorescent

devices generally using relatively large concentrations (6-20%) of the dopant, it should

also be easier to look for MS signals related to these types of materials compared to their

fluorescent counterparts that typically use dopant concentrations of 1-3%. An alternative

known technique [3-5] may be to extract the materials after degradation using solvents

and then use liquid chromatography mass spectrometry (LC-MS) in an attempt to identify

the degradation products related to Ir(iprpmi)3 or derivatives of imidazole-based Ir-

cyclometalated complexes. In Chapter 3 we found that FIrpic was unstable to hole

transport. Since Ir(iprpmi)3 has a shallow HOMO and effectively traps or transports holes

depending on concentration, additional studies are needed to determine how stable this

dopant is to charge transport.

While using Ir(iprpmi)3 is an excellent alternative to FIrpic for evaluating blue

PHOLED devices due to its structural differences and commercial availability, alternative

dopants should also be studied (Figure 6.1). As discussed in Chapter 3, imidazole

derivative Ir(tdmp)3, which is not commercially available, has been reportedly used as a

Page 157: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

129

Figure 6.1: Different types of blue phosphorescent dopants.

dopant to achieve a lifetime as high as 10,000 hours [6] in a device structure similar to

those reported by Lin et al. Ir(tdmp)3 has a more rigid structure compared to Ir(iprpmi)3

and also has a bulky dimethyl-phenyl substituent attached to the double bond of the

imidazole ring. It is possible that by making the molecule more rigid and bulky, it is less

susceptible to degradation simply because potentially reactive moieties such as the

nitrogen atoms or the double bond are more shielded from neighboring molecules.

Carbene complexes [7-10] are another class of blue emitting material that are showing

promise and can be compared with the imidazole derivatives. In any case, PHOLED

device and analytical (degradation) studies need to be conducted on these different

classes of dopants and their derivatives to compare and contrast how the differences in

molecular structure affects device performance, especially lifetime.

In addition to studying new blue phosphorescent dopants, there is also a need to

study new charge transport and host materials. In this thesis, triarylamine and carbazole

derivatives such as NPB, TAPC, TCTA, and mCBP have been used as HTL and host

materials. Since these types of materials are known to undergo C-N bond cleavage in

Blue Phosphorescent Dopants

FIrpic Ir(iprpmi)3Ir(tdmp)3 Carbene complex

Page 158: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

130

OLED devices, alternative materials that are either free of C-N bonds or that are

sterically hindered so that the C-N bond is less prone to cleavage are needed. From a

device engineering perspective, preventing electron-hole recombination from occurring

on the hole transport materials will also limit or prevent these types of materials from

degrading. For example, the HOMO level of Ir(iprpmi)3 is very shallow and therefore this

material is a very strong hole trap. This type of material may be able to prevent

recombination from occurring on any other hole transport materials. However, the

electron transport through the host also needs to be considered. With the LUMO level of

Ir(iprpmi)3 being rather high, electron transport occurs preferentially on the host material.

This requires the host to be stable to electron transport. We used mCBP for transporting

electrons, additional studies need to be conducted to determine if this material is stable

for this role. Electron-only devices can be made and the voltage rise with time will help

to determine the charge transport stability of mCBP. LDI-MS or LC-MS can then be used

in an attempt to identify degradation products. It is also important to use host materials

that are stable in the excited state. In Chapter 4 we found that forming the mCBP excited

state in a model OLED with Alq as the emitter resulted in enhanced device degradation.

For devices using mCBP as the host for Ir(iprpmi)3, further study is required to assess

whether mCBP is entering the excited state and factoring into the short device lifetimes.

Depending on the type of phosphorescent blue dopant and host materials, the

electron transport material can also play a critical role with respect to device

performance. If recombination occurs at the LEL/ETL interface, then high triplet energy

ETL materials are typically used to prevent exciton quenching. In order to have both high

Page 159: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

131

triplet energy and high electron mobility, ETL materials typically utilize pyridine-like

nitrogen atoms within the molecular structure. Unfortunately, these types of materials

often contribute to device instability. If recombination can be engineered to occur away

from the ETL, then degradation related to the ETL material may be avoided. However, a

host material that is stable to electron transport is required.

Since iridium cyclometalated complexes have been shown to achieve EQE’s

greater than 20%, there is often a desire to maintain these high efficiencies while also

utilizing device structures designed to enhance the lifetime. As we showed in Chapter 5,

often the correlation between efficiency and lifetime is unfortunately inverse – decreased

lifetimes for higher efficiency devices. In order to improve the device lifetime, one

strategy is to move away from Ir complexes with a high triplet energy that would produce

deeper blue emissions to Ir complexes with a lower triplet energy that would produce

lighter blue emissions. Although the color of the blue emission is somewhat

compromised, the lifetimes of the light blue PHOLEDs are often significantly higher than

the deep blue PHOLEDs. Engineering stable light-blue and high-efficiency PHOLEDs

may be a more expedient path to practical applications. Furthermore, deeper blue

emissions can be obtained with these light blue emitters by using device structures having

internal microcavity or external color filters. By easing the triplet energy requirements

and allowing for lower EQE blue devices, many new phosphorescent and host materials

are open to investigation and the prospects for achieving practical PHOLEDs should be

much higher.

Page 160: Investigation of Blue Phosphorescent Organic Light-Emitting Diode Instability

132

REFERENCES

[1] C. Lin, P.B. MacKenzie, R.W. Walters, J.-Y. Tsai, C.S. Brown, J. Deng, U.S.

Patent No. 7,902,374 (2011).

[2] I. Rabelo de Moraes, S. Scholz, B. Luessem, K. Leo, Org. Electron., 12 (2011)

341.

[3] D.Y. Kondakov, J. Appl. Phys., 104 (2008) 084520.

[4] D.Y. Kondakov, Proc. SPIE, 7051 (2008) 70510I.

[5] D.Y. Kondakov, T.D. Pawlik, W.F. Nichols, W.C. Lenhart, J. Soc. Inf. Disp., 16

(2008) 37.

[6] B. D'Andrade, P.B. MacKenzie, M.S. Weaver, J.J. Brown, U.S. Patent No.

8,557,399 (2013).

[7] P. Erk, M. Bold, M. Egen, E. Fuchs, T. Gessner, K. Kahle, C. Lennartz, O. Molt,

S. Nord, H. Reichelt, C. Schildknecht, H.-H. Johannes, W. Kowalsky, Dig. Tech.

Pap. - Soc. Inf. Disp. Int. Symp., 37 (2006) 131.

[8] C.-F. Chang, Y.-M. Cheng, Y. Chi, Y.-C. Chiu, C.-C. Lin, G.-H. Lee, P.-T. Chou,

C.-C. Chen, C.-H. Chang, C.-C. Wu, Angew. Chem., Int. Ed., 47 (2008) 4542.

[9] H. Sasabe, J.-i. Takamatsu, T. Motoyama, S. Watanabe, G. Wagenblast, N. Langer,

O. Molt, E. Fuchs, C. Lennartz, J. Kido, Adv. Mater. (Weinheim, Ger.), 22 (2010)

5003.

[10] C.-H. Hsieh, F.-I. Wu, C.-H. Fan, M.-J. Huang, K.-Y. Lu, P.-Y. Chou, Y.-H.O.

Yang, S.-H. Wu, I.C. Chen, S.-H. Chou, K.-T. Wong, C.-H. Cheng, Chem. - Eur.

J., 17 (2011) 9180.