An effective method in modulating thermally activated delayed fluorescence (TADF) emitters from green to blue emission: the role of the phenyl ring†
Received
20th November 2023
, Accepted 16th January 2024
First published on 16th January 2024
Abstract
Developing efficient blue emitters with high performance and low cost is crucial for the further development of organic light-emitting diodes (OLEDs). Based on the two experimentally reported green thermally activated delayed fluorescence (TADF) emitters, which are thioxanthone derivatives consisting of carbazole as an electron donor and 9H-thioxanthen-9-one-S,S-dioxide (SOXO) as an electron acceptor with donor–acceptor (D–A) or donor–acceptor–donor (D–A–D) structures, two new blue TADF emitters are designed by simply inserting a phenyl ring between D and A units. The TADF processes of the four thioxanthone derivatives are studied systematically through first-principles calculations. The role of the introduced phenyl ring in the excited state properties of the designed molecules is explored by analyzing the changes in molecular geometries, frontier molecular orbital distributions, the lowest singlet–triplet energy splitting (ΔEST), the spin orbit coupling (SOC) constants, the radiative decay rates (kr) and the nonradiative decay rates (knr), as well as the intersystem crossing rates (kISC) and reverse intersystem crossing rates (kRISC). The results show that when incorporating phenyl units into the D–A and D–A–D structures, both high kr and enhanced kRISC are achieved in Cz-Ph-SOXO and DCz-DPh-SOXO, demonstrating that incorporating the phenyl unit in D–A and D–A–D structures is an efficient way for developing new SOXO-based TADF molecules. It is worth noting that the kRISC values for Cz-Ph-SOXO and DCz-DPh-SOXO are significantly increased with respect to those of the experimental molecules. The present results would provide helpful guidelines for developing new SOXO-based TADF molecules experimentally.
1. Introduction
As the third generation of organic light emitting diode (OLED) emitters, thermally activated delayed fluorescence (TADF) materials, which can harvest both the singlet and triplet excitons through the efficient reverse intersystem crossing (RISC) process, have drawn more and more attention in recent years.1–3 Unlike the conventional fluorescent and phosphorescent materials, the TADF emitters can utilize all the singlet and triplet excitons and realize high efficiency without the assistance of noble metals. Thus, there is a high potentiality to develop lower cost, but more promising OLED devices based on this class of materials. For an efficient TADF emitter, two key parameters, a small singlet–triplet energy splitting (ΔEST) and high photoluminescence quantum yield (ΦPL), are requisite.4–7 To achieve a small ΔEST, the spatial separation of the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of the molecules are required and thereby the exchange integral (J) will be reduced based on the equation
.8,9 However, the demand of the separated distribution of the HOMO/LUMO for a narrow ΔEST can suppress the transition dipole moment between the ground state (S0) and the first singlet excited state (S1), leading to a small oscillator strength (f) and low ΦPL. Therefore, the delicate balance between the spatial separation of the HOMO–LUMO wave function and the large f could be a compromise design strategy for efficient TADF emitters.4–7 So far, versatile molecular systems with TADF properties have been developed, including spiro-acidine,10 triazine,11 spirobifluorene,12 phthalonitrile,13 triptycene,8 diphenyl sulfone derivatives,14–17 and so on. Highly efficient sky-blue OLEDs with 37% external quantum efficiency have been achieved, which can be comparable to the excellent performance of phosphorescent OLEDs.18 Despite significant progress having been made in recent years, developing highly efficient new TADF emitters with both small ΔEST and high ΦPL is still a great challenge. The traditional design strategy for TADF emitters is combining an electron donor (D) and an electron acceptor (A) unit to spatially separate the HOMOs and LUMOs.19,20 Generally, a phenyl linker is demanded to increase the D–A spatial separation distance and relieve the twisted angle between the D and the A.21,22 Compared with the D–A-type TADF emitters, D–A–D-type TADF emitters can facilitate a stronger intramolecular charge transfer and therefore smaller ΔEST, which can promote the highly efficient RISC progress and achieve excellent OLED performance.23–31
Thioxanthone (TXO) and 9-H-thioxanthen-9-one-10,10-dioxide (SOXO) units are excellent acceptor units for constructing highly efficient TADF materials. The ΔEST of SOXO is reported to be lower than 0.3 eV,32,33 and the SOXO core contains a ketone moiety which is alone able to produce delayed fluorescence.34,35 Several kinds of SOXO systems have been reported to exhibit excellent TADF properties. In 2014, Wang and co-workers reported two efficient TADF emitters of SOXO based derivatives with the D–A structure, using SOXO as the electron acceptor unit, and triphenylamine (TPA)/N-phenylcarbazole (PhCz) as the electron donor unit.36 Later on, they demonstrated a remarkable divergence in the photophysical properties of the SOXO based TADF emitters with different substitution positions of the PhCz donor.37 In 2019, they designed four D–A–D region isomers with TADF properties by connecting two PhCz as donors at the different substituted positions of the phenyl group in the SOXO unit, highlighting the beneficial role of the different substituted positions of the acceptor unit in facilitating the adjustment of ΔEST and f.38 In 2016, Su et al. investigated the structure–performance relationship of thioxanthone-based TADF emitters with substituted carbazole donors and inferior OLED performance was achieved for the symmetric D–A–D molecules with 2,7-substitutions.27 Interestingly, in the same year, they reported blue and yellow TXO-based TADF emitters with the D–A–D structure with an external quantum efficiency (EQE) of over 20%, in which they introduced two triphenylamine (TPA) donor units at the 3 and 6 position of the TX (9-H-thioxanthen-9-one) unit. Ultrahigh EQE values of 23.7% and 24.3% are achieved for the symmetrical blue emitter 3,6-2TPA-TX and the yellow emitter 3,6-2TPA-TXO, respectively.39 Although the SOXO-based TADF emitters have achieved excellent performance in OLED devices, the electroluminescence mechanism of this kind of material is still unclear. For the singlet-harvesting mechanism, both the singlet and triplet excited states are involved in the emission process, and the overall emission behavior relates to the individual properties of these states, ΔEST, f, the spin orbit coupling (SOC) and the related intersystem crossing (ISC) and RISC processes. All of these properties should be considered sufficiently when engineering the chemical structures of the TADF emitters.9,40,41 Therefore, a detailed understanding of the structural–property relationship of the organic TADF emitters is crucial for effective materials engineering, which facilitates the development of new efficient TADF emitters and promotes their applications in OLEDs.
In this work, the electroluminescence mechanisms of two experimentally reported D–A and D–A–D type SOXO based TADF emitters (the chemical structures are plotted in Fig. 1) are systematically investigated using density functional theory (DFT) and time-dependent density functional theory (TD-DFT). Two new SOXO derivatives are designed via simply introducing the phenyl ring between donor and acceptor units with D–π–A and D–π–A–π–D structures to further establish the structure–property relationship. The influence of the phenyl ring on the excited state properties of the SOXO-based molecules is investigated. The photophysical properties, the radiative and nonradiative decay rates, the ISC and RISC rates, as well as SOC constants of the investigated molecules, are explored in detail. It was found that the TADF emitters are modulated from green to blue emission, and the radiative decay rates and the rates of the RISC process are also enhanced through simply incorporating phenyl rings into the experimental molecules.
 |
| | Fig. 1 Chemical structures of the studied thioxanthone-based molecules with D–A (CzSOXO), D–π–A (Cz-Ph-SOXO), D–A–D (DCzSOXO) and D–π–A–π–D (DCz-DPh-SOXO) structures, θ and l denote the main dihedral angles and bond lengths. | |
2. Theoretical methods
The decay rate constants from the S1 to S0 states and the interconversion rate constants between the S1 and T1 states are key parameters in determining the electroluminescence mechanism of the SOXO-based TADF molecules. The radiative decay rate kr is calculated by the Einstein spontaneous emission equation as follows:42–44| |  | (1) |
where f is the oscillator strength and ΔEfi is the vertical emission energy with the unit of wavenumber (cm−1).
The nonradiative decay rate knr can be obtained based on Fermi's golden rule and the first-order perturbation theory45,46
| |  | (2) |
Here, the delta function δ is to ensure the conservation of energy; Piv is the Boltzmann distribution function for the initial vibronic manifold; Ĥfu,iv is the interaction operator between two different Born–Oppenheimer states, consisting of two contributions as below46
| |  | (3) |
here,
ĤBO is the nonadiabatic coupling and
ĤSO is the spin–orbit coupling,
r and
Q are the electronic and normal mode coordinates.
Φi is the electron wave function, and
Θiv is the nuclear vibrational wave function.
The internal conversion rate constant between two electronic states with the same spin multiplicity can be evaluated by using the first-order perturbation theory as45
| |  | (4) |
Here, the nonadiabatic electronic coupling matrix element
,
is the normal momentum operator of the kth normal mode in the final electronic states. Φf and Φi are the wavefunctions of the final state and the initial states, respectively. ρicfi is the thermal vibration correlation function (TVCF),42–44 which is
Here, Zi represents the partition function.
The intersystem crossing rate constant between the singlet and triplet states can be evaluated based on the following equation:42,43
| |  | (5) |
The detailed derivation of these equations can be found in Peng and Shuai's work.
45–47
3. Computational details
The geometric optimizations and frequency calculations of the S0 states are performed via density functional theory (DFT), while the excited singlet and triplet states are optimized through the time-dependent DFT (TD-DFT) approach. Since the excited state properties of TADF molecules with charge transfer character are functional dependent, several exchange–correlation functionals, including B3LYP,48 PBE0,49 PBE0-1/3,50 BMK51 and M062X52 with different HF exchange percentages in the XC functional of 20%, 25%, 33.33%, 42% and 54%, and two range-separated hybrid functionals, CAM-B3LYP53 and ωB97XD,54 are used to calculate the absorption and emission wavelengths of CzSOXO and DCzSOXO combining with the 6-31G(d)55–57 basis set (see Table 1). It was shown that the results based on the PBE0-1/3 functional agree well with the experimental values. The PBE0-1/3 functional and 6-31G(d) basis sets are therefore used to evaluate the ground and excited state properties of the studied molecules. All calculations are performed in a toluene solvent with the Polarizable Continuum Model (PCM)58 using the Gaussian 09 software package.59
Table 1 The calculated absorption (λab) and emission (λem) wavelengths of CzSOXO and DCzSOXO based on different functionals, together with the experimental values for comparison
|
|
CzSOXO
|
DCzSOXO
|
|
λ
ab (nm) |
λ
em (nm) |
λ
ab (nm) |
λ
em (nm) |
| B3LYP |
521 |
700 |
543 |
730 |
| PBE0 |
478 |
617 |
498 |
642 |
| PBE0-1/3 |
423 |
524 |
440 |
542 |
| BMK |
393 |
474 |
408 |
476 |
| M062X |
361 |
421 |
372 |
425 |
| CAM-B3LYP |
352 |
411 |
364 |
415 |
| ωB97XD |
341 |
389 |
349 |
395 |
| Expt.26 |
404 |
536 |
406 |
546 |
Based on the electronic properties, the calculations of the radiative and nonradiative decay rates, as well as the intersystem crossing rate and reverse intersystem crossing rate constants are finished through the Molecular Material Property Prediction Package (MOMAP).60–65 The SOC matrix elements are computed from the quadratic response function by the Dalton program package.66,67
In addition, to characterize the geometric changes between the S0 and S1 states and between the S1 and T1 states, the root of the mean of squared displacement (RMSD) is calculated using the following equation:
| |  | (6) |
where
i is the atomic ordinal number. To analyze the excitation properties of the investigated molecules, the electron–hole (e–h) distributions and the overlaps of e–h of the S
1 states are calculated through the Multiwfn
68,69 software.
4. Results and discussions
4.1 Molecular geometric structures
The molecular geometric structures play a crucial role in determining both the electronic structures and photophysical properties. A large distorted dihedral angle between D and A units generally allows efficient spatial separation of the HOMO and the LUMO to achieve a small ΔEST value. The geometric structures of the S0, S1 and excited triplet states for all the investigated molecules are optimized based on the PBE0-1/3/6-31G(d) level in a toluene solvent. The main dihedral angles and bond lengths (as shown in Fig. 1) of CzSOXO, DCzSOXO, Cz-Ph-SOXO and DCz-DPh-SOXO based on the optimized S0, S1 and T1 geometries in toluene are listed in Table 2. It can be seen that the molecular geometries between the S0 and S1 states are significantly altered when changing the molecular types. In the S0 states, the dihedral angle (θ1) between D and A moieties of CzSOXO is 50°, which is almost the same as that of DCzSOXO. However, when inserting the phenyl rings between the D and A unit in CzSOXO and DCzSOXO, the dihedral angles between the SOXO group and the phenyl ring are decreased to 37° for Cz-Ph-SOXO and DCz-DPh-SOXO, respectively. The corresponding bond lengths (l1) are also 0.08 Å larger than those of CzSOXO and DCzSOXO. When comparing the dihedral angles in Table 2 between the S0 and S1 states, it can be found that the dihedral angles between D and A units are increased by around 14° for CzSOXO and DCzSOXO. While the largest dihedral angle deviations between the S0 and S1 states for Cz-Ph-SOXO and DCz-DPh-SOXO have occurred between the carbazole unit and the inserted phenyl ring, and the value at S1 states is decreased by 6° with respect to those of S0 states. The decreased dihedral angles at S0 and S1 states make the geometries of Cz-Ph-SOXO and DCz-DPh-SOXO become more planar compared with those of CzSOXO and DCzSOXO. It can also be found that the T1 geometries of the studied molecules change significantly compared with the S1 geometries, and the selected dihedral angles of Cz-Ph-SOXO and DCz-DPh-SOXO at the T1 state show greater changes than those of CzSOXO and DCzSOXO with respect to those at the S1 state.
Table 2 The main dihedral angles (in degree) and bond lengths (in angstrom) of the designed SOXO-based molecules based on the optimized S0 and S1 geometries in toluene
| Molecules |
Bond parameters |
S0 geometry |
S1 geometry |
T1 geometry |
|
CzSOXO
|
θ
1
|
50 |
64 |
42 |
|
l
1
|
1.40 |
1.43 |
1.38 |
|
|
|
DCzSOXO
|
θ
11/θ12 |
49/49 |
64/53 |
49/43 |
|
l
11/l12 |
1.40/1.40 |
1.43/1.41 |
1.40/1.38 |
|
|
|
Cz-Ph-SOXO
|
θ
3/θ2 |
37/54 |
33/48 |
46/2 |
|
l
3/l2 |
1.48/1.41 |
1.47/1.40 |
1.41/1.39 |
|
|
|
DCz-DPh-SOXO
|
θ
31/θ32/θ21/θ22 |
36/37/53/126 |
35/33/56/132 |
4/36/46/126 |
|
l
31/l32/l21/l22 |
1.48/1.48/1.41/1.41 |
1.48/1.47/1.41/1.40 |
1.41/1.48/1.39/1.41 |
To quantitatively characterize the geometry changes between S0 and S1 states and between S1 and T1 states, the RMSD analysis of the investigated molecules is performed and the results are presented in Fig. 2. It can be found that the RMSD values between S0 and S1 states for CzSOXO and DCzSOXO are 0.344 and 0.394 Å, respectively. When bringing the phenyl ring between D and A units, the RMSD values decrease to 0.118 and 0.122 Å, respectively. The geometric changes between S1 and T1 states show much larger RMSD values than those between S0 and S1 states. The large geometry changes will lead to large reorganization energy between the corresponding two states and consequently lead to larger nonradiative decay rates. The geometry changes of the ground state and the excited states via changing the molecular types will definitely affect the distributions of the frontier molecular orbitals and the photophysical properties as discussed below.
 |
| | Fig. 2 Comparison of the geometries between S0 (pink), S1 (blue) and T1 (red) states in toluene for investigated molecules (RMSD values are presented). | |
4.2 Frontier molecular orbitals
The distributions of the frontier molecular orbitals play an important role in determining ΔEST, which is closely related to the steric hindrance of the molecular geometry. The more separated distributions between HOMOs and LUMOs, the smaller ΔEST will be. On the other hand, the spatial overlap between HOMO and LUMO is essential to achieve high radiative decay rates (kr). The distributions and energy levels of HOMOs and LUMOs for the investigated molecules in the S0 state are plotted in Fig. 3. It can be clearly seen that the HOMOs of CzSOXO and DCzSOXO are predominantly located on the donor units, and partially on the acceptor groups, and the LUMOs are mainly distributed on the acceptor moieties, and a small portion is distributed on the donor unit, ensuring a partial overlap between the distributions of HOMOs and LUMOs. When inserting the phenyl ring between D and A groups, smaller overlaps between HOMOs and LUMOs for Cz-Ph-SOXO and DCz-DPh-SOXO can also be found, which facilitates obtaining small ΔEST. It can also be found that the energy gaps between HOMOs and LUMOs for Cz-Ph-SOXO and DCz-DPh-SOXO are decreased by 0.14 and 0.10 eV compared with CzSOXO and DCzSOXO, respectively.
 |
| | Fig. 3 The distributions and energy levels of the frontier molecular orbitals for the studied molecules at the S0 state (isovalue is 0.02). | |
To analyze the excitation properties and quantitatively predict the degree of overlap between HOMOs and LUMOs, the electron–hole (e–h) distribution and overlap of e–h for the S1 state of the investigated molecules are provided in Fig. 4. It can be seen that there are obvious overlaps of e–h for CzSOXO and DCzSOXO, mainly occurring on the acceptor unit. For the designed molecules Cz-Ph-SOXO and DCz-DPh-SOXO, the e–h overlap became smaller than those of CzSOXO and DCzSOXO, indicating that they may possess smaller ΔEST. It can also be noticed that the e–h distributions are separated significantly for all the investigated molecules, indicating that the charge transfer process can occur in all the molecules. To further analyze the inside mechanisms, the spatial separation of holes and electrons Sr, the charge transfer (CT) length D, the degree of separation of holes and electrons t, the reflection of the overall average distribution breadth of electrons and holes H, and the Δr index for the S1 and S2 states used to measure the charge transfer length during the electron excitation, are calculated via Multiwfn software68 and the results are listed in Table 3. The calculated parameters in Table 3 also verify the above findings. When the t index is larger than zero, this implies that the distributions of holes and electrons are effectively separated because of the CT process. The Δr index for the S1 and S2 states is used to measure the charge transfer length during the electron excitation, and the larger the Δr value is, the more likely the excitation is a CT mode. It can be found that the Sr index of CzSOXO (0.32) and DCzSOXO (0.38) is larger than Cz-Ph-SOXO (0.24) and DCz-DPh-SOXO (0.29), indicating the larger overlap between holes and electrons, which is in accord with the e–h distributions as discussed above. It can also be found that t indices are all greater than zero, suggesting that sufficient hole and electron separations occurred in these molecules due to the CT process. However, the t index of DCzSOXO is as small as 0.08, demonstrating that there is some overlap between the hole and the electron. For the S1 and S2 states, Δr of all the molecules is larger than the threshold distinguishing the local excited (LE) state and the CT excited states of 2 Å, except that of DCzSOXO, indicating that they are LE states. It is also illustrated in Table 3 that the e–h overlaps can be decreased when incorporating the phenyl ring in Cz-Ph-SOXO and DCz-DPh-SOXO with respect to those of CzSOXO and DCzSOXO.
 |
| | Fig. 4 The electron–hole (e–h) distributions and overlaps of e–h for the S1 states of CzSOXO, DCzSOXO, Cz-Ph-SOXO and DCz-DPh-SOXO (isosurface value is 0.002). | |
Table 3 The Sr, D, t, H, and Δr index of the investigated molecules
| Molecules |
S
r/a.u. |
D/Å |
t/Å |
H/Å |
ΔrS1/Å |
ΔrS2/Å |
|
CzSOXO
|
0.32 |
5.08 |
4.62 |
2.81 |
5.66 |
6.49 |
|
DCzSOXO
|
0.38 |
1.44 |
0.08 |
4.40 |
1.58 |
1.62 |
|
Cz-Ph-SOXO
|
0.24 |
8.75 |
6.26 |
3.15 |
9.50 |
4.45 |
|
DCz-DPh-SOXO
|
0.29 |
3.52 |
1.78 |
6.02 |
4.88 |
4.94 |
4.3 ΔEST and transition properties
The singlet–triplet energy gap (ΔEST) for TADF molecules is one of the key parameters for the RISC process, which is correlated to the distributions of the HOMO and the LUMO. The excitation energies of S1 and T1 states based on the S0, S1 and T1 geometries and the vertical and adiabatic energy splitting between the S1 and T1 states for the investigated molecules are provided in Table 4. It can be noticed that the adiabatic energy gaps between the S1 and T1 states for CzSOXO and DCzSOXO are 0.35 and 0.31 eV, respectively. When inserting the phenyl ring between D and A units in CzSOXO and DCzSOXO, these values are increased to 0.63 and 0.59 eV, respectively. The results are in agreement with the analysis of the geometric changes. However, the energy gap of Cz-Ph-SOXO between the T3 and S1 states is much smaller than that of CzSOXO between T1 (T2) and S1 states (see Fig. 5), demonstrating that the ISC and the RISC processes may occur from the S1 to T3 states for Cz-Ph-SOXO. Similarly, for DCzSOXO and DCz-DPh-SOXO, the nearest triplet states to the S1 state are T2 (T3) and T3 (T4) states, respectively. So, the ISC process may occur from S1 to T2 (T3) and T3 (T4) states, respectively.
Table 4 The excitation energies of the S1 and T1 states based on S0 and S1 geometries and the energy difference between the S1 and T1 states for the investigated molecules (in unit of eV)
| Molecules |
S0 geometry |
S1 geometry |
T1 geometry |
Adiabatic |
| S1 |
T1 |
ΔEST |
S1 |
T1 |
ΔEST |
S1 |
T1 |
ΔEST |
ΔEST |
|
CzSOXO
|
2.93 |
2.63 |
0.30 |
2.37 |
2.24 |
0.13 |
2.49 |
1.85 |
0.64 |
0.35 |
|
DCzSOXO
|
2.82 |
2.54 |
0.28 |
2.29 |
2.16 |
0.13 |
2.41 |
1.85 |
0.56 |
0.31 |
|
Cz-Ph-SOXO
|
3.11 |
2.71 |
0.40 |
2.65 |
2.35 |
0.30 |
2.75 |
1.77 |
0.98 |
0.63 |
|
DCz-DPh-SOXO
|
3.06 |
2.68 |
0.38 |
2.61 |
2.33 |
0.28 |
2.71 |
1.77 |
0.94 |
0.59 |
 |
| | Fig. 5 The energy landscape of the vertical excitation of CzSOXO (a), Cz-Ph-SOXO (b), DCzSOXO (c) and DCz-DPh-SOXO (d). | |
The transition properties of excited states are also vital in determining the excited state properties. The natural transition orbital (NTO) analyses of the S1 and T1 states in toluene are performed for the investigated molecules. As shown in Fig. S1–S4 (ESI†), the highest occupied natural transition orbital (hole) and the lowest unoccupied natural transition orbital (particle) predominate the transition for S1 and T1 states. The S1 states of the studied molecules generally exhibit charge transfer (CT) characters. For the T1 and higher triplet states (T2, T3 or T4 states), it can be found that all the molecules not only show significant local excited (LE) properties, but also partial CT characters. The CT characters can help to achieve larger kRISC, therefore realizing a more efficient RISC process.22
4.4 Photophysical property
The photophysical property is vital in exploring the TADF process of the investigated molecules. The absorption and emission wavelength as well as the oscillator strength of the S1 state in a toluene solvent are provided in Table 5, together with the available experimental data. It can be found that the deviation between the theoretical and experimental absorption wavelengths of CzSOXO and DCzSOXO is less than 34 nm (with a relative difference of 7.7%). The calculated emission wavelengths are also in good agreement with the experimental results, and the deviations are within 12 nm. The emission wavelengths of CzSOXO and DCzSOXO are 524 and 542 nm, respectively, which are green emitters. While introducing the phenyl rings between the D and the A to CzSOXO and DCzSOXO make the emission wavelengths of Cz-Ph-SOXO and DCz-DPh-SOXO blue-shifted to 468 and 475 nm, respectively, which are blue emitters. When introducing the phenyl ring to CzSOXO and DCzSOXO between the D and A groups, the oscillator strengths of Cz-Ph-SOXO and DCz-DPh-SOXO are increased by 1.5 and 2 times, respectively, with respect to those of CzSOXO and DCzSOXO for the absorption process. While for the emission process, the oscillator strengths of Cz-Ph-SOXO and DCz-DPh-SOXO are also enhanced by 6 times than those of CzSOXO and DCzSOXO. From the perspective of oscillator strengths, it can be determined that the designed molecules (Cz-Ph-SOXO and DCz-DPh-SOXO) possess much better TADF properties compared with the experimental molecules (CzSOXO and DCzSOXO). Therefore, incorporating the phenyl ring between D and A moieties could be an efficient way for designing new efficient SOXO-based TADF molecules.
Table 5 The calculated absorption (λab) and emission (λem) wavelengths as well as the oscillator strength based on the PBE0-1/3/6-31G(d) level in a toluene solvent for the investigated molecules (the experimental values are also listed in the parentheses for comparison)
| Molecules |
λ
ab (nm) |
f
VA
|
λ
em (nm) |
f
VE
|
|
CzSOXO
|
423 (40426) |
0.0405 |
524 (53626) |
0.0145 |
|
DCzSOXO
|
440 (40626) |
0.0431 |
542 (54626) |
0.0157 |
|
Cz-Ph-SOXO
|
399 |
0.0613 |
468 |
0.0622 |
|
DCz-DPh-SOXO
|
405 |
0.0838 |
475 |
0.0665 |
4.5 Excited state dynamics
For TADF emitters, the rate constants of radiative (kr) and nonradiative (knr) processes from the S1 to S0 state, as well as the intersystem crossing (kISC) and reverse intersystem crossing (kRISC) rates play an important role in the dynamic process of the excited states. It is widely known that the ISC and RISC processes are related to not only the adiabatic energy gap between the S1 and T1 states, but also the SOC between the singlet and triplet states. The SOC constants between the S1 and T1 (T2, T3 and T4) states are calculated based on the S1 geometries through the Dalton66,67 package, as shown in Table 6. The kr and knr from the S1 to S0 state as well as kISC and kRISC between the S1 and T1 (T2, T3 and T4) states in toluene are provided in Table 7. The radiative decay rates of CzSOXO and DCzSOXO are 3.52 × 106 and 3.56 × 106 s−1, respectively. While this value for Cz-Ph-SOXO and DCz-DPh-SOXO is increased to 1.89 × 107 and 1.96 × 107 s−1, respectively, which agrees with the values of oscillator strengths. However, the nonradiative decay rates (knr) are all larger than the radiative rates (kr) for CzSOXO and DCzSOXO, which is not beneficial for the light-emitting process. This is because the nonradiative process is sensitive to the environment, and the calculated results are based on the single molecule model and without considering the intermolecular interaction. So when made into OLEDs, the nonradiative rates of the designed molecules can be reduced significantly than the calculated values as illustrated in previous studies.70–72 While knr of Cz-Ph-SOXO and DCz-DPh-SOXO is significantly decreased compared with kr, which is beneficial for the light-emitting process. It is also found in Table 6 that the calculated SOC values between the S1 state and the triplet state closest to the S1 state are larger than those values between the S1 and other triplet states.
Table 6 The spin–orbit coupling (SOC) constants between the S1 state and the selected triplet states in toluene for the investigated molecules (all the SOC values are in unit of cm−1)
| Molecules |
〈S1|HSO|T1〉 |
〈S1|HSO|T2〉 |
〈S1|HSO|T3〉 |
〈S1|HSO|T4〉 |
|
CzSOXO
|
0.228 |
0.358 |
— |
— |
|
DCzSOXO
|
0.215 |
0.183 |
0.366 |
— |
|
Cz-Ph-SOXO
|
0.158 |
0.432 |
0.374 |
— |
|
DCz-DPh-SOXO
|
0.141 |
0.158 |
0.043 |
0.365 |
Table 7 The rate constants of radiative (kr) and nonradiative (knr) from the S1 to S0 state as well as the ISC (kISC) and RISC (kRISC) between the S1 and T1 (T2, T3 and T4) states in toluene (all the rates are in unit of s−1)
| Molecules |
CzSOXO
|
DCzSOXO
|
Cz-Ph-SOXO
|
DCz-DPh-SOXO
|
|
k
r (S1 → S0) |
3.52 × 106 |
3.56 × 106 |
1.89 × 107 |
1.96 × 107 |
|
k
nr (S1 → S0) |
5.14 × 1010 |
8.40 × 1010 |
7.50 × 105 |
3.83 × 106 |
|
k
ISC (S1 → T1) |
8.27 × 106 |
1.02 × 106 |
7.00 × 104 |
5.65 × 104 |
|
k
RISC (T1 → S1) |
9.89 × 100 |
1.80 × 100 |
6.58 × 10−5 |
1.82 × 10−2 |
|
k
ISC (S1 → T2) |
2.06 × 107 |
2.53 × 106 |
5.09 × 105 |
1.59 × 104 |
|
k
RISC (T2 → S1) |
2.39 × 101 |
6.82 × 102 |
3.01 × 10−4 |
6.32 × 10−5 |
|
k
ISC (S1 → T3) |
— |
1.01 × 107 |
5.56 × 105 |
1.43 × 103 |
|
k
RISC (T3 → S1) |
— |
2.92 × 103 |
1.41 × 106 |
2.66 × 104 |
|
k
ISC (S1 → T4) |
— |
— |
— |
1.02 × 105 |
|
k
RISC (T4 → S1) |
— |
— |
— |
2.50 × 106 |
For CzSOXO, the T1 and T2 states are degenerate, and both the T1 and T2 states participate in the ISC and RISC processes. The kISC (2.06 × 107 s−1) and kRISC (2.39 × 101 s−1) between the S1 and T2 states are more effective than those between the S1 and T1 states due to the larger SOC values between the S1 and T2 states. It can be seen in Table 7 that the T1, T2 and T3 states of DCzSOXO are all involved in the ISC and RISC processes. The most efficient route of the ISC and RISC processes for DCzSOXO is between the S1 and T3 states, and kISC and kRISC are 1.01 × 107 s−1 and 2.92 × 103 s−1, respectively. It can also be found in Fig. 5 that the T2 and T3 states of DCzSOXO are degenerate. However, the energy gap between T1 and T2 (T3) states is as small as 0.19 eV, so the internal conversion process will occur quickly from the T2 (T3) to the T1 state, and the RISC process might also happen from the T1 to S1 state. For Cz-Ph-SOXO and DCz-DPh-SOXO, kISC values of the most effective route are decreased to 5.56 × 105 and 1.02 × 105 s−1, respectively. However, kRISC values of Cz-Ph-SOXO (1.41 × 106 s−1) and DCz-DPh-SOXO (2.50 × 106 s−1) of the most effective routes are significantly increased compared with those of CzSOXO (2.39 × 101 s−1) and DCzSOXO (2.92 × 103 s−1), demonstrating that Cz-Ph-SOXO and DCz-DPh-SOXO are more facilitated to the occurrence of the RISC process. It can also be found that the RISC from the T1 (T2) to S1 state is very small for Cz-Ph-SOXO, which is due to the larger energy gap (0.46 eV) between the T3 and T1 (T2) states. Therefore, the internal conversion process from the T3 to T1 (T2) state will be slow and the RISC process for Cz-Ph-SOXO mainly occurs from the T3 to S1 state. Similarly, for DCz-DPh-SOXO, the energy gap between T3 (T4) and T1 states is also as large as 0.46 eV, so the main ISC and RISC processes have happened between the S1 and T3 (T4) states.
In addition, the phosphorescence rates (kp) from the T1 to S0 states calculated through MOMAP60–65 software for all the investigated molecules are presented in Table S1 (ESI†). It can be found that the values of kRISC are apparently much larger than the corresponding kp, indicating that the reverse intersystem crossing process can successfully compete with the phosphorescence process. Based on the analysis of the dynamics of the excited states, it can be concluded that the designed molecules Cz-Ph-SOXO and DCz-DPh-SOXO with high kr and efficient kRISC can be used as excellent TADF emitters. Therefore, incorporating phenyl units into D–A and D–A–D type molecules is a good strategy for developing new efficient SOXO-based TADF emitters, and the emission color can also be modulated from green to blue simply by introducing the phenyl ring.
5. Conclusions
In summary, the photophysical processes and excited state dynamics of four SOXO-based molecules are investigated based on the first principles calculation. It is found that incorporating phenyl units between the donor and acceptor groups and changing the molecular types can significantly modulate the excited state properties of these molecules. The TADF mechanisms of the studied molecules are revealed by calculating kr, knr, kISC and kRISC, ΔEST and SOC based on the thermal vibration correlation function method and TD-DFT approach. Results show that the emission wavelengths of Cz-Ph-SOXO and DCz-DPh-SOXO are also modulated from green to blue by simply inserting the phenyl ring between the donor and acceptor groups. For the designed molecules Cz-Ph-SOXO and DCz-DPh-SOXO, the kr values increased significantly compared with those of CzSOXO and DCzSOXO. The kISC values of Cz-Ph-SOXO and DCz-DPh-SOXO are relatively smaller than those of CzSOXO and DCzSOXO. However, the kRISC values of Cz-Ph-SOXO and DCz-DPh-SOXO are significantly increased compared with those of CzSOXO and DCzSOXO, demonstrating that Cz-Ph-SOXO and DCz-DPh-SOXO are more facilitated to achieve the TADF process. In particular, the kRISC values of Cz-Ph-SOXO and DCz-DPh-SOXO reached up to 1.41 × 106 s−1 and 2.50 × 106 s−1, respectively, which is beneficial to the occurrence of delayed fluorescence. Our results would be helpful for developing new SOXO-based TADF materials experimentally.
Conflicts of interest
The authors have no conflicts to disclose.
Acknowledgements
This work was supported by the Natural Science Foundation of Shandong, China, Grant No. ZR2020QB074, the National Natural Science Foundation of China, Grant No. 21503056, the Fundamental Research Funds for the Central Universities, Grant No. HIT. NSRIF. 2016090. The authors gratefully acknowledge HZWTECH for providing computational facilities.
References
- Q. Zhang, J. Li, K. Shizu, S. Huang, S. Hirata, H. Miyazaki and C. Adachi, J. Am. Chem. Soc., 2012, 134, 14706–14709 CrossRef CAS PubMed.
- H. Uoyama, K. Goushi, K. Shizu, H. Nomura and C. Adachi, Nature, 2012, 492, 234–238 CrossRef CAS PubMed.
- J. Li, T. Nakagawa, J. MacDonald, Q. Zhang, H. Nomura, H. Miyazaki and C. Adachi, Adv. Mater., 2013, 25, 3319–3323 CrossRef CAS PubMed.
- Y. Im, M. Kim, Y. J. Cho, J. A. Seo, K. S. Yook and J. Y. Lee, Chem. Mater., 2017, 29, 1946–1963 CrossRef CAS.
- T. Hofbeck, U. Monkowius and H. Yersin, J. Am. Chem. Soc., 2015, 137, 399–404 CrossRef CAS PubMed.
- S. Hirata, Y. Sakai, K. Masui, H. Tanaka, S. Y. lee, H. Nomura, N. Nakamura, M. Yasumatsu, H. Nakanotani, Q. S. Zhang, K. Shizu, H. Miyazaki and C. Adachi, Nat. Mater., 2015, 14, 330–336 CrossRef CAS PubMed.
- H. T. Sun, Z. B. Hu, C. Zhong, X. K. Chen, Z. R. Sun and J.-L. Brédas, J. Phys. Chem. Lett., 2017, 8, 2393–2398 CrossRef CAS PubMed.
- K. Kawasumi, T. Wu, T. Y. Zhu, H. S. Chae, T. V. Voorhis, M. A. Baldo and T. M. Swager, J. Am. Chem. Soc., 2015, 137, 11908–11911 CrossRef CAS PubMed.
- Y. Tao, K. Yuan, T. Chen, P. Xu, H. Li, R. Chen, C. Zheng, L. Zhang and W. Huang, Adv. Mater., 2014, 26, 7931–7958 CrossRef CAS PubMed.
- G. Mehes, H. Nomura, Q. Zhang, T. Nakagawa and C. Adachi, Angew. Chem., Int. Ed., 2012, 51, 11311–11315 CrossRef CAS PubMed.
- S. Y. Lee, T. Yasuda, H. Nomura and C. Adachi, Appl. Phys. Lett., 2012, 101, 093306 CrossRef.
- T. Nakagawa, S. Y. Ku, K. T. Wong and C. Adachi, Chem. Commun., 2012, 48, 9580–9582 RSC.
- H. Nakanotani, K. Masui, J. Nishide, T. Shibata and C. Adachi, Sci. Rep., 2013, 3, 2127 CrossRef CAS PubMed.
- H. Tsujimoto, D.-G. Ha, G. Markopoulos, H. S. Chae, M. A. Baldo and T. M. Swager, J. Am. Chem. Soc., 2017, 139, 4894–4900 CrossRef CAS PubMed.
- A. E. Nikolaenko, M. Cass, F. Bourcet, D. Mohamad and M. Roberts, Adv. Mater., 2015, 27, 7236–7240 CrossRef CAS PubMed.
- P. Rajamalli, N. Senthilkumar, P. Gandeepan, P. Y. Huang, M. J. Huang, C. Z. Ren-Wu, C. Y. Yang, M. J. Chiu, L. K. Chu, H. W. Lin and C. H. Cheng, J. Am. Chem. Soc., 2016, 138, 628–634 CrossRef CAS PubMed.
- T. Hatakeyama, K. Shiren, K. Nakajima, S. Nomura, S. Nakatsuka, K. Kinoshita, J. Ni, Y. Ono and T. Ikuta, Adv. Mater., 2016, 28, 2777–2781 CrossRef CAS PubMed.
- T. A. Lin, T. Chatterjee, W. L. Tsai, W. K. Lee, M. J. Wu, M. Jiao, K. C. Pan, C. L. Yi, C. L. Chung, K. T. Wong and C. C. Wu, Adv. Mater., 2016, 28, 6976–6983 CrossRef CAS PubMed.
- M. Numata, T. Yasuda and C. Adachi, Chem. Commun., 2015, 51, 9443–9446 RSC.
- X. K. Chen, Y. Tsuchiya, Y. Ishikawa, C. Zhong, C. Adachi and J.-L. Brédas, Adv. Mater., 2017, 29, 1702767 CrossRef PubMed.
- Q. Zhu, X. Guo and J. Zhang, J. Comput. Chem., 2019, 40, 1578–1585 CrossRef CAS PubMed.
- L. J. Wang, T. Li, P. C. Feng and Y. Song, Phys. Chem. Chem. Phys., 2017, 19, 21639–21647 RSC.
- Y. C. Li, Z. H. Wang, X. L. Li, G. Z. Xie, D. C. Chen, Y. F. Wang, C. C. Lo, A. Lien, J. B. Peng, Y. Cao and S. J. Su, Chem. Mater., 2015, 27, 1100–1109 CrossRef CAS.
- X. F. Lu, S. H. Fan, J. H. Wu, X. W. Jia, Z. S. Wang and G. Zhou, J. Org. Chem., 2014, 79, 6480–6489 CrossRef CAS PubMed.
- P. Data, P. Pander, M. Okazaki, Y. Takeda, S. Minakata and A. P. Monkman, Angew. Chem., Int. Ed., 2016, 55, 5739–5744 CrossRef CAS PubMed.
- M. K. Etherington, F. Franchell, J. Gibson, T. Northey, J. Santos, J. S. Ward, H. F. Higginbotham, P. Data, A. Kurowska, P. L. D. Santos, D. R. Graves, A. S. Batsaov, F. B. Dias, M. R. Bryce, T. J. Penfold and A. P. Monkman, Nat. Commun., 2017, 8, 14987 CrossRef CAS PubMed.
- Z. H. Wang, Y. C. Li, X. Y. Cai, D. C. Chen, G. Z. Xie, K. K. Liu, Y. C. Wu, C. C. Lo, A. Lien, Y. Cao and S. J. Su, ACS Appl. Mater. Interfaces, 2016, 8, 8627–8636 CrossRef CAS PubMed.
- J. B. Im, R. Lampande, G. H. Kim, J. Y. Lee and J. H. Kwon, J. Phys. Chem. C, 2017, 121, 1305–1314 CrossRef CAS.
- C. C. Fan, C. B. Duan, C. M. Han, B. Han and H. Xu, ACS Appl. Mater. Interfaces, 2016, 8, 27383–27393 CrossRef CAS PubMed.
- C. Wang, X. L. Li, Y. Y. Pan, S. T. Zhang, L. Yao, Q. Bai, W. J. Li, P. Lu, B. Yang, S. J. Su and Y. G. Ma, ACS Appl. Mater. Interfaces, 2016, 8, 3041–3049 CrossRef CAS PubMed.
- S. Y. Lee, C. Adachi and T. Yasuda, Adv. Mater., 2016, 28, 4626–4631 CrossRef CAS PubMed.
- S. Ishijima, M. Higashi and H. Yamaguchi, J. Phys. Chem., 1994, 98, 10432–10435 CrossRef CAS.
- C. Ley, F. Morlet-Savary, P. Jacques and J. P. Fouassier, Chem. Phys., 2000, 255, 335–346 CrossRef CAS.
- A. M. Turek, G. Krishnamoorthy, K. Phipps and J. Saltiel, J. Phys. Chem. A, 2002, 106, 6044–6052 CrossRef CAS.
- S. Reineke and M. A. Baldo, Sci. Rep., 2014, 4, 3797 CrossRef PubMed.
- H. Wang, L. S. Xie, Q. Peng, L. Q. Meng, Y. Wang, Y. P. Yi and P. F. Wang, Adv. Mater., 2014, 26, 5198–5204 CrossRef CAS PubMed.
- X. F. Wei, Y. Z. Chen, R. H. Duan, J. J. Liu, R. F. Wang, Y. W. Liu, Z. Y. Li, Y. P. Yi, Y. Y. Takamura, P. F. Wang and Y. Wang, J. Mater. Chem. C, 2017, 5, 12077–12084 RSC.
- X. Wei, Z. Li, T. Hu, R. Duan, J. Liu, R. Wang, Y. Liu, X. Hu, Y. Yi, P. Wang and Y. Wang, Adv. Opt. Mater., 2019, 7, 1801767 CrossRef.
- Y. C. Li, X. L. Li, D. J. Chen, X. Y. Cai, G. Z. Xie, Z. Z. He, Y. C. Wu, A. Lien, Y. Cao and S. J. Su, Adv. Funct. Mater., 2016, 26, 6904–6912 CrossRef CAS.
- M. Y. Wong and E. Zysman-Colman, Adv. Mater., 2017, 29, 1605444 CrossRef PubMed.
- Z. Tu, G. Han, T. Hu, R. Duan and Y. Yi, Chem. Mater., 2019, 31, 6665–6671 CrossRef CAS.
- Q. Lu, M. Qin, S. Liu, L. Lin, C.-K. Wang, J. Fan and Y. Song, Chem. Phys. Lett., 2021, 764, 138260 CrossRef CAS.
- Y. Liu, L. Wang, L. Xu and Y. Song, J. Mater. Chem. C, 2023, 11, 13403–13417 RSC.
- Q. Lu, G. Y. Jiang, F. Y. Li, L. L. Lin, C. K. Wang, J. Z. Fan and Y. Z. Song, Spectrochim. Acta, Part A, 2020, 229, 117964 CrossRef CAS PubMed.
- Q. Peng, D. Fan, R. H. Duan, Y. P. Yi, Y. L. Niu, D. Wang and Z. G. Shuai, J. Phys. Chem. C, 2017, 121, 13448–13456 CrossRef CAS.
- Q. Peng, Y. L. Niu, Q. Shi, X. Gao and Z. G. Shuai, J. Chem. Theory Comput., 2013, 9, 1132–1143 CrossRef CAS PubMed.
- Y. L. Niu, Q. Peng, C. M. Deng, X. Gao and Z. G. Shuai, J. Phys. Chem. A, 2010, 114, 7817–7831 CrossRef CAS PubMed.
- B. Miehlich, A. Savin, H. Stoll and H. Preuss, Chem. Phys. Lett., 1989, 157, 200–206 CrossRef CAS.
- C. Adamo and V. Barone, Toward Reliable Density Functional Methods without Adjustable Parameters: The PBE0 Model, J. Chem. Phys., 1999, 110, 6158–6170 CrossRef CAS.
- C. A. Guido, E. Bremond, C. Adamo and P. Cortona, J. Chem. Phys., 2013, 138, 021104 CrossRef PubMed.
- A. D. Boese and J. M. L. Martin, J. Chem. Phys., 2004, 121, 3405–3416 CrossRef CAS PubMed.
- Y. Zhao and D. G. Truhlar, Theor. Chem. Acc., 2008, 120, 215–241 Search PubMed.
- T. Yanai, D. P. Tew and N. C. Handy, Chem. Phys. Lett., 2004, 393, 51–57 CrossRef CAS.
- Y.-S. Lin, G.-D. Li, S.-P. Mao and J.-D. Chai, J. Chem. Theory Comput., 2013, 9, 263–272 CrossRef CAS PubMed.
- H. Ma, W. Shi, J. Ren, W. Li, Q. Peng and Z. Shuai, J. Phys. Chem. Lett., 2016, 7, 2893–2898 CrossRef CAS PubMed.
- J. Liu, J. Fan, K. Zhang, Y. Zhang, C.-K. Wang and L. Lin, Chin. Phys. B, 2020, 29, 088504 CrossRef CAS.
- Q. Peng, Y. Yi, Z. Shuai and J. Shao, J. Chem. Phys., 2007, 126, 114302 CrossRef PubMed.
- J. Tomasi, B. Mennucci and R. Cammi, J. Cheminform., 2005, 105, 2999–3093 CAS.
-
M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian 09, Revision D.01, Gaussian Inc., Wallingford, CT, 2010 Search PubMed.
- Y. L. Niu, W. Q. Li, Q. Peng, H. Geng, Y. P. Yi, L. J. Wang, G. J. Nan, D. Wang and Z. G. Shuai, Mol. Phys., 2018, 116, 1078–1090 CrossRef CAS.
- Q. Peng, Y. P. Yi, Z. G. Shuai and J. S. Shao, J. Am. Chem. Soc., 2007, 129, 9333–9339 CrossRef CAS PubMed.
- Y. L. Niu, Q. Peng and Z. G. Shuai, Sci. China Ser. B-Chem., 2008, 51, 1153–1158 CrossRef CAS.
- Z. G. Shuai, Chin. J. Chem., 2020, 38, 1223–1232 CrossRef CAS.
- Z. G. Shuai and Q. Peng, Phys. Rep., 2014, 537, 123–156 CrossRef CAS.
- Z. G. Shuai and Q. Peng, Nat. Sci. Rev., 2017, 4, 224–239 CrossRef CAS.
- DALTON. A molecular electronic structure program, Release Dalton 2019. Alpha, 2019, https://daltonprogram.org/.
- K. Aidas, C. Angeli, K. L. Bak, V. Bakken, R. Bast, L. Boman, O. Christiansen, R. Cimiraglia, S. Coriani, P. Dahle, E. K. Dalskov, U. Ekström, T. Enevoldsen, J. J. Eriksen, P. Ettenhuber, B. Fernández, L. Ferrighi, H. Fliegl, L. Frediani, K. Hald, A. Halkier, C. Hättig, H. Heiberg, T. Helgaker, A. C. Hennum, H. Hettema, E. Hjertenæs, S. Høst, I.-M. Høyvik, M. F. Iozzi, B. Jansik, H. J. A. Jensen, D. Jonsson, P. Jørgensen, J. Kauczor, S. Kirpekar, T. Kjærgaard, W. Klopper, S. Knecht, R. Kobayashi, H. Koch, J. Kongsted, A. Krapp, K. Kristensen, A. Ligabue, O. B. Lutnæs, J. I. Melo, K. V. Mikkelsen, R. H. Myhre, C. Neiss, C. B. Nielsen, P. Norman, J. Olsen, J. M. H. Olsen, A. Osted, M. J. Packer, F. Pawlowski, T. B. Pedersen, P. F. Provasi, S. Reine, Z. Rinkevicius, T. A. Ruden, K. Ruud, V. Rybkin, P. Salek, C. C. M. Samson, A. Sánchez de Merás, T. Saue, S. P. A. Sauer, B. Schimmelpfennig, K. Sneskov, A. H. Steindal, K. O. Sylvester-Hvid, P. R. Taylor, A. M. Teale, E. I. Tellgren, D. P. Tew, A. J. Thorvaldsen, L. Thøgersen, O. Vahtras, M. A. Watson, D. J. D. Wilson, M. Ziolkowski and H. Ågren, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2014, 4, 269–284 CAS.
- T. Lu and F. W. Chen, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed.
- Z. Y. Liu, T. Lu and Q. X. Chen, Carbon, 2020, 165, 461–467 CrossRef CAS.
- J. Leng, Z. Zhang, Y. Zhang, J. Sun and H. Ma, J. Lumin., 2018, 204, 312–318 CrossRef CAS.
- J. Fan, L. Cai, L. Lin and C.-K. Wang, Phys. Chem. Chem. Phys., 2017, 19, 29872–29879 RSC.
- B. Liu, H. Nie, X. Zhou, S. Hu, D. Luo, D. Gao, J. Zou, M. Xu, L. Wang, Z. Zhao, A. Qin, J. Peng, H. Ning, Y. Cao and B. Z. Tang, Adv. Funct. Mater., 2016, 26, 776–783 CrossRef CAS.
|
| This journal is © the Owner Societies 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.