Next Article in Journal
The Functional Characterization of a Site-Specific Apigenin 4′-O-methyltransferase Synthesized by the Liverwort Species Plagiochasma appendiculatum
Previous Article in Journal
Multi-Anti-Parasitic Activity of Arylidene Ketones and Thiazolidene Hydrazines against Trypanosoma cruzi and Leishmania spp.
Previous Article in Special Issue
Spirobifluorene Core-Based Novel Hole Transporting Materials for Red Phosphorescence OLEDs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tuning the Geometrical Structures and Optical Properties of Blue-Emitting Iridium(III) Complexes through Dimethylamine Substitutions: A Theoretical Study

1
Low Carbon Energy Institute, School of Chemical Engineering & Technology, China University of Mining &Technology, Xuzhou 221008, China
2
Fukui Institute for Fundamental Chemistry, Kyoto University, 34-4 Takano Nishihiraki-cho, Sakyo, Kyoto 606-8103, Japan
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(5), 758; https://doi.org/10.3390/molecules22050758
Submission received: 29 March 2017 / Revised: 30 April 2017 / Accepted: 5 May 2017 / Published: 7 May 2017
(This article belongs to the Special Issue Organic Light Emitting Diodes)

Abstract

:
The geometrical structures and photophysical properties of Ir(4,6-dFppy)2(pic) (FIrpic) and its derivative (o-FIr, m-FIr, p-FIr) with dimethylamine substituted at the picolinic acid (NO) ligand were fully investigated by density functional theory and time-dependent density functional theory. The simulated electronic structure, as well as absorption and emission spectra of FIrpic are in good agreement with the experimental observations. The introduction of dimethylamine at the NO ligand at different positions is beneficial to extend the π-electron delocalization, increase HOMO energy levels, and hence improve the hole injection and transfer ability compared with those of FIrpic. Furthermore, o-FIr, m-FIr, and p-FIr have large absorption intensity and participation of metal-to-ligand charge transfer (MLCT) contribution in the main absorption spectra, which would be useful to improve the intersystem crossing (ISC) from the singlet to triplet excited state. More importantly, the high quantum yield of o-FIr (which is explained based on the detailed analysis of triplet energy, ET1), participation of 3MLCT contribution in the phosphorescent spectra, and energy difference between 3MLCT and triplet metal centered (3MC) d-d excited state compared with m-FIr and p-FIr indicate that o-FIr is expected to be an excellent blue phosphorescence emitter with high efficiency.

1. Introduction

Phosphorescent transition metal materials such as Re(I), Ru(II), and Ir(III) complexes are of great interest [1,2,3,4], because these materials have been extensively investigated for optoelectronic and microelectronic applications, such as dye-sensitized solar cells (DSSCs) and organic light-emitting diodes (OLEDs) [5,6,7]. Among these emitters, Ir(4,6-dFppy)2(pic) (FIrpic) is one of the most well-known blue phosphorescent emitters, which is widely used as a dopant in the electroluminescent layers of OLEDs [8,9,10]. However, the quantum yield and equilibrium between the hole and electron transportation of FIrpic are considered to be difficult for its further application. The development of high-performance blue-emitting materials is strongly required for the fabrication of OLEDs. Unfortunately, it is a challenging task to achieve stable and high-efficiency blue phosphorescent emitters because the blue phosphorescence complex has a high emission state compared with other color complexes. This enables thermal activation to triplet metal centered (3MC) d-d state, which is one of the main nonradiative processes from the emissive excited state for Ir(III) complexes [11,12].
In OLEDs, researchers can control the emission color and photophysical properties of emitters by a modification of the chelating ligands [13,14,15,16,17]. The common approach in tuning the photophysical properties of FIrpic is to modify the chemical structures of 2-(2,4-difluorophenyl) pyridine (CN) ligands. Baranoff et al. [18] replaced the F atoms on FIrpic by H, Cl, and Br. They found that the large halogen substituent contributed to sizable distortions of specific (CN) ligands, which were likely to play a role in the emissive and nonradiative properties when coupled with the heavy-atom effect. Owing to the introduction of the strong electron withdrawing perfluoro carbonyl group on (CN) ligands of FIrpic, the newly designed Ir complexes displayed deepest blue emissions and considerably high external quantum efficiencies (EQEs) [19]. Li et al. [20] investigated a series of Ir(III) complexes by introducing tert-butyl, n-heptyl, 3-ethylheptyl, and 2,4,6-trimethylphenyl substituted groups into the (CN) ligands of FIrpic. They revealed that both the substituted group and position had an effect on tuning the emission energies and quantum yield. However, the important chelating picolinic acid (NO) ligand of FIrpic has not received as much attention, to the best our knowledge. Herein, we will theoretically design and investigate a type of Ir complexes o-FIr, m-FIr, p-FIr (Figure 1) by introducing the dimethylamine substituted at the picolinic acid (NO) ligand, because the dimethylamine is one of the most popular functional groups used as hole transfer materials. We perform detailed theoretical calculations on the electronic structure, absorption and emission properties, quantum efficiency, as well as charge transport properties, aiming at revealing the relationship between the structures and photophysical properties. This work is expected to give some insights into the design and synthesis of blue emitters with high quantum efficiency.

2. Results and Discussion

2.1. The Optimized Geometries in the Ground and Lowest Lying Triplet Excited States

The schematic structures of FIrpic, o-FIr, m-FIr, and p-FIr are depicted in Figure 1. The ground-state (S0) geometry of FIrpic was optimized by B3LYP, CAMB3LYP, M06L, MPW1PW91 functionals with the same basis set. The detailed geometries and available crystal data of FIrpic [21] are collected in Table S1 in supporting information. It is found that M06L functional predicts more satisfied geometrical structures by compared with the experimental data. Therefore, the S0 geometries of o-FIr, m-FIr, p-FIr are calculated by M06L functional, and the optimized geometries as well as the numbering of some key atoms are depicted in Figure 2. The optimized primary metal–ligand bond lengths of these designed complexes in the ground-state (S0) geometry are collected in Table 1.
As listed in Table 1, the N1-Ir-N2 and O1-Ir-N2 of FIrpic are ca. 90°, and the dihedral angle of N1-O1-C2-C1 of FIrpic is 6.8°. Obviously, FIrpic has distorted octahedral geometry around the Ir(Ш) atom. Table 1 shows that the N1-Ir-N2, O1-Ir-N2, and N1-O1-C2-C1 of o-FIr, m-FIr, and p-FIr are similar to those of FIrpic. Clearly, the introduction of a dimethylamine moiety on the pyridine moiety of picolinic acid (NO) does not affect the distorted octahedral geometry of the Ir(Ш) complex. For o-FIr and p-FIr, because of the introduction of dimethylamine group on the pyridine, the Ir-N1 and Ir-O1 of o-FIr and p-FIr are shortened (0.005 Å and 0.01 Å for o-FIr, while 0.05 Å and 0.016 Å for p-FIr) in comparison to those of FIrpic. These calculated results indicate that there is a strong interaction between the Ir(Ш) and (NO) ligand. Furthermore, the Ir-N1 bond length of o-FIr, m-FIr, and p-FIr are significantly shortened compared with that of FIrpic; therefore, the possible of dissociation of Ir-N1 bond in the 3MC state is expected to be decreased by the incorporation of N(CH3)2 groups.
The lowest lying triplet excited states (T1) of studied complexes are also listed in Table 1. As listed in Table 1, the Ir-N1 bond length of these complexes is dramatically elongated in the T1 state in comparison with that in S0 state. This suggests that the interaction between the NO ligand and the metal is weakened in the T1 state. From the S0 to the T1 state, the Ir-C1/2 and Ir-N2/3 bond lengths of 2-(2,4-difluorophenyl)pyridine (CN) ligands in o-FIr, m-FIr, and p-FIr are contracted by the substitutions of the dimethylamine groups on the pyridine moiety. The strengthened metal–ligand bonds will strength the interaction between the metal and CN ligands and further benefit the charge transfer from the metal to ligand, which may improve the quantum efficiency.

2.2. Froniter Molecular Orbtials

Considering that frontier molecular orbitals are useful to provide the nature of the charge transfer and photophysical properties in the excited state, the HOMO and LUMO energies, HOMO–LUMO energy gaps, and the contour plots of these orbitals of studied complexes are drawn in Figure 3. Detailed descriptions of the frontier molecular compositions are collected in Table S2 in the supporting information.
Normally, when the interfacial vacuum energy shift can be neglected, the hole (or electron) injection energy barrier from the electrode to semiconductor (Φh) (or Φe) can be calculated by Φh = |HOMO| − Φm (or Φe = Φm − |LUMO|), where Φm is the work function of the electrode [22]. Consider that Au (Φm = 5.1 eV) is widely used as electrode in OLEDs; therefore, it is used as reference electrode [22]. As depicted in Figure 3, the HOMO and LUMO energies of o-FIr, p-FIr, and m-FIr increase remarkably (−4.65 eV and −2.15 eV for o-FIr, −4.59 eV and −2.07 eV for m-FIr, −4.65 eV and −2.16 eV for p-FIr) compared with FIrpic (−4.80 eV and −2.30 eV). Therefore, the calculated Φh of these complexes are (in the region of 0.30–0.51 eV) much smaller than those of the calculated Φh values (in the region of 2.80–3.03 eV). It is obvious that these emitters are suitable for use as p-type organic semiconductors [23]. Furthermore, the HOMO energy level of o-FIr (0.15 eV), m-FIr (0.21 eV), and p-FIr (0.15 eV) is significantly higher than that of FIrpic. The higher HOMO energies will benefit the hole injection abilities. Together with the increasing HOMO and LUMO energy, the HOMO–LUMO energy gaps of our designed molecules are similar to that of FIrpic, which means that absorption and phosphorescent spectra of these designed complexes might be similar to that of FIrpic when these spectra are predominantly contributed by the HOMO→LUMO transition.
For FIrpic, o-FIr, p-FIr, and m-FIr, the electron densities of HOMOs are mainly located on 50% d(Ir) orbital and 38% π(CN ligands) orbital, as shown in Figure 3 and Table S2. Owing to the introduction of N(CH3)2 groups, the LUMOs in o-FIr, m-FIr, and p-FIr are mainly focused on the CN ligands with about 90% compositions, while the LUMO in FIrpic is contributed by NO ligand. The HOMO-1 of these studied complexes is mainly focused on the ca. 50% d orbital the ca. 30% π orbitals of NO ligand, while theirs HOMO-2 are mainly contributed by NO ligand. The contributions of d(Ir) orbital on HOMO-3 of o-FIr (67.5%) and m-FIr (56.9%) are similar to that of FIrpic (67.1%). This means that the N(CH3)2 substitutions have a minor effect on the composition of the frontier molecular orbital. Therefore, the nature of electron transition upon excitations of o-FIr, p-FIr, and m-FIr might be similar with those of FIrpic, which will be further investigated in the following section.

2.3. Absorption Spectra

The calculated absorption spectra of studied complexes associated with oscillator strengths (f) and the corresponding compositions, as well as the experimental data [24] are given in Table 2. Simulated absorption spectra of these complexes with the data calculated at the TD-M06L method are drawn in Figure 4.
As listed in Table 2, the main calculated values of these bands are localized at 453.00, 380.95, and 265.86 nm for FIrpic, respectively, which are in good agreement with the experimental results of 458, 379, and 256 nm, respectively [24]. For the low-lying absorption spectra of FIrpic, the excitations of HOMO→LUMO + 1 and HOMO − 1→LUMO + 2 are dominantly responsible for the absorption peaks at 453.00 and 380.95 nm, respectively. As listed in Table S2, the HOMO and HOMO − 1 are delocalized on the Ir(d), NO, and NC ligand, while the LUMO + 1 and LUMO + 2 are mainly contributed by the NC ligand. Therefore, from the analysis of FMOs, the absorption peaks at 453.00 and 380.95 nm can be characterized as metal-to-ligand charge transfer (MLCT), inter-ligand charge transfer (ILCT), and ligand–ligand charge transfer (LLCT). The absorption peak with largest oscillator strength (f = 0.2104) of FIrpic is located at 265.86 nm, which is arising from the HOMO − 8→LUMO + 2 and HOMO→LUMO + 7 transitions. Because of the large of amount of Ir(d) in HOMO (50.7%) and HOMO – 8 (20.8%), the absorption peak at 265.86 nm can mainly be assigned to the MLCT/LLCT/ILCT character. From the analysis of frontier molecule orbital compositions, the mount of MLCT of the 453.00, 380.95, and 265.86 for FIrpic is 21.07%, 21.07%, and 7.16%, respectively. The significant MLCT contributions on the lower-lying absorptions of FIrpic are in good agreement with the experimental results [24].
Table 2 also shows the lowest-lying absorption peak of o-FIr, m-FIr, and p-FIr at 463.87, 462.95, and 463.29 nm, respectively, which is dominantly responsible for the excitation of HOMO→LUMO. Table S2 shows that the HOMO of o-FIr, m-FIr, and p-FIr is composed of 50.1% Ir(d) + 47.7% π(NO + NC), 50.3% Ir(d) + 47.9% π(NO + NC), 48.6% Ir(d) + 49.3% π(NO + NC), respectively, while their LUMOs are contributed by ca. 80% π*(NC). Thus, the lowest-lying absorptions of these complexes originate from the [Ir(d) + π(NO + NC)]→π*(NC) excited state with MLCT/ILCT/LLCT character. The mount of MLCT of the lowest-lying absorption of o-FIr, m-FIr, and p-FIr is ca.22%, which is similar to that of FIrpic. Furthermore, as shown in Figure 4, the absorption intensity of N(CH3)2-substituted Ir complexes (m-FIr, o-FIr, p-FIr) in the highest energy absorption spectra have been effectively enhanced, which may increase the intersystem crossing rates (ISC). For o-FIr, the highest energy absorption peak (262.33 nm) is contributed by the mixture excitations of HOMO − 9→LUMO + 2, HOMO − 5→LUMO + 5, HOMO − 4→LUMO + 5, HOMO − 2→LUMO + 5, HOMO − 1→ LUMO + 5. Since the Ir(d) in the HOMO − 9, HOMO − 4, HOMO is 22.7%, 67.5%, 50.1%, respectively, the MLCT contributed 8.9% to the 262.33 nm of o-FIr, which is slightly larger than that of FIrpic. For m-FIr and p-FIr, the absorption peak with large oscillator strength is located at 267.06 (attributed by HOMO − 8→LUMO + 2 and HOMO→LUMO + 7) and 271.08 nm (attributed by HOMO − 7→LUMO + 4, HOMO − 3→LUMO + 5, HOMO − 2→LUMO + 5, HOMO→LUMO + 7), respectively. Combining the transition configuration with the orbital composition, the mount of MLCT of the high energy absorption peaks for m-FIr and p-FIr are only 4.4% and 5.2%, respectively. Clearly, the contribution of MLCT in the whole absorption spectra of o-FIr are remarkably larger than those of FIrpic, m-FIr, and p-FIr. Normally, the participation of MLCT in the absorption spectra is a critical factor affecting the intensity of the singlet–triplet transitions, and thus improves the quantum efficiency. Thus, the calculated results indicate that the introduction of N(CH3)2 groups at the ortho-position of NO ligand is useful to enhance the intersystem crossing and singlet–triplet transitions, and therefore the possibility of the phosphorescent quantum yield might be enhanced.

2.4. Phosphorescence Emission Spectra

Based on the optimized lowest triplet excited state (T1) geometries, vertical T1–S0 transition energies (ΔEvert) and the emission wavelength (nm) obtained by CAMB3LYP functional are collected in Table 3, together with available experimental values [24]. Since the vertical T1–S0 transition energy (ΔEvert) is the electronic energy difference between the T1 and S0 state at the optimized T1 geometry, it is therefore strongly dependent on the functional. The molecular orbital compositions responsible for the emission spectra of the studied complex are given in Table S3. The 0-0 T1→S0 transition energy (ΔE0-0) is calculated by the energy difference between the S0 and T1 states at their respective optimized geometries obtained by M06 functional.
As listed in Table 3, the calculated emission peak for FIrpic based on the calculated Evert and E0-0 method is located at 490.49 and 480.86 nm, respectively. The slightly longer wavelength derived from Evert is caused by only one single configuration, while real emission is composed of more than one configuration. The phosphorescence at 490.49 nm of FIrpic is contributed by the excitation of HOMO→LUMO. According to the orbital composition analyses, the HOMO is contributed by 42.8% Ir(d) + 54.3% π(CN + NO), while the LUMO is mainly focused on CN ligand. Accordingly, the emission of FIrpic can be attributed to 3MLCT/3LLCT/3ILCT character. The d(Ir) orbital contributes to 42.8% for HOMO, and thus the emission of FIrpic at 490.49 nm has 3MLCT (20.31%) transition character. For N(CH3)2-substituted Ir complexes, the phosphorescent emissions based on Evert method are located at 492.78, 484.32, and 485.83 nm for o-FIr, m-FIr, and p-FIr, respectively, which is dominantly responsible for excitations of HOMO→LUMO and HOMO − 4→LUMO. This is because the HOMO and HOMO − 4 of o-FIr, m-FIr, and p-FIr are mainly contributed by d(Ir) and π(CN + NO) (besides the HOMO − 4 of m-FIr), while the LUMOs of o-FIr and m-FIr are located at π*(CN). Table 3 shows that the calculated E0–0 for these designed complexes are ca. 480nm. Accordingly, these designed complexes have similar phosphorescent emissions and transition character with that of FIrpic, which might be a potential candidate for blue-emitting materials. Furthermore, the d(Ir) orbital contributions to HOMO and HOMO − 4 (42.8% and 0.97% for m-FIr, 30.68% and 36.68% for p-FIr) are much smaller than that of HOMO (41.35%) and HOMO − 4 (63.39%) for o-FIr. The larger d(Ir) orbital composition results in slightly larger 3MLCT% in its phosphorescence of o-FIr (18.78%) than m-FIr (9.11%) and p-FIr (11.56%). Therefore, the o-FIr may have improved Φp compared with m-FIr and p-FIr.
To give some insight into the nonradiative rate constant (knr), the high lying metal-centered (3MC d-d) excited states of studied complexes—which is one of the most important deactivation pathways for the phosphoresce emission from T1 in transition-metal complexes—were optimized by unrestricted M06L method. The energy difference between 3MLCT/π-π* and 3MC d-d excited states are drawn in Figure 5, along with the spin density distribution in the 3MC d–d state. As shown in Figure 5, the 3MC d-d state of studied complexes are lying at a relatively higher energy than that of 3MLCT/π-π* state, which suggests that the possibility of a radiationless pathway is low. Furthermore, the energy difference between the 3MLCT/π-π* and 3MC d-d excited states of o-FIr (0.30 eV) and m-FIr (0.30 eV) is larger than that of FIrpic (0.24 eV). Therefore, these calculation results show that ortho- and meta- substituted Ir complexes are useful to decrease the Knr. Based on the above viewpoint, it becomes clear that introducing the N(CH3)2 at the ortho-position of NO ligand in FIrpic led to larger ISC and kr values, as well as the smaller knr value relative to others. Therefore, o-FIr is considered to be used as potential blue-emitting emits with high quantum efficiency.

2.5. Performance in OLEDs

To evaluate the performance of OLED devices, it is highly important to investigate the charge injection and transfer ability. The ionization potential (IP) and electron affinity (EA) are usually employed to evaluate the energy barrier for the hole and electron injections. Normally, the lower IP (larger EA) value suggests that it is easier to inject hole (electron). According to the Marcus theory [25], the electron transfer rate constants can be expressed by the following formula:
k = ( π λ k B T ) 1 / 2 V 2 exp ( λ 4 k B T )
where T and kB is the temperature and Boltzmann constant, respectively. The λ is the reorganization energy, and V is the coupling matrix element. According to Equation (1), the intermolecular charge transfer rate k is determined by two factors, λ and V. Due to the limited intermolecular charge transfer range in the solid state, the mobility of charges has been demonstrated to be primarily associated with the reorganization energy λ for OLED materials [26]. The detailed description of the calculations of vertical ionization potential (IP(v)), vertical electron affinity (EA(v)), hole extraction potential, (HEP), electron extraction potential (EEP) and electron reorganization energy (λelectron), as well as hole reorganization energy (λhole) are listed in our previous report [16]. According to the previous report [16], the IP, EA, and λ of studied complexes are calculated and listed in Table 4.
Compared with FIrpic, all these designed complexes have smaller IP(v) and HEP, which is consistent with their higher HOMO energy levels. Thus, the hole injection ability of these designed complexes are improved by introduction of dimethylamine substitutions at the CO ligand. However, o-FIr, m-FIr, and p-FIr have smaller EA values than FIrpic, which will lead to worse electron injection ability compared with FIrpic. According to Equation (1), the λ is required to be small for efficient charge transport. As listed in Table 4, the λelectron of studied complexes are slightly smaller than that of FIrpic, indicating that these complexes have excellent electron transfer ability. The difference between the λhole and λelectron of FIrpic, o-FIr, and p-FIr is less than 0.05 eV. Hence, it is expected that the FIrpic, o-FIr, and p-FIr might have excellent charge transfer balance, which is one of the key factors affecting the performance of OLEDs.

3. Computational Details

The geometrical structures of the ground state (S0) of FIrpic were fully optimized by B3LYP [27,28], CAMB3LYP [29], M06L [30], and mPW1PW91 [31] methods, respectively, combined with 6-31G(d) basis set for the non-metal atoms (H, C, N, O, and F) and a “double-j” quality basis set LANL2DZ for Ir atom. The calculated vibrational frequencies with no imaginary frequencies based on the optimized geometries verify that each of the optimized structures was a true minimum on the potential energy surface. The calculated results are collected in Table S1 in the supporting information, together with the crystal data [21]. As depicted in Table S1, the bond lengths obtained for B3LYP and CAMBLYP overestimated the bond lengths, especially Ir-N1 (0.024 Å for CAMB3LYP and 0.031 Å for B3LYP) and Ir-N3 (0.049 Å for CAMB3LYP and 0.073 Å for B3LYP), while the bond length differences between the calculated results obtained by M06L and crystal data were small. Therefore, the ground-state geometries of the studied complexes were calculated by M06L method. Based on the ground-optimized structure, the vertical excitation energies were calculated by TD-M06L method with continuum polarizable continuum model (CPCM). The lowest-lying triplet excited state geometries (T1) and triplet-centered (3MC) d-d states were optimized by unrestricted M06L method. Based on the T1 equilibrium geometries, four different functionals—PBE0, B3LYP, CAMB3LYP, and M06L—were performed on the emission spectra calculations (as listed in Table S4). The CAMB3LYP functional was adopted to predict the emission spectra of these complexes, because the calculated results obtained by CAMB3LYP were consistent with the experimental results. All calculations were performed with the Gaussian 09 software package [32].

4. Conclusions

In this work, the geometrical and electronic structures, absorption and emission spectra, as well as phosphorescence efficiencies for a series of dimethylamine-substituted iridium(Ш) complexes FIrpic have been investigated by density functional theory (DFT) and time-dependent density functional theory (TDDFT) methods. The results indicate that the introduction of dimethylamine is useful to extend the π-electron delocalization, increase HOMO energy levels, and enhance hole injection and transfer ability. Furthermore, the absorption intensity and participation of MLCT of these dimethylamine substitutions are significantly stronger than that of FIrpic, suggesting the possibility of singlet-to-triplet transitions are enhanced, and therefore the phosphorescent quantum yield might be increased by introducing dimethylamine. Importantly, through the analyses of triplet energy (ET1), participation of 3MLCT in phosphorescent spectra, as well as the energy difference between the 3MLCT/π-π and 3MC d-d state, it can be inferred that o-FIr has larger kr values and smaller knr values compared with other molecules. So, the strategy of introducing the dimethylamine in NO ligand of FIrpic should be useful in the design of more-efficient blue emitters.

Supplementary Materials

Supplementary materials are available on line.

Acknowledgments

Financial supports from the Fundamental Research Funds for the Central Universities (No. 2015XKMS042). We are grateful to the High Performance Computing Center of China University of Mining and Technology for the award of CPU hours to accomplish this work.

Author Contributions

Hong-Qu Tang performed the calculation work; Guo-Jun Kang and Xue-Feng Ren analyzed the data and wrote the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lin, C.-H.; Chiu, Y.-C.; Chi, Y.; Tao, Y.-T.; Liao, L.-S.; Tseng, M.-R.; Lee, G.-H. Mechanistic Investigation of Improved Syntheses of Iridium(III)-Based OLED Phosphors. Organometallics 2012, 31, 4349–4355. [Google Scholar] [CrossRef]
  2. Xia, H.; Zhu, Y.; Lu, D.; Li, M.; Zhang, C.; Yang, B.; Ma, Y. Ruthenium(II) Complexes with the Mixed Ligands 2,2′-Bipyridine and 4,4′-Dialkyl Ester-2,2′-bipyridine as Pure Red Dopants for a Single-Layer Electrophosphorescent Device. J. Phys. Chem. 2006, 110, 18718–18723. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, T.-T.; Jia, J.-F.; Ren, Y.; Wu, H.-S. Ligand Effects on Structures and Spectroscopic Properties of Pyridine-2-aldoxime Complexes of Re(CO)3+: DFT/TDDFT Theoretical Studies. J. Phys. Chem. 2011, 115, 3174–3181. [Google Scholar] [CrossRef] [PubMed]
  4. Ho, C.-L.; Wong, W.-Y. Small-molecular blue phosphorescent dyes for organic light-emitting devices. New J. Chem. 2013, 37, 1665–1683. [Google Scholar] [CrossRef]
  5. Won, D.-I.; Lee, J.-S.; Ji, J.-M.; Jung, W.-J.; Son, H.-J.; Pac, C.; Kang, S.O. Highly Robust Hybrid Photocatalyst for Carbon Dioxide Reduction: Tuning and Optimization of Catalytic Activities of Dye/TiO2/Re(I) Organic–Inorganic Ternary Systems. J. Am. Chem. Soc. 2015, 137, 13679–13690. [Google Scholar] [CrossRef] [PubMed]
  6. Lamansky, S.; Djurovich, P.; Murphy, D.; Abdel-Razzaq, F.; Lee, H.-E.; Adachi, C.; Burrows, P.E.; Forrest, S.R.; Thompson, M.E. Highly Phosphorescent Bis-Cyclometalated Iridium Complexes:  Synthesis, Photophysical Characterization, and Use in Organic Light Emitting Diodes. J. Am. Chem. Soc. 2001, 123, 4304–4312. [Google Scholar] [CrossRef] [PubMed]
  7. Velmurugan, G.; Ramamoorthi, B.K.; Venuvanalingam, P. Are Re(I) phenanthroline complexes suitable candidates for OLEDs? Answers from DFT and TD-DFT investigations. Phys. Chem. Chem. Phys. 2014, 16, 21157–21171. [Google Scholar] [CrossRef] [PubMed]
  8. Rausch, A.F.; Thompson, M.E.; Yersin, H. Matrix Effects on the Triplet State of the OLED Emitter Ir(4,6-dFppy)2(pic) (FIrpic): Investigations by High-Resolution Optical Spectroscopy. Inorg. Chem. 2009, 48, 1928–1937. [Google Scholar] [CrossRef] [PubMed]
  9. Kozhevnikov, V.N.; Zheng, Y.H.; Clough, M.; Al-Attar, H.A.; Griffiths, G.C.; Abdullah, K.; Raisys, S.; Jankus, V.; Bryce, M.R.; Monkman, A.P. Cyclometalated Ir(III) Complexes for High-Efficiency Solution-Processable Blue PhOLEDs. Chem. Mater. 2013, 25, 2352–2358. [Google Scholar] [CrossRef]
  10. Rausch, A.F.; Thompson, M.E.; Yersin, H. Blue Light Emitting Ir(III) Compounds for OLEDs - New Insights into Ancillary Ligand Effects on the Emitting Triplet State. J. Phys. Chem. 2009, 113, 5927–5932. [Google Scholar] [CrossRef] [PubMed]
  11. Sajoto, T.; Djurovich, P.I.; Tamayo, A.B.; Oxgaard, J.; Goddard, W.A.; Thompson, M.E. Temperature Dependence of Blue Phosphorescent Cyclometalated Ir(III) Complexes. J. Am. Chem. Soc. 2009, 131, 9813–9822. [Google Scholar] [CrossRef] [PubMed]
  12. Yi, S.; Kim, J.-H.; Cho, Y.-J.; Lee, J.; Choi, T.-S.; Cho, D.W.; Pac, C.; Han, W.-S.; Son, H.-J.; Kang, S.O. Stable Blue Phosphorescence Iridium(III) Cyclometalated Complexes Prompted by Intramolecular Hydrogen Bond in Ancillary Ligand. Inorg. Chem 2016, 55, 3324–3331. [Google Scholar] [CrossRef] [PubMed]
  13. Stringer, B.D.; Quan, L.M.; Barnard, P.J.; Wilson, D.J.D.; Hogan, C.F. Iridium Complexes of N‑Heterocyclic Carbene Ligands: Investigation into the Energetic Requirements for Efficient Electrogenerated Chemiluminescence. Organometallics 2014, 33, 4860–4872. [Google Scholar] [CrossRef]
  14. Chu, W.-K.; Yiu, S.-M.; Ko, C.-C. Neutral Luminescent Bis(bipyridyl) Osmium(II) Complexes with Improved Phosphorescent Properties. Organometallics 2014, 33, 6771–6777. [Google Scholar] [CrossRef]
  15. You, Y.; Park, S.Y. Phosphorescent iridium(III) complexes: toward high phosphorescence quantum efficiency through ligand control. Dalton Trans. 2009, 1269, 1267–1282. [Google Scholar] [CrossRef]
  16. Kang, G.J.; Ren, X.F.; He, Q.Q. Theoretical study on effect of thiophene substitution on the structure and phosphorescence quantum yields of red-emitting iridium(III) emitters in OLEDs. J. Photochem. Photobiol. 2016, 319, 25–33. [Google Scholar] [CrossRef]
  17. Shang, X.H.; Han, D.M.; Zhou, D.F.; Zhang, G. Shedding light on the photophysical properties of iridium(III) complexes with a dicyclometalated phosphate ligand via N-substitution from a theoretical viewpoint. New J. Chem. 2017, 41, 1645–1652. [Google Scholar] [CrossRef]
  18. Baranoff, E.; Curchod, B.F.E.; Monti, F.; Steimer, F.; Accorsi, G.; Tavernelli, I.; Rothlisberger, U.; Scopelliti, R.; Grätzel, M.; Nazeeruddin, M.K. Influence of Halogen Atoms on a Homologous Series of Bis-Cyclometalated Iridium(III) Complexes. Inorg. Chem. 2012, 51, 799–811. [Google Scholar] [CrossRef] [PubMed]
  19. Lee, S.; Kim, S.-O.; Shin, H.; Yun, H.-J.; Yang, K.; Kwon, S.-K.; Kim, J.-J.; Kim, Y.-H. Deep-Blue Phosphorescence from Perfluoro Carbonyl-Substituted Iridium Complexes. J. Am. Chem. Soc. 2013, 135, 14321–14328. [Google Scholar] [CrossRef] [PubMed]
  20. Li, J.; Zhang, Q.; He, H.; Wang, L.; Zhang, J. Tuning the electronic and phosphorescence properties of blue-emitting iridium(III) complexes through different cyclometalated ligand substituents: A theoretical investigation. Dalton Trans. 2015, 44, 8577–8589. [Google Scholar] [CrossRef] [PubMed]
  21. Xu, M.-L.; Che, G.-B.; Li, X.-Y.; Xiao, Q. Bis[3,5-difluoro-2-(2-pyridyl)phenyl]-(picolinato)iridium(III). Acta Cryst. 2009, E65, m28. [Google Scholar] [CrossRef] [PubMed]
  22. Wen, Y.; Liu, Y. You have full text access to this content, Recent Progress in n-Channel Organic Thin-Film Transistors. Adv. Mater. 2010, 22, 1331–1345. [Google Scholar] [CrossRef] [PubMed]
  23. Zhang, S.-F.; Chen, X.-K.; Fan, J.-X.; Ren, A.-M. Charge transport properties in a series of five-ring-fused thienoacenes: A quantum chemistry and molecular mechanic study. Org. Electron. 2013, 14, 607–620. [Google Scholar] [CrossRef]
  24. Baranoff, E.; Curchod, B.F.E. FIrpic: Archetypal blue phosphorescent emitter for Electroluminescence. Dalton Trans. 2015, 44, 8318–8329. [Google Scholar] [CrossRef] [PubMed]
  25. Lin, B.C.; Cheng, C.P.; Lao, Z.P.M. Reorganization Energies in the Transports of Holes and Electrons in Organic Amines in Organic Electroluminescence Studied by Density Functional Theory. J. Phys. Chem. 2003, 107, 5241–5251. [Google Scholar] [CrossRef]
  26. Marcus, R.A. On the Theory of Oxidation-Reduction Reactions Involving Electron Transfer. J. Chem. Phys. 1956, 24, 966–978. [Google Scholar] [CrossRef]
  27. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. 1988, 37, 785–789. [Google Scholar] [CrossRef]
  28. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  29. Yanai, T.; Tew, D.P.; Handy, N.C. A new hybrid exchange–correlation functional using the Coulomb-attenuating method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef]
  30. Zhao, Y.; Truhlar, D.G. A new local density functional for main-group thermochemistry, transition metal bonding, thermochemical kinetics, and noncovalent interactions. J. Chem. Phys. 2006, 125, 194101–194118. [Google Scholar] [CrossRef] [PubMed]
  31. Adamo, C.; Barone, V. Exchange functionals with improved long-range behavior and adiabatic connection methods without adjustable parameters: The mPW and mPW1PW models. J. Chem. Phys. 1998, 108, 664–675. [Google Scholar] [CrossRef]
  32. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Revision B.04; Gaussian: Wallingford, CT, USA, 2009. [Google Scholar]
Sample Availability: Not available.
Figure 1. Schematic structures of the studied complexes.
Figure 1. Schematic structures of the studied complexes.
Molecules 22 00758 g001
Figure 2. Optimized structures of o-FIr, m-FIr, and p-FIr in the ground state, together with the number of some key atoms.
Figure 2. Optimized structures of o-FIr, m-FIr, and p-FIr in the ground state, together with the number of some key atoms.
Molecules 22 00758 g002
Figure 3. The orbital energies and contour plot of the HOMO and LUMO of studied complexes.
Figure 3. The orbital energies and contour plot of the HOMO and LUMO of studied complexes.
Molecules 22 00758 g003
Figure 4. Simulated absorption spectra of studied complexes with the data calculated at the TD-M06L level.
Figure 4. Simulated absorption spectra of studied complexes with the data calculated at the TD-M06L level.
Molecules 22 00758 g004
Figure 5. Energy level diagram of the studied complexes in the T1 and triplet metal centered (3MC) excited states, together with the contour plots of the spin density distribution in the 3MC state.
Figure 5. Energy level diagram of the studied complexes in the T1 and triplet metal centered (3MC) excited states, together with the contour plots of the spin density distribution in the 3MC state.
Molecules 22 00758 g005
Table 1. Calculated metal–ligand bond lengths (Å), bond angles (°), and dihedral angles (°) for studied complexes in the ground state (S0) and first excited triplet state(T1).
Table 1. Calculated metal–ligand bond lengths (Å), bond angles (°), and dihedral angles (°) for studied complexes in the ground state (S0) and first excited triplet state(T1).
MoleculeStateIr-N1Ir-O1Ir-N2Ir-C1Ir-N3Ir-C2N1-Ir-N2O1-Ir-N2N1-O1-C2-C1
FIrpicS02.2152.1902.0662.0012.0581.99788.493.56.8
T12.3132.1722.0691.9792.0601.97597.188.7−3.4
o-FIrS02.2102.1802.0662.0022.0541.99787.394.05.0
T12.3162.1542.0571.9812.0671.97784.493.73.1
m-FIrS02.2062.1932.0642.0022.0561.99588.693.07.2
T12.2852.1712.0691.9792.0571.97895.989.0−4.6
p-FIrS02.1652.1772.0552.0102.0452.00187.692.96.3
T12.3052.1602.0681.9822.0581.97797.488.7−3.0
Table 2. The calculated maximum absorption wavelength (nm), oscillator strengths (f), and major contribution transition composition for studied complexes obtained by TD-M06L method. FIrpic: Ir(4,6-dFppy)2(pic).
Table 2. The calculated maximum absorption wavelength (nm), oscillator strengths (f), and major contribution transition composition for studied complexes obtained by TD-M06L method. FIrpic: Ir(4,6-dFppy)2(pic).
MoleculeTransitionλ(nm)fCompositionCIExp. [24] (nm)
FIrpicS0→S2453.000.0250HOMO→LUMO + 10.68971455
S0→S7380.950.0614HOMO − 1→LUMO + 20.55936379
S0→S50265.860.2104HOMO − 8→LUMO + 20.38300256
HOMO→LUMO + 70.24359
o-FIrS0→S1463.870.0258HOMO→LUMO0.70168
S0→S7395.150.0266HOMO − 2→LUMO + 20.56244
HOMO − 1→LUMO + 2−0.40890
S0→S52269.040.1135HOMO − 9→LUMO + 10.25480
HOMO − 5→LUMO + 50.42176
HOMO − 1→LUMO + 70.23321
S0→S53268.660.1642HOMO − 7→LUMO + 4−0.35667
HOMO − 1→LUMO + 70.36355
S0→S54266.760.1318HOMO − 9→LUMO + 20.21920
HOMO − 2→LUMO + 70.26962
HOMO − 1→LUMO + 7−0.21838
HOMO→LUMO + 70.41903
S0→S55266.320.0836HOMO − 5→LUMO + 50.42068
S0→S56264.540.1094HOMO − 9→LUMO + 20.48856
HOMO − 8→LUMO + 2−0.19242
HOMO − 1→LUMO + 50.27865
S0→S57262.330.1716HOMO − 9→LUMO + 2−0.24371
HOMO − 5→LUMO + 5−0.24091
HOMO − 4→LUMO + 5−0.25407
HOMO − 2→LUMO + 50.20626
HOMO − 2→LUMO + 70.30782
HOMO − 1→LUMO + 50.32002
m-FIrS0→S1462.950.0263HOMO→LUMO0.70166
S0→S6393.550.0423HOMO − 2→LUMO−0.41398
HOMO − 2→LUMO + 10.42182
HOMO − 1→LUMO + 10.31148
S0→S53267.060.3060HOMO − 8→LUMO + 2−0.24031
HOMO→LUMO+70.25154
p-FIrS0→S1463.290.0286HOMO→LUMO0.70148
S0→S8388.020.0581HOMO − 1→LUMO + 1−0.43685
HOMO→LUMO + 30.47613
S0→S52271.080.2512HOMO − 7→LUMO + 4−0.27219
HOMO − 3→LUMO + 50.21898
HOMO − 2→LUMO + 50.29024
HOMO→LUMO + 7−0.21723
Table 3. The calculated phosphorescence emission wavelength (nm), along with metal–ligand charge transfer character (3MLCT, %) and available experimental values.
Table 3. The calculated phosphorescence emission wavelength (nm), along with metal–ligand charge transfer character (3MLCT, %) and available experimental values.
MoleculeEvertE0-0Exp [24]
λ (nm)E (eV)Composition3MLCT (%)λ (nm)E (eV)λ (nm)
FIrpic490.492.530.49 (HOMO→LUMO) 20.31480.862.58468
o-FIr492.782.520.45 (HOMO − 4→LUMO) 0.43 (HOMO→LUMO)18.78482.582.60
m-FIr484.322.56−0.36 (HOMO − 4→LUMO) 0.40 (HOMO→LUMO) 9.11481.782.57
p-FIr485.832.55−0.36 (HOMO − 4→LUMO) 0.49 (HOMO→LUMO) 11.56481.762.57
Table 4. The calculated ionization potential (IP, eV), electron affinities (EA, eV), hole extraction potential, (HEP, eV), electron extraction potential (EEP, eV) hole/electron reorganization energy (λholeelectron, eV).
Table 4. The calculated ionization potential (IP, eV), electron affinities (EA, eV), hole extraction potential, (HEP, eV), electron extraction potential (EEP, eV) hole/electron reorganization energy (λholeelectron, eV).
MoleculeIP(v)HEPEA(v)EEPλholeλelectron
FIrpic6.566.290.720.980.270.26
o-FIr6.326.060.480.690.260.21
m-FIr6.396.090.560.750.300.19
p-FIr6.346.110.530.730.230.20

Share and Cite

MDPI and ACS Style

Ren, X.-F.; Tang, H.-Q.; Kang, G.-J. Tuning the Geometrical Structures and Optical Properties of Blue-Emitting Iridium(III) Complexes through Dimethylamine Substitutions: A Theoretical Study. Molecules 2017, 22, 758. https://doi.org/10.3390/molecules22050758

AMA Style

Ren X-F, Tang H-Q, Kang G-J. Tuning the Geometrical Structures and Optical Properties of Blue-Emitting Iridium(III) Complexes through Dimethylamine Substitutions: A Theoretical Study. Molecules. 2017; 22(5):758. https://doi.org/10.3390/molecules22050758

Chicago/Turabian Style

Ren, Xue-Feng, Hong-Qu Tang, and Guo-Jun Kang. 2017. "Tuning the Geometrical Structures and Optical Properties of Blue-Emitting Iridium(III) Complexes through Dimethylamine Substitutions: A Theoretical Study" Molecules 22, no. 5: 758. https://doi.org/10.3390/molecules22050758

Article Metrics

Back to TopTop