Sie sind auf Seite 1von 7

Organic Electronics 14 (2013) 12041210

Contents lists available at SciVerse ScienceDirect

Organic Electronics
journal homepage: www.elsevier.com/locate/orgel

Solution-processed hexaazatriphenylene hexacarbonitrile as a universal hole-injection layer for organic light-emitting diodes
Hao-Wu Lin a,, Wei-Chieh Lin a, Jung-Hung Chang b, Chih-I Wu b
a b

Department of Materials Science and Engineering, National Tsing Hua University, HsinChu 300, Taiwan, ROC Department of Electrical Engineering and Graduate Institute of Electro-optical Engineering, National Taiwan University, Taipei 106, Taiwan, ROC

a r t i c l e

i n f o

a b s t r a c t
The use of 1,4,5,8,9,11-hexaazatriphenylene hexacarbonitrile (HAT-CN) thin layers, particularly the solution-processed type, as an efcient hole-injection layer (HIL) for organic optoelectronic devices is demonstrated herein. Among the solvents commonly used for solution processing, 2-propanone was found to selectively dissolve HAT-CN, allowing the fabrication of a rigid lm. The alignment of the electronic energy levels of the solution-processed HAT-CN and thermally polymerized 2,7-disubstituted uorene-based triaryldiamine (VB-FNPD) species was evaluated using ultraviolet photoelectron spectroscopy. The results revealed that the lowest unoccupied molecular orbital of HAT-CN and the highest occupied molecular orbital of VB-FNPD were very close to the Fermi level, which facilitated charge transfer at the interface and improved hole injection. The utilization of HATCN as HIL resulted in a dramatic enhancement of the performance of solution-processed red, green, and blue organic light-emitting diodes. The external quantum efciency, current efciency, and power efciency of the HAT-CN-based devices were higher than or almost similar to those of optimized poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS)-based devices. Because of the efcient carrier-injection capability and the capacity to prevent interfacial mixing and erosion during fabrication, solution-processed HAT-CN is promising as a novel alternative to conventional PEDOT:PSS HILs. 2013 Elsevier B.V. All rights reserved.

Article history: Received 24 December 2012 Received in revised form 6 February 2013 Accepted 13 February 2013 Available online 27 February 2013 Keywords: Solution-process Organic light-emitting diodes Hole injection layer HAT-CN

1. Introduction Solution-processed organic optoelectronic devices such as organic light-emitting diodes (OLEDs) and organic solar cells have gained immense attention owing to their potential advantages such as ultra-low-cost and very large-area production [13]. The conductive polymer poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS) has been widely used as the hole-injection layer (HIL) in conventional solution-processed organic optoelectronic devices, which have exhibited good performances [410]. However, PEDOT:PSS suffers from certain limitations such as its strongly acidic and hygroscopic nature, exciton quenching properties, and aqueous solubility, resulting in
Corresponding author.
E-mail address: hwlin@mx.nthu.edu.tw (H.-W. Lin). 1566-1199/$ - see front matter 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.orgel.2013.02.011

a decrease in the device performance and stability [11 14]. Even though PEDOT:PSS replacements such as inorganic metal oxides [1518] and graphene oxides [19,20] have been successfully adopted for use in solution-processed organic solar cells, PEDOT:PSS continues to be used exclusively as the HIL in solution-processed OLEDs [46]. This is probably because of the nonideal adhesion of the layers formed on top of the HIL, or owing to the poorer performance encountered, when these other materials are used as the HIL. Therefore, nding an effective HIL for solution-processed OLEDs is required for enhancing the device stability and furthering mass production. In this study, we explore the use of a new kind of nanostructured, roomtemperature solution-processed, ultrathin layer of 1,4,5,8,9,11-hexaazatriphenylene hexacarbonitrile (HATCN) as a HIL. In particular, it is demonstrated that HATCN is selectively soluble in a very specic solvent and is

H.-W. Lin et al. / Organic Electronics 14 (2013) 12041210

1205

resistant against most other solvents commonly used for casting top layers. This prevents interfacial mixing and erosion of the HAT-CN underlayer during further processing. Analysis of the electronic energy levels of the HAT-CN layer and the hole-transporting layer (HTL) formed on top of it suggests that an efcient charge transfer occurs at the HAT-CN/HTL interface, which facilities hole injection from the anode. Solution-processed phosphorescent red, green, and blue OLEDs fabricated using HAT-CN HILs are demonstrated to exhibit a dramatic enhancement of the device efciency compared to devices without HILs. The efciency values are largely comparable with those of optimized PEDOT:PSS-based devices for all colors. The results reveal the potential of solution-processed HAT-CN as a promising universal hole-injection material for OLEDs.

2. Experimental 2.1. Device fabrication The chemical structures of the compounds and the device architecture utilized in this study are shown in

Fig. 1. Prior to organic thin-lm deposition, prepatterned indium tin oxide (ITO) glass substrates (sheet resistance of $12 X/sq) were cleaned in an ultrasonic bath using deionized water, 2-propanone, and methanol in sequence for 15 min each. The substrates were then UV-ozone treated before depositing the HILs. For the PEDOT:PSS-based devices, PEDOT:PSS ($40 nm) was rst spin-coated onto the substrates and then baked in air at 135 C for 30 min to remove any residual water. In the case of the HAT-CNbased devices, a solution of HAT-CN (3 mg/ml) was prepared in a nitrogen glove box using 2-propanone as the solvent, and subsequently, the solution was spin-coated onto the substrates. After the deposition of the HILs, a thermally polymerized 2,7-disubstituted uorene-based triaryldiamine (VB-FNPD) solution (10 mg/ml) was prepared using chlorobenzene as the solvent and was spin-coated onto the samples. The casted VB-FNPD lms were subsequently baked on a hot plate at 100 C for 30 min to remove any residual solvent and again at 230 C for 90 min to form a cross-linked HTL [21,22]. The emission layers (EMLs) which consisted of a mixture of the host material 4,40 ,400 tris(N-carbazolyl)-triphenylamine (TCTA) (10 mg/ml) and

(a)
N NC N NC NC N N N CN N N O CN O N

CN

HAT-CN

VB-FNPD
CF3

N F Ir

N O O N

N Ir N

S N N S

N Os N

N P P CF3

FIrpic

Ir(mppy)3

Os(btfp)2(pp2b)

(b)

Fig. 1. (a) Chemical structures of materials and (b) the device architecture utilized in this study.

1206 Table 1 Solubilities of HAT-CN in various solvents. Solvent Solubility 2-Propanone 10 mg/ml

H.-W. Lin et al. / Organic Electronics 14 (2013) 12041210

Toluene x

CB x

DCB x

TCB x

Chloroform x

DCM x

x: Insoluble (solubility <0.1 mg/ml), CB: chlorobenzene, DCB: 1.2-dichlorobenzene, TCB: 1,2,4-trichlorobenzene, DCM: 1,2-dichloromethane.

the phosphorescent dopants Os(btfp)2(pp2b) [22], tris-(24(4-toltyl)phenylpyridine)iridium(III) (Ir(mppy)3) [23], and bis(40 ,60 -diuorophenylpyridinato)iridium(III) picolinate (FIrpic) [24] in toluene, were spin-cast onto the cross-linked VB-FNPD layer to fabricate red, green, and blue OLEDs, respectively. The triplet energy of the TCTA host is $2.8 eV, which is larger than that of FIrpic (2.65 eV), Ir(mppy)3 (2.4 eV), and Os(btfp)2(pp2b) (1.96 eV) dopants. Vacuum-deposited 2,20 ,200 -(1,3,5-benzenetriyl)tris[1-phenyl-1H-benzimidazole] (TPBi) or 1,3,5tri[(3-pyridyl)-phen-3-yl]benzene (TmPyPB) was used as the electron-transporting layers (ETLs), and cesium uoride (CsF) was employed as the electron-injection layer. For the blue devices, an additional TCTA layer was inserted between the VB-FNPD layer and the EML to prevent exciton quenching near the VB-FNPD layer. The detailed device structures were as follows:  Device A: ITO (150 nm)/HAT-CN (4 nm)/VB-FNPD (35 nm)/TCTA: Os(btfp)2(pp2b) 10 wt% (20 nm)/TPBi (60 nm)/CsF (1 nm)/Al (120 nm).  Device B: ITO (150 nm)/PEDOT:PSS ($40 nm)/VB-FNPD (35 nm)/TCTA: Os(btfp)2(pp2b) 10 wt% (20 nm)/TPBi (60 nm)/CsF (1 nm)/Al (120 nm).  Device C: ITO (150 nm)/VB-FNPD (35 nm)/TCTA: Os(btfp)2(pp2b) 10 wt% (20 nm)/TPBi (60 nm)/CsF (1 nm)/Al (120 nm).

 Device D: ITO (150 nm)/HAT-CN (4 nm)/VB-FNPD (35 nm)/TCTA:Ir(mppy)3 10 wt% (20 nm)/TPBi (60 nm)/ CsF (1 nm)/Al (120 nm).  Device E: ITO (150 nm)/PEDOT:PSS ($40 nm)/VB-FNPD (35 nm)/TCTA:Ir(mppy)3 10 wt% (20 nm)/TPBi (60 nm)/ CsF (1 nm)/Al (120 nm).  Device F: ITO (150 nm)/VB-FNPD (35 nm)/ TCTA:Ir(mppy)3 10 wt% (20 nm)/TPBi (60 nm)/CsF (1 nm)/Al (120 nm).  Device G: ITO (150 nm)/HAT-CN (4 nm)/VB-FNPD (35 nm)/TCTA (10 nm)/TCTA: FIrpic 10 wt% (20 nm)/ TmPyPB (60 nm)/CsF (1 nm)/ Al (120 nm).  Device H: ITO (150 nm)/PEDOT:PSS ($40 nm)/VB-FNPD (35 nm)/TCTA (10 nm)/TCTA: FIrpic 10 wt% (20 nm)/ TmPyPB (60 nm)/CsF (1 nm)/Al (120 nm).  Device I: ITO (150 nm)/VB-FNPD (35 nm)/TCTA (10 nm)/TCTA: FIrpic 10 wt% (20 nm)/TmPyPB (60 nm)/CsF (1 nm)/Al (120 nm).

2.2. Measurements Thin lms for atomic force microscopy (AFM) and ultraviolet photoemission spectroscopy (UPS) characterization were prepared under the same conditions as those for the OLEDs. The AFM images were obtained using a Bruker Dimension Icon atomic force microscope in the tapping mode. The UPS measurements were performed using He I

(a) ITO
20 nm 10 nm 1 m

(b) PEDOT: PSS

1 m

20 nm 10 nm 1 m

1 m

0 m

0 m

(c) HAT-CN
20 nm 10 nm 1 m

(d) HAT-CN / VB-FNPD


20 nm 10 nm 1 m

1 m

1 m

0 m

0 m

Fig. 2. AFM images of (a) ITO (Rq = 0.65 nm), (b) ITO/PEDOT:PSS (40 nm) (Rq = 0.86 nm), (c) ITO/HAT-CN (4 nm) (Rq = 2.82 nm), and (d) ITO/HAT-CN (4 nm)/ VB-FNPD (35 nm) (Rq = 1.33 nm).

H.-W. Lin et al. / Organic Electronics 14 (2013) 12041210

1207

3. Results and discussion 3.1. Solubility test First, the solubility of HAT-CN in various common solvents was determined; the results are summarized in Table 1. Interestingly, HAT-CN dissolved only in 2propanone (acetone) with a solubility (10 mg/ml) suitable for thin-lm deposition, but remained undissolved in toluene, chlorobenzene, 1,2-dichlorobenzene, 1,2,4-trichlorobenzene, chloroform, and 1,2-dichloromethane. This selective solubility of HAT-CN allowed us to form rigid lms of it using the solution process, with the lms remaining intact during the fabrication processes that followed. The morphology of the HAT-CN layer was evaluated using AFM and is shown in Fig. 2. Unlike the smooth PEDOT:PSS lm, it was found that HAT-CN formed nanoaggregates on the ITO-coated substrate, possibly owing to extremely dense molecular packing. The HAT-CN thin layer with an average thickness of 4 nm exhibited a root-meansquare (RMS) roughness (Rq) of up to 2.82 nm. It is possible that these nanoaggregates facilitate carrier injection, as has been previously observed in nanopatterned cathodes [26 28]. It has been found that a charge-carrier-generating interface is formed between layers of vacuum-deposited HATCN and the hole-transporting material, N,N0 -di(naphthalene-1-yl)-N,N0 -diphenylbenzidine (NPD) [29]. A holeinjection efciency of close to 100% was demonstrated in such a HAT-CN/NPD system [30]. Moreover, high-efciency tandem OLEDs utilizing vacuum-deposited HAT-CN/NPD as the interconnection layers have also been successfully demonstrated. Hence, it is logical to adopt a molecular system that is similar to NPD, but suitable for solution processing, to be combined with the solution-processed HAT-CN lms. To achieve this purpose, a cross-linkable derivative of NPD, i.e., VB-FNPD, was used as the HTL in this study. The cross-linked VB-FNPD lm showed smooth morphology with Rq of 1.33 nm. 3.2. UPS measurement The electronic energy levels of ITO/HAT-CN/VB-FNPD were determined using UPS. Fig. 3a and b shows the evolution of the valence band spectra of HAT-CN and VB-FNPD lms deposited on an ITO substrate. The ITO layer exhibited a work function of 4.8 eV in response to a UV-ozone treatment of 15 min. Upon HAT-CN deposition, the work

Intensity (a.u.)

(21.2 eV) and He II (40.8 eV) as excitation sources in an ultra-high-vacuum chamber with a base pressure of 109 Torr [25]. The current densityvoltageluminance (JVL) characteristics of the OLEDs were obtained using a 2636A SourceMeter (Keithley Instruments Inc.) and a silicon photodetector calibrated using a PR-650 SpectraScan colorimeter (Photo Research, Inc.). The electroluminescence (EL) spectra of the devices were recorded using an Ocean Optics spectrometer. The thicknesses of the organic lms were determined using a V-VASE (JA Woollam Inc.) variable-angle spectroscopic ellipsometer.

function increased to $5.6 eV and the position of the highest occupied molecular orbital (HOMO) of the coated HATCN layer was observed to be $3.6 eV. The energy levels determined herein are compatible with those reported previously for vacuum-deposited HAT-CN [29,31]. Slight deviations between the values for solution-processed and

(a)

He I h =
HAT-CN 19 nm 9 nm 4 nm ITO substrate

16

14

0 =E F

Binding energy (eV)

eV

Intensity (a.u.)

(b)
Intensity (a.u.)

He I h =

VB-FNPD: 0 21.4 nm

Binding energy (eV)

HAT-CN 4 nm

VB-FNPD 21.4 nm 18.7 nm 2.9 nm 1.6 nm 0 nm

16

14

0 =E

Binding energy (eV)

(c)

Fig. 3UPS spectra of lms of different thickness of (a) ITO, ITO/HAT-CN and (b) ITO/HAT-CN/VB-FNPD. The inset shows the low-energy region at a larger magnication. (c) The corresponding energy level diagram of ITO/HAT-CN/VB-FNPD.

1208

H.-W. Lin et al. / Organic Electronics 14 (2013) 12041210

vacuum-deposited layers of HAT-CN/NPD may be due to variations in the molecular packing derived from the various deposition methods. Nevertheless, the results provide direct evidence that the HAT-CN lm was successfully deposited onto the ITO substrate by solution processing, and consequently, the properties of the anode were altered signicantly. The HAT-CN/ITO samples were coated with layers of VB-FNPD of various thicknesses. The results of UPS-based analyses of these samples are shown in Fig. 3b; the lower binding energy regions of the curves are magnied in the inset of the gure. The onset of the HOMO is observed at $0.3 eV for a thickness of 1.6 nm of the VB-FNPD lm. This energy gradually increases to $0.5 eV as the VB-FNPD thickness increases. The corresponding energy level diagram is shown in Fig. 3c. The small difference between the Fermi level, the lowest unoccupied molecular orbital (LUMO) level of HAT-CN, and the HOMO level of VB-FNPD indicates that

electrons can be excited from the HOMO level of VB-FNPD to the LUMO level of HAT-CN effortlessly, which was equivalent to VB-FNPD being p-doped by HAT-CN [29,32]. Thus, the interface between HAT-CN and VB-FNPD layers can be assumed to be an ohmic contact and should exhibit good hole-injection ability [30]. 3.3. OLEDs characteristics The promising results from UPS prompted further exploration of the HAT-CN layers in the solution-processed phosphorescent OLEDs. Fig. 4a and b shows the JVL characteristics and the efciencies of red OLEDs fabricated with the novel solution-processed HAT-CN HIL (Device A), with a conventional PEDOT:PSS HIL (Device B) and without HIL (Device C). The current densities of the device without HIL (Device C) were characterized by a very high turn-on voltage of up to 6.8 V, owing to the large energy barrier between the ITO substrate and the hole-transporting VB-

Luminance cd/m

10 10 10 10

Device A Device B Device C

Current Efficiency (cd/A)

(a)
2

10

100 80 60 40 20

(b)

21 18 15 12 9 6 3 0 1

Device A Device B Device C

21 18 15 12 9 6 3 0 10000

Current Density mA/cm2

Power Efficiency (lm/W)

-1

-3

10

12

0 14

10

100

1000

Voltage (V)

Luminance cd/m2
Current Efficiency (cd/A)
100 80 60 40 20

(c)
Luminance cd/m
2

10 10 10 10 10

Device D Device E Device F

(d)

70 60 50 40 30 20 10 0 1

Device D Device E Device F

80 70 60 50 40 30 20 10 0 10000

Current Density mA/cm2

Power Efficiency (lm/W)

-1

-3

0 10 12 14 16

10

100

1000

Voltage (V)
Current Efficiency (cd/A)

Luminance cd/m2
100 80 60 40 20

(e)
2

10 10 10 10 10

Luminance cd/m

Device G Device H Device I

(f)

30 25 20 15 10 5 0 1

Device G Device H Device I

21 18 15 12 9 6 3 0 10000

Current Density mA/cm2

Power Efficiency (lm/W)

-1

-3

0 10 12 14 16

10

100

1000

Voltage (V)

Luminance cd/m2

Fig. 4. Current densityvoltageluminance characteristics, luminous efcacy, and current efciency of solution-processed (a and b) red, (c and d) green, and (e and f) blue phosphorescent OLEDs. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)

H.-W. Lin et al. / Organic Electronics 14 (2013) 12041210 Table 2 EL characteristics of solution-processed OLEDs with various hole-injection materials. Device EQE (%) [a] A B C D E F G H I 13.9 11.4 11.2 14.7 14.5 10.7 11.4 11.5 10.1 [b] 12.9 11.2 11.2 14.6 14.4 10.5 11.3 11.1 10.0 Current efciency (cd/A) [a] 14.5 11.8 11.2 50.9 49.6 36.7 23.3 24.5 21.7 [b] 13.5 11.6 11.2 50.8 49.2 36.1 23.3 23.8 21.4 Power efciency (lm/W) [a] 15.2 8.5 4.5 55.0 44.3 12.5 15.8 13.4 6.6 [b] 8.1 7.3 3.5 39.9 38.7 10.3 13.0 12.9 6.3 Driving voltage (V) [c] 3.0 3.1 6.8 2.7 2.7 8.1 3.7 3.3 7.4 [d] 7.6 7.4 12.8 5.7 5.7 12.9 7.3 7.8 13.2

1209

[a] Maximum efciencies; [b] measured at 100 cd/m2; [c] measured at 1 cd/m2; [d] measured at 1000 cd/m2.

Fig. 5. Images of OLEDs utilizing a solution-processed HAT-CN injection layer.

FNPD layer. The introduction of the HAT-CN layer between the ITO and VB-FNPD HTL layers decreased the turn-on voltage signicantly to 3.0 V, a value even lower than that for the PEDOT:PSS-based device. The HAT-CN-based device showed a maximum efciency of 13.9% (external quantum efciency, EQE), a current efciency of 14.5 cd/A, and a luminous efcacy of 15.2 lm/W, which outperformed the PEDOT:PSS-based device (11.4%, 11.8 cd/A, and 8.5 lm/W) and the device without HIL (11.2%, 11.2 cd/A, and 4.5 lm/ W). This superior performance could be attributed to the extremely effective hole injection associated with the use of the HAT-CN layer, which prohibits exciton recombination near the conductive anodes. The EL characteristics are summarized in Table 2. On the basis of the favorable results obtained for the HAT-CN-based red devices, the green- and blue-emitting phosphorescent iridium complexes Ir(mppy)3 and FIrpic were selected and combined with the host TCTA for the fabrication of green (Devices D, E, and F) and blue devices (Devices G, H, and I), respectively. Fig. 4c and d shows the JVL characteristics and the efciencies of the green devices. As anticipated, the devices fabricated with an HATCN HIL (Device D) exhibited a notable augmentation of the current density compared to the device without HILs (Device F); this increase can also be ascribed to a decrease in the hole-injection barrier. An EQE of 14.7%, a current efciency of up to 50.9 cd/A, and a luminous efcacy as high as 55 lm/W were achieved in the HAT-CN-based green device (Device D). An excellent current efciency ($50 cd/ A) was maintained across a varying luminance from 1 cd/ m2 to 103 cd/m2 in Device D. Fig. 4e and f shows the JVL characteristics and the efciencies of the blue devices. Considering that effective

exciton connement is a key point in fabricating phosphorescent blue OLEDs and accounting for the low triplet energy of VB-FNPD (2.29 eV), solution-processed TCTA was used as an interlayer to prevent exciton quenching by VB-FNPD. Moreover, TmPyPB was used as an electrontransporting layer instead of TPBi owing to its hole-blocking nature derived from the deeper lying HOMO level and a higher electron mobility relative to TPBi [33]. The maximum EQE and current efciency of the HAT-CN-based blue device (Device G) were fully comparable to those of the PEDOT:PSS-based device (Device H) (11.4% and 23.3 cd/A vs. 11.5% and 24.5 cd/A, respectively). Furthermore, the luminous efcacy, which is the parameter most relevant to carrier injection, was evidently superior for the HATCN-based device than for the PEDOT:PSS device (15.8 lm/ W vs. 13.4 lm/W). Images of the HAT-CN-based devices (Device A, D, G) are shown in Fig. 5.

4. Conclusion In conclusion, it was demonstrated that HAT-CN is selectively soluble in 2-propanone and is insoluble in other solvents commonly utilized for solution processing (such as toluene, chlorobenzene, 1,2-dichlorobenzene, 1,2,4-trichlorobenzene, chloroform, and 1,2-dichloromethane). This unique characteristic facilitated the formation of a rigid HIL in multilayer solution-processed devices. The energy levels of HAT-CN and of the VB-FNPD HTLs formed on top of them were determined. On the basis of the obtained values, it was determined that the high hole-injection capability was owing to the charge transfer between the HAT-CN and VB-FNPD layers. Solution-processed phos-

1210

H.-W. Lin et al. / Organic Electronics 14 (2013) 12041210 [9] D.W. Li, L.J. Guo, Appl. Phys. Lett. 88 (2006) 063513. [10] W.S. Chung, H. Lee, W. Lee, M.J. Ko, N.G. Park, B.K. Ju, K. Kim, Org. Electron. 11 (2010) 521. [11] J.S. Kim, P.K.H. Ho, C.E. Murphy, A. Seeley, I. Grizzi, J.H. Burroughes, R.H. Friend, Chem. Phys. Lett. 386 (2004) 2. [12] J.S. Kim, R.H. Friend, I. Grizzi, J.H. Burroughes, Appl. Phys. Lett. 87 (2005) 023506. [13] K.H. Yim, R. Friend, J.S. Kim, J. Chem. Phys. 124 (2006) 184706. [14] M.P. de Jong, L.J. van Ijzendoorn, M.J.A. de Voigt, Appl. Phys. Lett. 77 (2000) 2255. [15] Y.M. Sun, G.C. Welch, W.L. Leong, C.J. Takacs, G.C. Bazan, A.J. Heeger, Nat. Mater. 11 (2012) 44. [16] J. Subbiah, P.M. Beaujuge, K.R. Choudhury, S. Ellinger, J.R. Reynolds, F. So, Org. Electron. 11 (2010) 955. [17] A.K.K. Kyaw, X.W. Sun, C.Y. Jiang, G.Q. Lo, D.W. Zhao, D.L. Kwong, Appl. Phys. Lett. 93 (2008) 221107. [18] H. Choi, B. Kim, M.J. Ko, D.K. Lee, H. Kim, S.H. Kim, K. Kim, Org. Electron. 13 (2012) 959. [19] S.S. Li, K.H. Tu, C.C. Lin, C.W. Chen, M. Chhowalla, ACS Nano 4 (2010) 3169. [20] J.M. Yun, J.S. Yeo, J. Kim, H.G. Jeong, D.Y. Kim, Y.J. Noh, S.S. Kim, B.C. Ku, S.I. Na, Adv. Mater. 23 (2011) 4923. [21] C.Y. Lin, Y.C. Lin, W.Y. Hung, K.T. Wong, R.C. Kwong, S.C. Xia, Y.H. Chen, C.I. Wu, J. Mater. Chem. 19 (2009) 3618. [22] B.-S. Du, J.-L. Liao, M.-H. Huang, C.-H. Lin, H.-W. Lin, Y. Chi, H.-A. Pan, G.-L. Fan, K.-T. Wong, G.-H. Lee, P.-T. Chou, Adv. Funct. Mater. 22 (2012) 3491. [23] M. Cai, T. Xiao, E. Hellerich, Y. Chen, R. Shinar, J. Shinar, Adv. Mater. 23 (2011) 3590. [24] 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. [25] C.I. Wu, C.T. Lin, Y.H. Chen, M.H. Chen, Y.J. Lu, C.C. Wu, Appl. Phys. Lett. 88 (2006) 152104. [26] D. Liu, M. Fina, Z. Chen, X. Chen, G. Liu, S. Johnson, S.S. Mao, Appl. Phys. Lett. 91 (2007) 093514. [27] T. Earmme, E. Ahmed, S.A. Jenekhe, Adv. Mater. 22 (2010) 4744. [28] T. Earmme, S.A. Jenekhe, J. Mater. Chem. 22 (2012) 4660. [29] Y.K. Kim, J.W. Kim, Y. Park, Appl. Phys. Lett. 94 (2009) 063305. [30] C.E. Small, S.-W. Tsang, J. Kido, S.K. So, F. So, Adv. Funct. Mater. 22 (2012) 3261. [31] S.M. Park, Y.H. Kim, Y. Yi, H.Y. Oh, J.W. Kim, Appl. Phys. Lett. 97 (2010) 063308. [32] K.S. Yook, S.O. Jeon, J.Y. Lee, Thin Solid Films 517 (2009) 6109. [33] S.J. Su, T. Chiba, T. Takeda, J. Kido, Adv. Mater. 20 (2008) 2125.

phorescent OLEDs fabricated using HAT-CN HILs exhibited outstanding performances for all colors. To the best of our knowledge, this is the rst demonstration of a solutionprocessable alternative to PEDOT:PSS, with the resulting OLEDs exhibiting performances that were highly comparable in every aspect to those of PEDOT:PSS-based ones. We believe that these results should open up new avenues for the development of high-performance and high-stability solution-processed OLEDs for mass production. Acknowledgements The authors would like to thank Prof. Ken-Tsung Wong for supplying the sample of VB-FNPD and Prof. Yun Chi for supplying the sample of Os(btfp)2(pp2b) used in this study. The work was nancially supported by National Science Council of Taiwan (NSC 101-2112-M-007-017-MY3) and the Low Carbon, Energy Research Center, National Tsing Hua University. References
[1] C.-Y. Chen, H.-W. Chang, Y.-F. Chang, B.-J. Chang, Y.-S. Lin, P.-S. Jian, H.-C. Yeh, H.-T. Chien, E.-C. Chen, Y.-C. Chao, H.-F. Meng, H.-W. Zan, H.-W. Lin, S.-F. Horng, Y.-J. Cheng, F.-W. Yen, I.F. Lin, H.-Y. Yang, K.-J. Huang, M.-R. Tseng, J. Appl. Phys. 110 (2011) 094501. [2] T. Winkler, H. Schmidt, H. Flugge, F. Nikolayzik, I. Baumann, S. Schmale, T. Weimann, P. Hinze, H.H. Johannes, T. Rabe, S. Hamwi, T. Riedl, W. Kowalsky, Org. Electron. 12 (2011) 1612. [3] A.C. Arias, J.D. MacKenzie, I. McCulloch, J. Rivnay, A. Salleo, Chem. Rev. 110 (2010) 3. [4] L.D. Hou, L.A. Duan, J.A. Qiao, D.Q. Zhang, G.F. Dong, L.D. Wang, Y. Qiu, Org. Electron. 11 (2010) 1344. [5] Y.J. Doh, J.S. Park, W.S. Jeon, R. Pode, J.H. Kwon, Org. Electron. 13 (2012) 586. [6] B.H. Zhang, G.P. Tan, C.S. Lam, B. Yao, C.L. Ho, L.H. Liu, Z.Y. Xie, W.Y. Wong, J.Q. Ding, L.X. Wang, Adv. Mater. 24 (2012) 1873. [7] F.C. Krebs, Org. Electron. 10 (2009) 761. [8] R.J. Wang, D. Liu, R. Zhang, L.J. Deng, J.Y. Li, J. Mater. Chem. 22 (2012) 1411.

Das könnte Ihnen auch gefallen