Next Article in Journal
MUC1 Tissue Expression and Its Soluble Form CA15-3 Identify a Clear Cell Renal Cell Carcinoma with Distinct Metabolic Profile and Poor Clinical Outcome
Next Article in Special Issue
Effect of Synthesis Method on Reaction Mechanism for Hydrogen Evolution over CuxOy/TiO2 Photocatalysts: A Kinetic Analysis
Previous Article in Journal
Insight into the Spatial Arrangement of the Lysine Tyrosylquinone and Cu2+ in the Active Site of Lysyl Oxidase-like 2
Previous Article in Special Issue
The Lattice Model of Particles with Orientation-Dependent Interactions at Solid Surfaces: Wetting Scenarios
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Interplay between ESIPT and TADF for the 2,2′-Bipyridine-3,3′-diol: A Theoretical Reconsideration

Institute of Atomic and Molecular Physics, Jilin University, Changchun 130012, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 13969; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232213969
Submission received: 7 October 2022 / Revised: 1 November 2022 / Accepted: 9 November 2022 / Published: 12 November 2022
(This article belongs to the Collection Feature Papers in 'Physical Chemistry and Chemical Physics')

Abstract

:
Organic molecules with excited-state intramolecular proton transfer (ESIPT) and thermally activated delayed fluorescence (TADF) properties have great potential for realizing efficient organic light-emitting diodes (OLEDs). Furthermore, 2,2′-bipyridine-3,3′-diol (BP(OH)2) is a typical molecule with ESIPT and TADF properties. Previously, the double ESIPT state was proved to be a luminescent state, and the T2 state plays a dominant role in TADF for the molecule. Nevertheless, whether BP(OH)2 undergoes a double or single ESIPT process is controversial. Since different ESIPT channels will bring different TADF mechanisms, the previously proposed TADF mechanism based on the double ESIPT structure for BP(OH)2 needs to be reconsidered. Herein, reduced density gradient, potential energy surface, IR spectra and exited-state hydrogen-bond dynamics computations confirm that BP(OH)2 undergoes the barrierless single ESIPT process rather than the double ESIPT process with a barrier. Moreover, based on the single ESIPT structure, we calculated spin-orbit coupling matrix elements, nonradiative rates and electron-hole distributions. These results disclose that the T3 state plays a predominant role in TADF. Our investigation provides a better understanding on the TADF mechanism in hydrogen-bonded molecular systems and the interaction between ESIPT and TADF, which further provides a reference for developing efficient OLEDs.

1. Introduction

Organic molecules with thermally activated delayed fluorescence (TADF) properties have aroused extensive attention due to their significant applications in organic light-emitting diodes (OLEDs) [1,2,3,4,5]. These applications are attributed to their high utilization of both singlet and triplet excitons. For conventional fluorescent OLEDs, only 25% singlet excitons are available, which causes low internal quantum efficiency [6,7]. In contrast, despite the fact that phosphorescent OLEDs can take full advantage of excitons by doping iridium or platinum complexes, their presence brings high costs and environmental pollution [8,9]. In 2012, Adachi’s team realized 100% internal quantum efficiency using TADF OLEDs [10]. TADF is the phenomenon of converting triplet excitons to singlet states via a fast RISC process under thermal activation [11,12]. Indeed, the luminous efficiency of molecules with the TADF properties suffer from the self-absorption effect, which hampers the efficiency improvement and affects practical application [13]. It is worth noting that the organic molecules with excited-state intramolecular proton transfer (ESIPT) properties can avoid this problem due to the significant Stokes shift [14,15,16,17,18]. ESIPT can be regarded as an ultrafast photo-tautomerization process in which a proton is transferred from proton donor to proton acceptor along with a strong intramolecular hydrogen bond (IHB) [19,20,21,22,23]. Therefore, the research of the organic TADF molecules capable of ESIPT effectively facilitates internal quantum efficiency, characterized by high luminous efficiency, low cost and long lifetime [24,25].
Additionally, 2,2′-bipyridine-3,3′-diol (BP(OH)2) is a typical molecule that simultaneously exhibits the ESIPT properties and TADF features. Bulska et al. reported that BP(OH)2 under the S1 state would undergo a cooperative double ESIPT reaction [26]. Based on the double ESIPT form, Tokumura et al. put forward the TADF mechanism for BP(OH)2 in hexane [27]. Concretely, after photoexcitation, the rapid double ESIPT happens, and then the T2 structure can be obtained through the intersystem crossing (ISC) process in competition with emissions from the S1 state. Following the ISC process, the reverse intersystem crossing (RISC) process from T2 to S1 state occurs, which thereby results in the TADF phenomenon. It was found that only the S1 and T2 states participate in the ISC and RISC processes. Plasser et al. deemed that BP(OH)2 prefers to occur in the stepwise double ESIPT process rather than in the concerted double ESIPT process, based on the TDDFT and resolution-of-identity second-order approximate coupled-cluster methods [28]. Recently, Zhao et al. suggested that, in low-polar solvents, the single ESIPT is more likely to occur in BP(OH)2 [29]. Accordingly, whether the molecule undergoes single or double proton transfer has not been identified completely. However, different geometric structures of the S1 state significantly affect the spin-orbit coupling interaction between the singlet and triplet states, which further brings different TADF mechanisms [30,31]. Therefore, the previously proposed TADF mechanism based on the double ESIPT structure needs to be reconsidered.
In this study, we have demonstrated the ESIPT pathway and reconsidered the TADF mechanism of BP(OH)2 theoretically in low-polar hexane solvent on account of density functional theory (DFT) and time-dependent density functional theory (TDDFT) methods. Specifically, the optimized geometric structures and absorption and emission peaks were analyzed to investigate the excited-state behavior for BP(OH)2. In addition, the calculations of the frontier molecular orbitals (FMOs), infrared (IR) spectra, reduced density gradient (RDG), potential energy curves (PECs), potential energy surface (PES), and excited-state hydrogen-bond dynamics were applied to explore the ESIPT pathway. Moreover, the spin-orbit coupling (SOC) constants, energy level gaps and nonradiative decay rates were calculated to describe the TADF mechanism of BP(OH)2. Furthermore, the electron-hole distribution was applied to analyze the electron excitation nature. In a word, our theoretical investigations demonstrate that BP(OH)2 will undergo the single ESIPT process rather than the double ESIPT process, and provide reconsiderations of the mechanisms of the ISC and RISC processes involved in the TADF phenomenon of this molecule. These findings provide a theoretical foundation for understanding how the ESIPT mechanism affects TADF, which favors achieving highly efficient photoluminescence.

2. Results and Discussion

2.1. Structural Analysis and Absorption and Fluorescence Spectra

Figure 1 presents optimized geometric structures of S0 and S1 states for BP(OH)2 in hexane at the B3LYP/def2-TZVP level with the dispersion correction (gd3bj). Meanwhile, the primary structural parameters involved in the IHBs of BP(OH)2 in hexane are collected in Table 1. The optimized configurations show that the Enol* form for BP(OH)2 does not exist in the S1 state. In addition, there are two possible ESIPT pathways in the S1 state: the single ESIPT path and double ESIPT path. For the single proton transfer path, the distance between N1 and H1 decreases from 1.659 Å to 1.026 Å, and that between O1 and H1 is elongated from 1.005 Å to 1.788 Å. For the double proton transfer path, the N1-H1 length is shortened from 1.659 Å to 1.034 Å, and O1-H1 is stretched from 1.005 Å to 1.725 Å. Similar to N1-H1 and O1-H1, the N2-H2 distance is decreased, while the length of O2–H2 exhibits a tendency of elongation. Therefore, the ESIPT pathways for BP(OH)2 can be either the single ESIPT path or the double ESIPT path.
The absorption and fluorescence peaks for BP(OH)2 in hexane were calculated at the B3LYP/def2-TZVP (gd3bj) level. The relevant data are recorded in Table 2. The theoretical absorption maximum of the Enol form locates at 337 nm, being consistent with the experimental data (345 nm) [27]. This illustrates the dependability of our chosen calculation method for BP(OH)2 system. The fluorescence peaks emitted from the optimized Keto*1 and Keto*2 forms are found at 484 nm and 479 nm, respectively, which coincide with the experimental peak (499 nm). Thus, the fluorescence is believed to be derived from the Keto*1 and Keto*2 structures combined with the optimized structures. However, in Table 2, the fluorescence peak of the Keto*1 conformation is in better agreement with 499 nm than that of the Keto*2 conformation.

2.2. Frontier Molecular Orbitals

The tendency of the ESIPT behavior can be effectively affected by charge distribution changes [32]. For exploring the mechanism of the ESIPT process, we analyzed the charge distribution for the FMOs (HOMO and LUMO) of the S0 and S1 states in BP(OH)2, as drawn in Figure 2. We find that the transition from HOMO to LUMO is corresponding to a dominant π-π∗ type character, which plays a significant role in promoting the proton transfer. Herein, we focus only on the charge distribution around the IHB, affecting the excited-state hydrogen bond dynamics. As detailed in Figure 2, the electron density distribution of both hydroxyl oxygens, which are related to the ESIPT process, decreases, whereas that of the proton acceptors increases from HOMO to LUMO. These changes mean that the electronegativities of the nitrogen atoms are stronger than those of the oxygen atoms for BP(OH)2 after the transition from HOMO to LUMO. Therefore, the FMOs analysis illustrates that both IHBs in BP(OH)2 are reinforced upon excitation to S1, which in turn facilitates the occurrence of the proton transfer process.

2.3. Reduced Density Gradient Analysis

RDG can clearly characterize the interactions of molecules in real space and the intensity of the interactions [15,33]. Specifically, RDG can be expressed as the following Formula (1):
RDG ( r ) = 1 2 ( 3 π 2 ) 1 3 | ρ ( r ) | ρ ( r ) 4 3
In addition, the electron density ρ(r) is associated with the eigenvalues I2 in the Hessian matrix of electron density, as shown in Equation (2)
Ω ( r ) = sign ( I 2 ( r ) ) ρ ( r )
Herein, the Ω(r) < 0 exhibits hydrogen-bonding interactions. The Ω(r) 0 stands for the van der Waals interaction. The Ω(r) > 0 refers to the steric effect.
As for BP(OH)2, to unveil the location and strength of IHB interactions and further judge the ESIPT pathway, the RDG image and gradient isosurfaces of this molecule are provided in Figure 3. The spikes inside the circles denote the strength of the hydrogen bond. Additionally, there is a blue disk between the oxygen and hydrogen atoms, and the hydrogen and nitrogen atoms, corresponding to the spikes in the circles. As detailed in Figure 3, the spike position around −0.06 for BP(OH)2 reveals the hydrogen-bonding interactions of O2-H2…N2 and O1-H1…N1 at the S0 state. For the Keto*1 state, the two spike peaks located around −0.06 and −0.04 are attributed to O1-H1…N1 and O2…H2-N2, respectively. The IHBs corresponding to the Keto*2 state are O2…H2-N2 and O1…H1-N1 with the spike peak of −0.05. During the process from the S0 to Keto*1 state, the IHB changes; namely, the proton H1 is transferred to N1 atoms. Moreover, the peak position of O2-H2…N2 shifted from −0.06 of the S0 state to −0.04 of the Keto*1 state, implying that the intensity of IHB (O2-H2…N2) is weakened. The change of two IHBs means that only single ESIPT reaction can occur in BP(OH)2. During the process from the S0 to the Keto*2 state, the hydrogen bond positions change from O1-H1…N1 and O2 H2…N2 to O1…H1-N1 and O2…H2-N2. However, the absolute spike peaks (0.04) of the Keto*2 state are smaller than that of the O1…H1-N1 in the Keto*1 state (0.05). This indicates that the strength of the newly formed IHB after the single ESIPT process is significantly stronger than those of the two IHBs after the double ESIPT process. Therefore, these results confirm that the intramolecular single ESIPT reaction is more likely to occur than the double ESIPT reaction.

2.4. Potential Energy Surface and Potential Energy Curves

In order to determine whether BP(OH)2 takes place in the single or double proton transfer process, we constructed the PES of the S1 state as a function of the bond lengths of O1-H1 and O2-H2. The PES is drawn in Figure 4a, where the black and red arrows represent the single ESIPT pathway and double ESIPT pathway, respectively. Importantly, the four stable points can be observed in Figure 4a. The PES of the S1 state is symmetrical along the diagonal curve as a result of the symmetrical structure of BP(OH)2. This means that BP(OH)2 has three stable conformations in the S1 state: the Enol* form, the Keto*1 form and the Keto*2 form. Specifically, in order to achieve the stable Keto*2 form, the double ESIPT pathway needs to surpass an energy barrier from the Enol* form to the Keto*2 form. However, there is no potential barrier in the single ESIPT pathway along with O1-H1…N1 or O2-H2…N2. That is to say, the single ESIPT process will happen spontaneously when forming the Keto*1 conformation upon excitation to S1. Consequently, the BP(OH)2 in hexane will occur during the single ESIPT process instead of the double ESIPT, even though it possesses two IHBs.
We further plotted the PECs in Figure 4b to compare the difference between the double and single proton transfer pathways clearly. The S0 single curve shows that BP(OH)2 exists in the Enol configuration in the S0 state. For the S0 double path, the high energy barrier prevents the double proton transfer behavior from taking place. These results mean that neither the single nor double proton transfer process can occur in the S0 state. In addition, we notice that, for the S1 double path, BP(OH)2 needs to undergo an energy barrier forming two new hydrogen bonds (N1-H1 and N2-H2), while for the S1 single path, there is only one stable structure, and the coordinate of this stable configuration is 1.724 Å. In other words, the BP(OH)2 will take place in a barrierless single proton transfer process. Thus, it can be further attested that the double ESIPT pathway is ruled out due to the existence of a potential barrier.

2.5. Infrared Spectra

IR spectra are found to be an invaluable tool to describe hydrogen-bonding dynamics processes by analyzing the changes in vibration frequency [34]. To affirm our viewpoint once again—that only the single ESIPT occurs in the S1 state—Figure 5 presents IR spectra of BP(OH)2 (scaling factor 0.959) in the S0 and S1 states. The calculated IR spectra range from 2600 cm−1 to 3400 cm−1, which involve vibration peaks for O1-H1, O2-H2 and N1-H1 groups. Specifically, in the S0 state, there is only a characteristic peak around 2869 cm−1 due to the structural symmetry of BP(OH)2, assigned to the O1-H1 and O2-H2 stretching vibration. When excited to S1, the O2-H2 and N1-H1 stretching vibrational frequency of BP(OH)2 are at 2918 cm−1 and 3153 cm−1, respectively. Obviously, the IR spectrum of the S1 state for BP(OH)2 is different from that of the S0 state. The vibration peak of O2-H2 exhibits a blue shift from 2869 cm−1 of the S0 state to 2918 cm−1 of the S1 state. This reveals that, following excitation to S1, the IHB (O2-H2…N2) is weakened, which is not conducive to the transfer of H2 protons. Interestingly, the O1-H1 vibration peak (2869 cm−1) disappears in the S1 state while the new N1-H1 peak (3153 cm−1) forms, indicating the transfer of proton H1 from O1 to N1. Consequently, these results above demonstrate the occurrence of the single proton transfer process upon excitation.

2.6. Excited State Hydrogen Bond Dynamics

Considering that the timescale and characteristics of the ESIPT process may well be theoretically simulated and further revealed by excited-state hydrogen-bonding dynamics [35], the time-evolution trajectories of all bond lengths involved in IHBs are presented in Figure 6a. The Enol* form is the initial photoexcited configuration. In addition, at ~15 fs, the bond length of O1-H1 increases gradually with the time evolution, and the H1-N1 decreases gradually with the time evolution in comparison with the bond lengths of the Enol* form. These changes are due to the fact that the proton H1 moves along with a hydrogen bond from O1 to N1 rapidly until the proton H1 is transferred. After 15 fs, all the bond lengths related to the intramolecular hydrogen bond tend toward dynamic equilibrium. The Keto*1 form after 15 fs can be seen in Figure 6a. By comparing the Enol* form with the Keto*1 form, only the single ESIPT takes place in BP(OH)2 upon excitation to the S1 state.
Due to the intramolecular charge redistribution having momentous effects on the ESIPT behavior, we also calculated the time evolution of the NBO charge for the atoms related to the proton transfer displayed in Figure 6b. For O1-H1…N1, within 15 fs, the negative charge of the N1 atom decreases, while the electronegativity of O1 hardly obviously changes with time evolution. This phenomenon is due to the fact that the positively charged proton H1 gradually moves away from O1 to the N1 atom, which offsets the negative charge on N1. At the same time, the electronegativity of O2 increases gradually with time evolution, but the negative charge of N2 decreases with time evolution. These changes mean that the interaction between O2 and H2 is enhanced and the interaction between N2 and H2 atoms is weakened within 15 fs. Thus, it is not possible to transfer the proton H2 from O2 to N2. The NBO charge of all atoms related to the intramolecular hydrogen bond tends toward dynamic equilibrium after 15 fs. Based on the above analysis, the variety of the NBO charge once again supports the viewpoint that the BP(OH)2 molecule will carry out the single proton transfer reaction instead of the double proton transfer upon excitation.

2.7. Spin-Orbit Coupling Interaction

The BP(OH)2 molecule was found to exhibit not only single proton transfer properties but also the TADF phenomenon [27]. TADF phenomenon is closely correlated with the ISC and RISC processes. According to Fermi’s golden rule [36], the larger SOC constant and the smaller energy level gap can promote the ISC and RISC processes. Therefore, to verify which triplet states are involved in the ISC process and elucidate the mechanism of the delayed fluorescence phenomenon, the energy level gaps and SOC constants are recorded in Table 3. To be specific, the energy level gaps for S1-T1, S1-T2 and S1-T3 are 0.980 eV, 0.120 eV and −0.046 eV, respectively. The energy level gaps of S1-T3 and S1-T2 are smaller than that of S1-T1, which makes the T1 state difficult to induce the RISC process from T1 to S1. Based on that, the SOCISC constants of S1-T3 and S1-T2 are 31.029 cm−1 and 0.411 cm−1. Obviously, the SOCISC constant of S1-T3 is much larger than that of S1-T2. Thus, the energy level gap, combined with spin-orbit coupling interaction, confirms that the ISC process is mainly the transition from the S1 to the T3 state. Correspondingly, it can be found that the SOCRISC constants of S1-T1, S1-T2 and S1-T3 are 0.095 cm−1, 0.585 cm−1 and 0.321 cm−1, respectively. The SOCRISC constants of S1-T3 and S1-T2 are larger than those of T1-S1, implying that the RISC process of T2-S1 and T3-S1 would be realized facilely.
Since differences in the electron-hole distribution between the singlet and triplet states will affect the ISC process [37], we calculated the electron-hole distribution diagrams of the singlet (S1 Keto*1) and triplet (T2 Keto*1 and T3 Keto*1) states, as plotted in Figure 7. In accordance with the El-Sayed rule [38], the transition between identical vibrational states is forbidden. Although the S1, T2 and T3 states are characteristic of charge transfer, the electron-hole distribution of the S1 state is different from that of T2 and T3 states. This means that there are two ISC transition channels, which are from the S1 state to T2 and from the S1 state to T3. During the excitation process, the T3 state undergoes significant electron transfer and drastic alternation of electron distribution structure. In other words, the electron transfer from the S1 state to T3 is more pronounced than that from the S1 state to T2, which is the reason for the large SOCISC value between the S1 and T3 states. However, the degree of electron transfer in the S1 state is close to that in the T2 state, leading to a smaller SOCISC value compared with SOCISC of the S1 and T3 states. Meanwhile, we find from Table 3 that the T2 energy level is rather close to the T3 energy level, which means that there may be an internal conversion process from the T3 state to T2. Thus, it is demonstrated that the T3 energy level plays a significant role in collecting triplet state excitons.
For a more in-depth discussion of the ISC and RISC processes associated with TADF, the ISC and RISC rates of BP(OH)2 in hexane are listed in Table 3. The ISC rates of S1-T1, S1-T2 and S1-T3 were calculated to be 1.40 × 106 s−1, 1.73 × 106 s−1 and 1.05 × 1010 s−1, respectively. The ISC rate of S1-T3 is four orders of magnitude larger than that of S1-T1 and S1-T2 due to the small energy level gap and the large SOC constant. Consequently, the result demonstrates that the ISC process will primarily happen from the S1 state to T3. Furthermore, the RISC rates of S1-T2 and S1-T3 are 3.52 × 106 s−1 and 1.12 × 106 s−1, which are one order of magnitude larger than that of S1-T1. Thus, it is confirmed once again that the RISC process belongs to the transition from the T3 state to S1 and the one from the T2 state to S1. Through the above analysis, our calculated results illustrate that the ISC process is mainly assigned to the transition from the S1 state to T3, and the RISC process in BP(OH)2 belongs to the transition from the T3 state to S1 and the one from the T2 state to S1. In the previous paper of Tokumura et al. [27], for BP(OH)2, the ISC and RISC processes after the double proton transfer are the transition between the S1 and T2 states. In contrast, herein, we reconsider the two processes based on the single proton transfer structure (Keto*1 form).

2.8. TADF Mechanism

Based on the analysis above, we describe the photophysical processes that give rise to the TADF phenomenon in BP(OH)2, with the ESIPT feature displayed in Figure 8. Upon excitation to the S1 state, BP(OH)2 occurs in the intramolecular single proton transfer, forming the S1 Keto*1 structure. Subsequently, the normal emission from the S1 Keto*1 conformation occurs in competition with the efficient ISC process from the S1 Keto*1 state to T3 Keto*1. Following the ISC process, there are two possible RISC pathways: the transition from the T3 Keto*1 state to S1 Keto*1 and the one from the T2 Keto*1 state to S1 Keto*1. The BP(OH)2 molecule that returns to the S1 Keto*1 state decays to the S0 Keto state by emitting delayed fluorescence. In summary, the luminescence is derived from the single ESIPT structure, and the T3 state plays a crucial role in the TADF process due to the large electron-hole separation.

3. Materials and Methods

All calculations in our work were completed using the DFT and TDDFT methods at the B3LYP/def2-TZVP level with the dispersion correction (gd3bj) [39,40,41,42,43,44]. The effect of the hexane solvent was considered with the integral equation formalism variant of the polarizable continuum model [45]. In this way, the optimized geometric structures, absorption and emission properties, FMOs, IR spectra, PECs, PES and excited-state hydrogen-bond dynamics were calculated in the Gaussian 16 package [46]. Vibrational frequencies were computed by determining the second derivatives of the energy with respect to the Cartesian nuclear coordinates and then transforming to mass-weighted coordinates. Concerning the ab initio molecular dynamics study of ESIPT, the classical trajectory calculation (i.e., Born–Oppenheimer molecular dynamics (BOMD)) was undertaken in the S1 state [47,48]. The RDG image and the electron-hole distribution of the singlet and triplet states were analyzed using Multiwfn software [49]. In addition, the PySOC program interfaced with Gaussian 16 software was used to calculate the SOC constants [50,51]. Moreover, the nonradiative decay rates between singlet and triplet states were investigated in MOMAP [52,53,54,55].

4. Conclusions

The ESIPT-based TADF mechanisms of BP(OH)2 were theoretically reconsidered at the B3LYP/def-2TZVP (gd3bj) level. Geometric structure, RDG scatter plots, PES and IR spectra indicate that BP(OH)2 in hexane will undergo the single ESIPT process rather than the double ESIPT upon excitation to S1 state. Furthermore, the BOMD computations prove once again the occurrence of the single ESIPT process and further provide the timescale for this process. Moreover, the energy level gaps, SOC constants and nonradiative decay rates of S1-T2, S1-T2 and S1-T2 were calculated based on the single ESIPT structure (Keto*1 form). These results elucidate that the ISC process is mainly the transition from the S1 Keto*1 state to T3 Keto*1, and the RISC process belongs to the transition from the T3 Keto*1 state to S1 Keto*1 and the one from the T2 Keto*1 state to S1 Keto*1. Furthermore, we find that the different electron-hole distributions between the singlet and triplet states of BP(OH)2 facilitate the ISC process utilizing the electron-hole analysis. Thus, the T3 state plays a dominant role in the TADF process, which is induced by the significant electron-hole separation between S1 and T3 states. Our work will not only be helpful to further understand the interaction between ESIPT and TADF but also contribute to achieving high luminescence efficiency for organic molecules with TADF properties.

Author Contributions

Data curation, X.Z.; Formal analysis, X.Z. and Q.L.; Funding acquisition, H.Y. and Y.S.; Methodology, H.Y.; Software, X.Z. and L.Z.; Supervision, H.Y. and Y.S.; Validation, H.Y.; Writing—original draft, X.Z.; Writing—review and editing, H.Y. and Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by [the National Natural Science Foundation of China] grant number [12174149 and 11874180].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interest.

References

  1. Vazquez, R.J.; Yun, J.H.; Muthike, A.K.; Howell, M.; Kim, H.; Madu, I.K.; Kim, T.; Zimmerman, P.; Lee, J.Y.; Iii, T.G. New Direct Approach for Determining the Reverse Intersystem Crossing Rate in Organic Thermally Activated Delayed Fluorescent (TADF) Emitters. J. Am. Chem. Soc. 2020, 142, 8074–8079. [Google Scholar] [CrossRef]
  2. Bryden, M.A.; Zysman-Colman, E. Organic thermally activated delayed fluorescence (TADF) compounds used in photocatalysis. Chem. Soc. Rev. 2021, 50, 7587–7680. [Google Scholar] [CrossRef] [PubMed]
  3. Chen, X.K.; Kim, D.; Bredas, J.L. Thermally Activated Delayed Fluorescence (TADF) Path toward Efficient Electroluminescence in Purely Organic Materials: Molecular Level Insight. Acc. Chem. Res. 2018, 51, 2215–2224. [Google Scholar] [CrossRef] [PubMed]
  4. Gibson, J.; Penfold, T.J. Nonadiabatic coupling reduces the activation energy in thermally activated delayed fluorescence. Phys. Chem. Chem. Phys. 2017, 19, 8428–8434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Sheng, R.; Yang, L.; Li, A.; Chen, K.; Zhang, F.; Duan, Y.; Zhao, Y.; Chen, P. Highly efficient orange and white OLEDs based on ultrathin phosphorescent emitters with double reverse intersystem crossing system. J. Lumin. 2022, 246, 118852. [Google Scholar] [CrossRef]
  6. Liu, H.; Liu, Z.; Li, G.; Huang, H.; Zhou, C.; Wang, Z.; Yang, C. Versatile Direct Cyclization Constructs Spiro-acridan Derivatives for Highly Efficient TADF emitters. Angew. Chem. Int. Ed. 2021, 60, 12376–12380. [Google Scholar] [CrossRef]
  7. Fan, J.; Zhang, Y.; Zhou, Y.; Lin, L.; Wang, C.-K. Excited State Properties of a Thermally Activated Delayed Fluorescence Molecule in Solid Phase Studied by Quantum Mechanics/Molecular Mechanics Method. J. Phys. Chem. C 2018, 122, 2358–2366. [Google Scholar] [CrossRef]
  8. Hu, Y.-Y.; Luo, W.; Hu, C.-G.; Wang, Y.; Tong, B.-H.; Fung, M.-K.; Tian, Y.-P.; Zhang, Q.-F. The one-pot synthesis of homoleptic phenylphthalazine iridium(III) complexes and their application in high efficiency OLEDs. J. Lumin. 2020, 219, 116846. [Google Scholar] [CrossRef]
  9. Fan, J.; Zhang, Y.; Ma, Y.; Song, Y.; Lin, L.; Xu, Y.; Wang, C.-K. The role of intermolecular interactions in regulating the thermally activated delayed fluorescence and charge transfer properties: A theoretical perspective. J. Mater. Chem. C 2020, 8, 8601–8612. [Google Scholar] [CrossRef]
  10. Lin, L.; Wang, Z.; Fan, J.; Wang, C. Theoretical insights on the electroluminescent mechanism of thermally activated delayed fluorescence emitters. Org. Electron. 2017, 41, 17–25. [Google Scholar] [CrossRef]
  11. Dos Santos, P.L.; Etherington, M.K.; Monkman, A.P. Chemical and conformational control of the energy gaps involved in the thermally activated delayed fluorescence mechanism. J. Mater. Chem. C 2018, 6, 4842–4853. [Google Scholar] [CrossRef]
  12. Zhou, D.; Liu, D.; Gong, X.; Ma, H.; Qian, G.; Gong, S.; Xie, G.; Zhu, W.; Wang, Y. Solution-Processed Highly Efficient Bluish-Green Thermally Activated Delayed Fluorescence Emitter Bearing an Asymmetric Oxadiazole-Difluoroboron Double Acceptor. ACS Appl. Mater. Interfaces 2019, 11, 24339–24348. [Google Scholar] [CrossRef] [PubMed]
  13. Azarias, C.; Budzak, S.; Laurent, A.D.; Ulrich, G.; Jacquemin, D. Tuning ESIPT fluorophores into dual emitters. Chem. Sci. 2016, 7, 3763–3774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Li, Y.; Sun, C.; Han, J.; Zhou, Q.; Cao, B.; Yin, H.; Shi, Y. Theoretical investigation of intermolecular hydrogen bond induces fluorescence quenching phenomenon for Coumarin-1. J. Lumin. 2020, 221, 117110. [Google Scholar] [CrossRef]
  15. Sun, C.; Zhang, X.; Diao, L.; Cao, B.; Yin, H.; Shi, Y. How the atomic electron-accepting ability affect the double ESIPT process of 2,5-bis(benzoxazol-2-yl)thiophene-3,4-diol? J. Lumin. 2020, 225, 117329. [Google Scholar] [CrossRef]
  16. Yin, H.; Zhang, Y.-M.; Zhao, H.-F.; Yang, G.; Shi, Y.; Zhang, S.X.-A.; Ding, D.-J. Optical anti-counterfeiting of a single molecule by two solvents based on intra- and intermocular excited state proton transfer mechanisms. Dyes Pigments 2018, 159, 506–512. [Google Scholar] [CrossRef]
  17. Zhao, J.; Dong, H.; Zheng, Y. Elaborating the excited state multiple proton transfer mechanism for 9Hpyrido[3,4-b]indole. J. Lumin. 2018, 195, 228–233. [Google Scholar] [CrossRef]
  18. Zhu, L.; Zhou, Q.; Cao, B.; Li, B.; Wang, Z.; Zhang, X.; Yin, H.; Shi, Y. Theoretical reconsideration of the mechanism of the excited state proton transfer of indigo carmine in water. J. Mol. Liq. 2022, 347, 118365. [Google Scholar] [CrossRef]
  19. Yang, Y.; Zhao, J.; Li, Y. Theoretical Study of the ESIPT Process for a New Natural Product Quercetin. Sci. Rep. 2016, 6, 32152. [Google Scholar] [CrossRef] [Green Version]
  20. Song, P.; Ma, F.-C. Intermolecular hydrogen-bonding effects on photophysics and photochemistry. Int. Rev. Phys. Chem. 2013, 32, 589–609. [Google Scholar] [CrossRef]
  21. Yin, H.; Shi, Y.; Wang, Y. Time-dependent density functional theory study on the excited-state intramolecular proton transfer in salicylaldehyde. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2014, 129, 280–284. [Google Scholar] [CrossRef] [PubMed]
  22. Sun, L.; Chen, Y.; Sun, M. Exploring Nonemissive Excited-State Intramolecular Proton Transfer by Plasmon-Enhanced Hyper-Raman Scattering and Two-Photon Excitation Fluorescence. J. Phys. Chem. C 2021, 126, 487–492. [Google Scholar] [CrossRef]
  23. Yang, Y.; Zhang, Q.; Liu, Y.; Jiang, Z.; Qin, C.; Jiang, K.; Liu, Y. Ultrafast dynamics of dual fluorescence of 2-(2′-hydroxyphenyl) benzothiazole and its derivatives by femtosecond transient absorption spectroscopy. J. Lumin. 2022, 248, 118922. [Google Scholar] [CrossRef]
  24. Cao, Y.; Eng, J.; Penfold, T.J. Excited State Intramolecular Proton Transfer Dynamics for Triplet Harvesting in Organic Molecules. J. Phys. Chem. A 2019, 123, 2640–2649. [Google Scholar] [CrossRef] [PubMed]
  25. Wu, K.; Zhang, T.; Wang, Z.; Wang, L.; Zhan, L.; Gong, S.; Zhong, C.; Lu, Z.H.; Zhang, S.; Yang, C. De Novo Design of Excited-State Intramolecular Proton Transfer Emitters via a Thermally Activated Delayed Fluorescence Channel. J. Am. Chem. Soc. 2018, 140, 8877–8886. [Google Scholar] [CrossRef] [PubMed]
  26. Bulska, H. Intramolecular cooperative double proton transfer in [2,2′-bipyridyl]-3,3′-diol. Chem. Phys. Lett. 1983, 98, 398–402. [Google Scholar] [CrossRef]
  27. Kunihiro, T.; Mutsuo, K.; Osamu, O. Reverse intersystem crossing from higher triplet to excited singlet in 2,2’-bipyridine-3,3’-diol phototautomer. J. Photochem. Photobiol. A Chem. 1994, 81, 151–158. [Google Scholar] [CrossRef]
  28. Plasser, F.; Barbatti, M.; Aquino, A.J.A.; Lischka, H. Excited-state diproton transfer in [2,2′-bipyridyl]-3,3′-diol: The mechanism is sequential, not concerted. J. Phys. Chem. A 2009, 113, 8490–8499. [Google Scholar] [CrossRef]
  29. Zhao, J.; Liu, X.; Zheng, Y. Controlling Excited State Single versus Double Proton Transfer for 2,2’-Bipyridyl-3,3′-diol: Solvent Effect. J. Phys. Chem. A 2017, 121, 4002–4008. [Google Scholar] [CrossRef]
  30. Chen, K.; Kurganskii, I.V.; Zhang, X.; Elmali, A.; Zhao, J.; Karatay, A.; Fedin, M.V. Intersystem Crossing and Electron Spin Selectivity in Anthracene-Naphthalimide Compact Electron Donor-Acceptor Dyads Showing Different Geometry and Electronic Coupling Magnitudes. Chem. Eur. J. 2021, 27, 7572–7587. [Google Scholar] [CrossRef]
  31. Xu, C.; Yu, L.; Zhu, C.; Yu, J.; Cao, Z. Intersystem crossing-branched excited-state intramolecular proton transfer for o-nitrophenol: An ab initio on-the-fly nonadiabatic molecular dynamic simulation. Sci. Rep. 2016, 6, 26768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Wang, W.; Marshall, M.; Collins, E.; Marquez, S.; Mu, C.; Bowen, K.H.; Zhang, X. Intramolecular electron-induced proton transfer and its correlation with excited-state intramolecular proton transfer. Nat. Commun. 2019, 10, 1170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Chen, J.; Yu, T.; Ubba, E.; Xie, Z.; Yang, Z.; Zhang, Y.; Liu, S.; Xu, J.; Aldred, M.P.; Chi, Z. Achieving Dual-Emissive and Time-Dependent Evolutive Organic Afterglow by Bridging Molecules with Weak Intermolecular Hydrogen Bonding. Adv. Opt. Mater. 2019, 7, 1801593. [Google Scholar] [CrossRef]
  34. Samant, V.; Singh, A.K.; Ramakrishna, G.; Ghosh, H.N.; Ghanty, T.K.; Palit, D.K. Ultrafast Intermolecular Hydrogen Bond Dynamics in the Excited State of Fluorenone. J. Phys. Chem. A 2005, 109, 8693–8704. [Google Scholar] [CrossRef] [PubMed]
  35. Dommett, M.; Crespo-Otero, R. Excited state proton transfer in 2′-hydroxychalcone derivatives. Phys. Chem. Chem. Phys. 2017, 19, 2409–2416. [Google Scholar] [CrossRef]
  36. Lawetz, V.; Orlandi, G.; Siebrand, W. Theory of Intersystem Crossing in Aromatic Hydrocarbons. J. Chem. Phys. 1972, 56, 4058–4072. [Google Scholar] [CrossRef]
  37. Jiang, G.; Li, F.; Fan, J.; Song, Y.; Wang, C.-K.; Lin, L. Theoretical perspective for luminescent mechanism of thermally activated delayed fluorescence emitter with excited-state intramolecular proton transfer. J. Mater. Chem. C 2020, 8, 98–108. [Google Scholar] [CrossRef]
  38. Samanta, P.K.; Kim, D.; Coropceanu, V.; Bredas, J.L. Up-Conversion Intersystem Crossing Rates in Organic Emitters for Thermally Activated Delayed Fluorescence: Impact of the Nature of Singlet vs Triplet Excited States. J. Am. Chem. Soc. 2017, 139, 4042–4051. [Google Scholar] [CrossRef]
  39. Zhang, M.; Ren, B.; Wang, Y.; Zhao, C. A DFT/TDDFT study on the excited-state hydrogen bonding dynamics of 6-aminocoumarin in water solution. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 101, 191–195. [Google Scholar] [CrossRef]
  40. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  41. Lanke, S.K.; Sekar, N. Pyrazole based NLOphores: Synthesis, photophysical, DFT, TDDFT studies. Dyes Pigments 2016, 127, 116–127. [Google Scholar] [CrossRef]
  42. Salomon, O.; Reiher, M.; Hess, B.A. Assertion and validation of the performance of the B3LYP⋆ functional for the first transition metal row and the G2 test set. J. Chem. Phys. 2002, 117, 4729–4737. [Google Scholar] [CrossRef]
  43. Navarro, L.; Rodriguez, F.; Cirera, J. Controlling the spin-crossover behavior of the [Cr(indenyl)2] family via ligand functionalization. Dalton Trans. 2021, 50, 8704–8710. [Google Scholar] [CrossRef] [PubMed]
  44. Chrayteh, M.; Huet, T.R.; Drean, P. Microsolvation of myrtenal studied by microwave spectroscopy highlights the role of quasi-hydrogen bonds in the stabilization of its hydrates. J. Chem. Phys. 2020, 153, 104304. [Google Scholar] [CrossRef] [PubMed]
  45. Mennucci, B.; Cancès, E.; Tomasi, J. Evaluation of Solvent Effects in Isotropic and Anisotropic Dielectrics and in Ionic Solutions with a Unified Integral Equation Method: Theoretical Bases, Computational Implementation, and Numerical Applications. J. Phys. Chem. B 1997, 101, 10506–10517. [Google Scholar] [CrossRef]
  46. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16, Revision B 01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  47. Nguyen, P.D.; Ding, F.; Fischer, S.A.; Liang, W.; Li, X. Solvated First-Principles Excited-State Charge-Transfer Dynamics with Time-Dependent Polarizable Continuum Model and Solvent Dielectric Relaxation. J. Phys. Chem. Lett. 2012, 3, 2898–2904. [Google Scholar] [CrossRef]
  48. Yin, H.; Li, B.; Zhao, X.; Liu, Y.; Shi, Y.; Ding, D. Restriction of intramolecular torsion induces abnormal blue-shifted fluorescence in the aggregate state. Dyes Pigments 2022, 201, 110192. [Google Scholar] [CrossRef]
  49. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef]
  50. Gao, X.; Bai, S.; Fazzi, D.; Niehaus, T.; Barbatti, M.; Thiel, W. Evaluation of Spin-Orbit Couplings with Linear-Response Time-Dependent Density Functional Methods. J. Chem. Theory Comput. 2017, 13, 515–524. [Google Scholar] [CrossRef] [Green Version]
  51. Li, Q.; Wan, Y.; Zhou, Q.; Li, Y.; Li, B.; Zhu, L.; Wan, Y.; Yin, H.; Shi, Y. Exploring the effect of nitrile substituent position on fluorescence quantum yield of ESIPT-based oxazoline derivatives: A TDDFT investigation. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2022, 272, 120953–120959. [Google Scholar] [CrossRef]
  52. Niu, Y.; Li, W.; Peng, Q.; Geng, H.; Yi, Y.; Wang, L.; Nan, G.; Wang, D.; Shuai, Z. MOlecular MAterials Property Prediction Package (MOMAP) 1.0: A software package for predicting the luminescent properties and mobility of organic functional materials. Mol. Phys. 2018, 116, 1078–1090. [Google Scholar] [CrossRef]
  53. Peng, Q.; Yi, Y.; Shuai, Z.; Shao, J. Toward Quantitative Prediction of Molecular Fluorescence Quantum Efficiency: Role of Duschinsky Rotation. J. Am. Chem. Soc. 2007, 129, 9333–9339. [Google Scholar] [CrossRef] [PubMed]
  54. Niu, Y.; Peng, Q.; Shuai, Z. Promoting-mode free formalism for excited state radiationless decay process with Duschinsky rotation effect. Sci. China Chem. 2008, 51, 1153–1158. [Google Scholar] [CrossRef]
  55. Shuai, Z. Thermal Vibration Correlation Function Formalism for Molecular Excited State Decay Rates. Chin. J. Chem. 2020, 38, 1223–1232. [Google Scholar] [CrossRef]
Figure 1. The optimized structure of BP(OH)2 in the S0 and S1 states at the B3LYP/def2TZVP (gd3bj) IEFPCM (Hexane) level. Enol form: the initial conformation at the S0 state; Keto*1 form: the conformation based on single ESIPT in the S1 state; Keto*2 form: the conformation based on double ESIPT in the S1 state. The blue atom represents N, the cyan atom represents C, the grey atom represents H, and red atom represents O.
Figure 1. The optimized structure of BP(OH)2 in the S0 and S1 states at the B3LYP/def2TZVP (gd3bj) IEFPCM (Hexane) level. Enol form: the initial conformation at the S0 state; Keto*1 form: the conformation based on single ESIPT in the S1 state; Keto*2 form: the conformation based on double ESIPT in the S1 state. The blue atom represents N, the cyan atom represents C, the grey atom represents H, and red atom represents O.
Ijms 23 13969 g001
Figure 2. The HOMO and LUMO of BP(OH)2 in hexane solvent.
Figure 2. The HOMO and LUMO of BP(OH)2 in hexane solvent.
Ijms 23 13969 g002
Figure 3. Diagrams of RDG (Function 1) vs. Ω(r) (Function 2) and gradient isosurfaces of BP(OH)2 at Enol, Keto*1 and Keto*2 states.
Figure 3. Diagrams of RDG (Function 1) vs. Ω(r) (Function 2) and gradient isosurfaces of BP(OH)2 at Enol, Keto*1 and Keto*2 states.
Ijms 23 13969 g003
Figure 4. Constructed (a) PES in the S1 state and (b) PECs in the S0 and S1 states for BP(OH)2.
Figure 4. Constructed (a) PES in the S1 state and (b) PECs in the S0 and S1 states for BP(OH)2.
Ijms 23 13969 g004
Figure 5. Calculated S0 state and S1 state infrared spectrum based on the single ESIPT structure of BP(OH)2 in hexane.
Figure 5. Calculated S0 state and S1 state infrared spectrum based on the single ESIPT structure of BP(OH)2 in hexane.
Ijms 23 13969 g005
Figure 6. (a) The important bond lengths with time evolution, including the Enol* form (left) and Keto*1 form (right) and (b) the NBO charge of primary atoms with time evolution for BP(OH)2, including the Enol* form (left) and Keto*1 form (right). Atoms of different colors in the Enol* and Keto* are represented as in Figure 1.
Figure 6. (a) The important bond lengths with time evolution, including the Enol* form (left) and Keto*1 form (right) and (b) the NBO charge of primary atoms with time evolution for BP(OH)2, including the Enol* form (left) and Keto*1 form (right). Atoms of different colors in the Enol* and Keto* are represented as in Figure 1.
Ijms 23 13969 g006
Figure 7. Electron-hole distribution based on the Keto*1 form in the excited states for BP(OH)2. Blue color and green color represent holes and electrons, respectively. The golden atom represents C, the blue atom represents N, the red atom represents O, and light grey atom represents H.
Figure 7. Electron-hole distribution based on the Keto*1 form in the excited states for BP(OH)2. Blue color and green color represent holes and electrons, respectively. The golden atom represents C, the blue atom represents N, the red atom represents O, and light grey atom represents H.
Ijms 23 13969 g007
Figure 8. The schematic diagram of photoluminescence processes for BP(OH)2 in hexane solvent.
Figure 8. The schematic diagram of photoluminescence processes for BP(OH)2 in hexane solvent.
Ijms 23 13969 g008
Table 1. Obtained primary bond lengths for the optimized structure of BP(OH)2 at the S0 and S1 states (Unit Å).
Table 1. Obtained primary bond lengths for the optimized structure of BP(OH)2 at the S0 and S1 states (Unit Å).
Unit ÅEnolKeto*1Keto*2
N1-H1 (Å)1.6591.0261.034
O1-H1 (Å)1.0051.7881.725
N2-H2 (Å)1.6591.6861.034
O2-H2 (Å)1.0001.0001.724
Table 2. The absorption and emission peaks of BP(OH)2 in hexane solvent.
Table 2. The absorption and emission peaks of BP(OH)2 in hexane solvent.
Abs. (nm)Fluo. (nm)
Exp.345499
Theor.337(Enol)484(Keto*1)479(Keto*2)
Table 3. Calculated SOC constants, energy level difference and rates of ISC (KISC) and RISC (KRISC) between the singlet and triplet states of BP(OH)2 in hexane.
Table 3. Calculated SOC constants, energy level difference and rates of ISC (KISC) and RISC (KRISC) between the singlet and triplet states of BP(OH)2 in hexane.
S1-T1S1-T2S1-T3
ΔE (eV)0.9800.120−0.046
SOCISC (cm−1)0.1520.41131.029
SOCRISC (cm−1)0.0950.5850.321
KISC (s−1)1.40 × 1061.73 × 1061.05 × 1010
KRISC (s−1)5.48 × 1053.52 × 1061.12 × 106
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, X.; Zhu, L.; Li, Q.; Yin, H.; Shi, Y. The Interplay between ESIPT and TADF for the 2,2′-Bipyridine-3,3′-diol: A Theoretical Reconsideration. Int. J. Mol. Sci. 2022, 23, 13969. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232213969

AMA Style

Zhao X, Zhu L, Li Q, Yin H, Shi Y. The Interplay between ESIPT and TADF for the 2,2′-Bipyridine-3,3′-diol: A Theoretical Reconsideration. International Journal of Molecular Sciences. 2022; 23(22):13969. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232213969

Chicago/Turabian Style

Zhao, Xin, Lixia Zhu, Qi Li, Hang Yin, and Ying Shi. 2022. "The Interplay between ESIPT and TADF for the 2,2′-Bipyridine-3,3′-diol: A Theoretical Reconsideration" International Journal of Molecular Sciences 23, no. 22: 13969. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms232213969

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop