Eun Young Parka,
Da Hwan Leea,
Thi Na Lea,
Chol-Min Shinb,
Jihoon Lee*b and
Min Chul Suh*a
aDepartment of Information Display, Kyung Hee University, Dongdaemun-gu, Seoul 02447, Republic of Korea. E-mail: mcsuh@khu.ac.kr
bDepartment of Polymer Science and Engineering, Department of IT·Energy Convergence (BK21 FOUR), Korea National University of Transportation, Chungju 27469, Republic of Korea. E-mail: jihoonli@ut.ac.kr
First published on 1st September 2021
We designed and synthesized a new host material with a highly soluble and thermally stable indenocarbazole derivative (7,7-dimethyl-5-phenyl-2-(9-phenyl-9H-carbazol-3-yl)-5,7-dihydro-indeno[2,1-b]carbazole) that can make green phosphorescent organic light-emitting diodes (PHOLEDs) in a solution process. In particular, these are used in a blue common layer structure in which green and red-emitting layers are formed by a solution process and blue common layers are thermally evaporated. The new host material possesses excellent hole transport capability and high triplet energy (T1). Mainly we designed the hole dominant material to keep the exciton forming area away from the hole transport layer (HTL) and emitting layer (EML) interface, an interfacial mixing area to improve device performance. As a result, the greatest lifetime of 1300 hours was achieved and a high current efficiency of up to 66.3 cd A−1 was recorded when we used the optimized device structure of a 5 nm thick bipolar exciton blocking layer (B-EBL). It may be a good agreement of exciton confinement and reduced electron accumulation at the HTL and EML interface.
On the other hand, even if the blue light-emitting material forms a thin film by thermal evaporation technology, the lifetime is still short, so the long lifetime can't be guaranteed by the solution process. Therefore, a method of thermally evaporating blue EML as a common layer on the top patterned by a wet process such as ink-jet printing was introduced to overcome an issue of intrinsically short operational lifetime of blue OLED device which may show much shorter one if they are formed by solution process. In other words, to realize a blue common layer structure, we could print hole injection layer (HIL), HTL, red and green EML, and then use a thermal evaporation process to deposit a bipolar exciton blocking layer (B-EBL), blue EML, electron transport layer (ETL), electron injection layer (EIL), and cathode continuously. Among the many functional layers mentioned above, B-EBL plays an important role in inhibiting the triplet exciton energy of green or red dopant from being transmitted to the triplet state of the blue-light emitting layer to prevent blue light emission by TTA. Nevertheless, the performance of red and green devices manufactured by the solution process still shows poor performance compared to the thermal vacuum evaporation process. Hence, many reports have attempted to improve the performance of devices fabricated by solution processes. Here, we present a thermally stable device and a highly efficient solution-processed OLED with such a blue common layer structure.9–12
To understand the molecular geometry of the ground and the excited state of the new synthetic host material (PCIC), we carried out a molecular simulation following density functional theory (DFT) with the Schrödinger Materials Science program (version 4.1). We also proceeded with a time-dependent DFT (TD-DFT) calculation to obtain singlet energy level (S1) and triplet energy level (T1) of an optimized geometry of the host molecule.19 The theory and basis set we applied for all calculations was B3LYP/6-31G (d,p). Looking at the PCIC from the side of the geometry optimized structure, it can be seen that the indenocarbazole moiety and carbazole moiety are bent with a dihedral angle of about 42°.
This structure has a slightly smaller dihedral angle compared to the previously developed materials, but it was judged that additionally introduced methyl groups and phenyl groups hindered the stacking or ordering of molecules, indicating that they would be suitable for the solution process.20 Looking at the electron cloud distributions of HOMO and LUMO predicted from molecular simulation, it was confirmed that the electron cloud bands of HOMO spread widely along the two carbazoles, and then converge to one side in LUMO as shown in Fig. 1. The calculated HOMO and LUMO energy levels estimated from the corresponding optimized molecular geometry were −5.53 eV and −2.34 eV, respectively, and T1 was estimated to be 2.76 eV. In addition, some electronic properties of PCIC material were measured experimentally (Fig. 4S†). From absorption edge analysis, the Eg of 3.18 eV was obtained, and the HOMO level of 5.48 eV was estimated by cyclic voltammetry. From those data, the LUMO level was calculated as 2.3 eV. Meanwhile, from the photoluminescence (PL) spectrum, we estimated a T1 energy level. In other words, the T1 energy level of PCIC measured at low temperature (77 K) was determined to be 2.75 eV (Fig. 2). However, according to a recent report by Lee et al., the phosphorescent dopant can be quasi-dewet from EML and escape during annealing procedure during solution processes if the thermal stability of the host materials is very low.3 To confirm such a possibility, we proceeded with differential scanning calorimetry (DSC) analysis of the PCIC material and, in the end, confirmed that the glass transition temperature (Tg) is very low, about 98 °C, and the Tm is about 234 °C (Table 1). So, we performed molecular dynamics simulation to predict how the relative distance between molecules can be changed on series of annealing processes. However, we simulated the molecular stacking or ordering behavior by using imaginary spaces filled with PCIC materials mixed with green dopant molecules (Ir(mppy)3, 5 mol%) after the annealing process (Fig. 3).
Fig. 1 HOMO, LUMO distributions and energy level of PCIC predicted through DFT and TD-DFT calculations. |
Host | λabs (nm) | λPL (nm) | EHOMOb (eV) | ELUMOc (eV) | Eoptgc (eV) | Etd (eV) | Tge (°C) | Tme (°C) | Tdf (°C) | ||
---|---|---|---|---|---|---|---|---|---|---|---|
Sola | Film | Sola | Film | ||||||||
a Toluene solution (1.0 × 10−5 M).b Measured by cyclic voltammetry.c Estimated from the absorption spectra threshold.d Estimated from the phosphorescence spectra in toluene at 77 K.e Estimated from DSC measurements.f Estimated from TGA measurements corresponding to 5% weight loss. | |||||||||||
PCIC | 318 | 318 | 387, 407 | 395, 415 | −5.48 | −2.30 | 3.18 | 2.75 | 98 | 234 | 482 |
For molecular dynamics simulation, we created a tangled chain type unit cell containing a total of 226 molecules (10 Ir(mppy)3 and 216 PCIC molecules) by performing the Disordered System Builder mode with OPLS3e force filed applied.21 Then, molecular dynamics (MD) simulation to reach the equilibrium system with 10.0 ns and NPT conditions as well as the MD protocols (relaxation protocol) of three stage conditions was performed.22 Next, simulations were performed with 50.0 ns and NPT conditions at 100 °C, 120 °C, and 140 °C to see the annealing effect, and finally MD simulations were performed under 10.0 ns and 300 K conditions (Fig. 3(left)). Then, the value of g(r) was calculated as the probability of finding a molecule within the distance r through radial distribution functions (RDF) simulation. The calculated data are shown in Fig. 3(right).
Fig. 3 Molecular structure containing PCIC and 5 mol% of Ir(mppy)3 in a unit cell (left), radial distribution functions (RDFs) calculated by simulation method (right). |
Very interestingly, the intermolecular distance decreased from 3.3 Å to 2.7 Å as the temperature increased from 120 °C to 140 °C. In other words, when the annealing temperature of PCIC materials is increased, the distance between molecules does not increase, but rather gets closer, which means that PCIC molecules have less chance of segmental motion. By analogy with these facts, it means that PCIC materials do not cause severe segmental motion. We also performed AFM analysis to identify the surface roughness of synthetic PCIC film as well as the film composed of PCIC doped with 5 mol% of Ir(mppy)3. First, the average roughness (Ra), peak to valley height (Rpv), root mean square roughness (Rq) values of the pure PCIC surface were measured as 0.42, 4.19, and 0.52 nm, respectively. In addition, Ra, Rpv, and Rq were measured as 0.27, 3.67, and 0.35 nm, respectively, from the thin film of PCIC doped with 5 mol% of Ir(mppy)3. From these results, we confirmed the PCIC molecule, as well as PCIC doped with 5 mol% of Ir(mppy)3, is a good approach for the solution process device (Fig. 4 and Table 2).
Film | Ra (nm) | Rpv (nm) | Rq(nm) |
---|---|---|---|
PCIC | 0.42 | 4.19 | 0.52 |
PCIC:Ir(mppy)3 (5 mol%) | 0.27 | 3.67 | 0.35 |
Fig. 6(a) shows the J–V–L characteristics of green solution process OLEDs. The turn-on voltage (Von) at a constant luminance of 1 cd m−2 was 3.5 V for device A. The driving voltage (Vop) obtained at a luminance of 1000 cd m−2 was 6.5 V for device A. Likewise, the current density at 6.0 V was 0.82 mA cm−2 for device A. Meanwhile, at a constant luminance of 1000 cd m−2, the current efficiency was recorded to 55.2 cd A−1 for device A (Fig. 6(b)). Besides, the maximum current efficiency was 55.4 cd A−1 for device A in Table 3. The peak wavelength of the EL spectrum of the resulting device was 512 nm as shown in Fig. 6(c). Although we recorded high values in terms of efficiency, operational stability (Fig. 6(d)), the best half-life obtained here is only 5.0 hours for device A. As a solution device, the efficiency was quite high, but device A had a relatively low operational stability and a lifetime of only 5 hours, so we recognized that device optimization is required.24
Fig. 6 (a) J–V–L curves, (b) current efficiency, (c) EL spectra property, and (d) operational lifetime characteristics of PHOLEDs. |
Therefore, in order to overcome the low driving performance of green phosphorescent device (device A) manufactured by the solution process, we used a blue common layer approach that can more effectively limit the inflow of electrons and move the exciton formation region away from the interface mixing zone. Even if the blue common layer is deposited as a common layer on the top, electrons passing through the blue light-emitting layer on the top must be quickly transferred to green in order to emit light only from the wet-processed green emitter on the bottom. The holes through the lower green layer should enter very limitedly into the blue light-emitting layer. Since the layer that can perform this function is B-EBL, and B-EBL has a role to prevent migration of the triplet exciton made in the lower green emission layer to the blue emission layer, too. Thus, the bipolar materials with high T1 should be selected. The reason for this is as follows: if energy corresponding to T1 of the under-lying phosphorescent green is transferred to the T1 state of the blue light-emitting material on the top due to its low T1, side light emission may occur in the blue light emitter through the TTA process inside blue light emitting layer. Thus, we inserted a bipolar exciton blocking layer (B-EBL) with a high T1 energy level (DIC-TRZ, T1 = 2.82 eV)25 between the green emitter and the blue emitter to suppress the migration of green exciton to the adjacent blue layer as shown in the top left part in Fig. 7.
Therefore, we first tried to optimize the thickness of B-EBL and ETL because they are the most important layers to realize the blue common layer approach. For this experiment, in order to maintain the total electric field applied to the entire device, the total thickness of B-EBL, BCL, and ETL was kept at 40 nm. The detailed configuration of the devices fabricated is as follows:
ITO/AQ1200/x-HTL/PCIC:Ir(mppy)3 (5 mol%)/B-EBL (x nm)/BH:BD/TPBI (25–x nm)/LiF/Al (where x = 3, 5, 7 nm for devices B, C and D) (device configurations, including thickness conditions, are given in ESI†).
Fig. 8(a) shows the J–V–L characteristics of green solution process OLEDs. The turn-on voltages (Von) at a constant luminance of 1 cd m−2 were 4.5 V for devices B, C and D. The driving voltages (Vop) obtained at a luminance of 1000 cd m−2 were 8.6 V, 7.9 V, and 7.8 V for devices B, C and D, respectively. Likewise, the current density at 6.0 V was 0.13 mA cm−2 for device B, 0.15 mA cm−2 for device C and 0.22 mA cm−2 for device D, respectively. Although the thickness of B-EBL increased from 3 nm to 7 nm, it can be seen that the amount of current reaching the EML increases as the thickness of TPBI decreases from 22 nm to 18 nm. This might be due to the higher electron mobility of B-EBL (2.5 × 10−3 cm2 V−1 s−1) than TPBI (∼10−5 cm2 V−1 s−1) so that the recombination region more influences the thickness change of TPBI. In other words, the driving voltage gradually decreases as the thickness of TPBI, which has low mobility, decreases. More specifically, at a constant luminance of 1000 cd m−2, the current efficiency was recorded to 53.6 cd A−1 for device B, 60.3 cd A−1 for device C and 56.7 cd A−1 for device D, respectively (Fig. 8(b)). Besides, the maximum current efficiency was 53.7 cd A−1 for device B, 60.3 cd A−1 for device C and 60.9 cd A−1 for device D, respectively, as summarized in Table 4. The EL spectra of the three devices were obtained identically as shown in Fig. 8(c), despite the different conditions of the part corresponding to the ETL side. In conclusion, it was confirmed that when the B-EBL is 3 nm, the efficiency and driving voltage characteristics are inferior. On the other hand, when the B-EBL is increased to 7 nm, the driving voltage is rather lowered, but the efficiency is somewhat reduced. Therefore, the 5 nm condition for the B-EBL was judged to be the most suitable condition for this BCL approach. In other words, if we use about 5 nm of DIC-TRZ material (B-EBL) and about 20 nm of TPBI, electrons are supplied to the green EML coated on the bottom without any problem, resulting in efficient light emission.
Fig. 8 (a) J–V–L curves, (b) current efficiency (CE) and luminance results, (c) EL spectra property, and (d) dependence of operating voltage and CE on the thickness of B-EBL at 1000 cd m−2. |
Therefore, we selected the thickness that fits the conditions of device C on the top of the EML and optimized the thickness of the materials on the bottom of the EML. In other words, for this experiment, the control device was manufactured in the structure of device C. Besides, device E and device F were additionally prepared as follows:
ITO/AQ1200/x-HTL/PCIC:Ir(mppy)3 (5 mol%)/B-EBL/BH:BD/TPBI/LiF/Al (where thickness of AQ1200 and x-HTL: 30 nm and 25 nm for device E; 20 nm and 35 nm for device F) (device configurations, including thickness conditions, are given in ESI†).
The J–V–L characteristics of those fabricated devices (devices C, E and F) were shown in Fig. 9(a). The Von at a constant luminance of 1 cd m−2 were measured to 4.5 V for devices C, E and F. And the Vop obtained at a luminance of 1000 cd m−2 were 7.9 V, 7.8 V and 7.4 V for devices C, E and F, respectively. Likewise, the current density at 6.0 V was 0.22 mA cm−2 for device C, 0.67 mA cm−2 for device E, and 0.15 mA cm−2 for device F, respectively. As mentioned earlier, the current density tended to diminish as the ETL thickness decreases. However, we found that J does not decrease according to the increase in HTL thickness. The J was increased from 0.15 mA cm−2 to 0.67 mA cm−2 when the thickness of HTL was changed from 15 nm to 25 nm. This might be due to a shift of the recombination zone toward the central region of EML (device F) from the ETL side (device C). However, J decreased rapidly when the thickness of the HTL increased to 35 nm presumably due to an additional shift of recombination zone toward the HTL side, which is an interface mixing zone. Meanwhile, at a constant luminance of 1000 cd m−2, the current efficiency was recorded to 60.3 cd A−1 for device C, 66.2 cd A−1 for device E and 53.5 cd A−1 for device F, respectively. This can be said to be consistent with the analogy of the change according to the movement of the recombination zone mentioned above. Besides, the maximum current efficiency values were measured as 60.3 cd A−1 for device C, 66.4 cd A−1 for device E and 54.4 cd A−1 for device F, respectively, as summarized in Table 5. Meanwhile, a different trend was observed in the operational stability (Fig. 9(d)), the best half-lifetime we obtained here is 1300 hours for device C, while the half-lifetime for device E and device F is 420 hours and 26 hours, respectively. A very interesting fact is also shown in Fig. 9(c). In other words, as the thickness of HTL increases from device C to device E or device F, it can be seen that the secondary peak intensity of the EL spectrum is increasing. This change is the optical effect caused by the recombination zone shift. Indeed, it is well-known that as the secondary peak intensity increases, the recombination zone gets closer to the HTL (Fig. 9).26
To find out why there is a difference between the conditions of high efficiency and long lifetime, we prepared hole only devices (HOD) and electron only devices (EOD) as shown in Fig. 10. The detailed configuration of the devices fabricated is as follows:
Fig. 10 (a) HOD and EOD energy diagram consisting of pure PCIC and PCIC:Ir(mppy)3 (5 mol%), (b) J–V characteristics of HOD A, HOD B, EOD A and EOD B, (c) data results of J (HOD A)/J (EOD A) and J (HOD B)/J (EOD B) (device configuration with a thickness information is shown in the ESI†). |
ITO/PEDOT:PSS/PCIC (HOD A) & PCIC:Ir(mppy)3 (5 mol%, HOD B)/MoO3/Al.
ITO/ZnO/PEI/PCIC (EOD A) & PCIC:Ir(mppy)3 (5 mol%, EOD B)/Al.
Considering the electrical energy characteristics of PCIC, the energy levels of HOMO and LUMO are shallow, so it is expected that the injection of holes is basically easier than the injection of electrons. We fabricated two HODs and two EODs devices to more accurately find out this characteristic. The difference in HOMO energy level between PCIC and Ir(mppy)3 is only 0.2 eV, but since the hole cannot be injected only into the host and can be trapped by the dopant, the hole current density is reduced by about 1 order of magnitude by doping. On the other hand, although the LUMO energy level of PCIC is too shallow, the difference in LUMO energy from the dopant remains as much as 0.5 eV, but the decrease in current density due to electron trapping was not observed significantly. As a result, the host material itself is expected to have a hole mobility value of about 60 times greater than that of electron mobility, but in the case of a doped device, the hole mobility value is expected to be about 2 times greater than that of electron mobility (Fig. 10).
Therefore, when the thickness of the X-HTL is thin, the resistance of the cross-linked X-HTL is relatively small, and the recombination zone can be skewed toward the B-EBL, and the area can become very narrow so that the efficiency decreases (device C). On the other hand, if we can move it toward the HTL side, the recombination zone can be widened and the efficiency can be increased. In fact, the efficiency was the highest under the conditions of device E. But, the efficiency decreased again in device F. However, unlike efficiency, in the case of a lifetime, it was found that device C, which has the recombination zone furthest from the HTL and EML interface mixing zone, is the most advantageous for operational stability (e.g. device lifetime). This fact has been proven in several previously published papers,27–29 and it can be seen that the trend is in good agreement with such previous results. Fig. 11 shows a schematic diagram of the location according to the recombination zone.
In general, a hole-transporting host materials are known to have very low anionic state stability, so the lifetime of device E and device F, where electrons reach deeper, is a disadvantage. To find out the actual bond dissociation energy (BDE) value of PCIC, the molecular simulation was performed and summarized in Table 6 (Fig. 12).
Area | Neutral | Anion | Cation |
---|---|---|---|
1 | 3.59 | 1.20 | 3.65 |
3.15 | 5.84 | ||
2 | 3.65 | 1.37 | 3.83 |
3.21 | 5.91 |
We confirmed that the anion state of PCIC is very weak through the BDE calculation result, whereas the cation state has a relatively strong bond (Table 6). In the case of device F, as electrons move toward the HTL and EML interface, the width of electrons exposed to the host increases. Therefore, deterioration occurs due to the weak anion state, resulting in reduced efficiency and lifetime. However, for device C, the thin HTL thickness results in the formation of recombination zone at the EML and B-EBL interfaces. The relatively strong cation state is stable against deterioration and so, in the case of device C, the lifetime characteristic is 50 times higher than that of device F, and it showed a stable characteristic. In fact, as mentioned earlier, due to the shallow LUMO of PCIC and the deep electron trapping level of the green dopant, it will be possible to keep the recombination zone close to the B-EBL side, and we expect that there will be room for further improvement in a lifetime.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1ra04855d |
This journal is © The Royal Society of Chemistry 2021 |