Abstract
Thermally activated delayed fluorescence (TADF) materials and multi-resonant (MR) variants are promising organic emitters that can achieve an internal electroluminescence quantum efficiency of ~100%. The reverse intersystem crossing (RISC) is key for harnessing triplet energies for fluorescence. Theoretical modeling is thus crucial to estimate its rate constant (kRISC) for material development. Here, we present a comprehensive assessment of the theory for simulating the RISC of MR-TADF molecules within a perturbative excited-state dynamics framework. Our extended rate formula reveals the importance of the concerted effects of nonadiabatic spin-vibronic coupling and vibrationally induced spin-orbital couplings in reliably determining kRISC of MR-TADF molecules. The excited singlet-triplet energy gap is another factor influencing kRISC. We present a scheme for gap estimation using experimental Arrhenius plots of kRISC. Erroneous behavior caused by approximations in Marcus theory is elucidated by testing 121 MR-TADF molecules. Our extended modeling offers in-depth descriptions of kRISC.
Converted spin-vibronic coupling effects enhance internal quantum transfer processes for molecular OLED materials.
INTRODUCTION
Organic light-emitting diodes (OLEDs) (1) have gained widespread interest owing to their remarkable features such as lightness and wide viewing range. Materials that emit light through fluorescence (2, 3) and phosphorescence (4–6) have been developed for their application in OLEDs. Electrical excitations that occur in OLED devices produce singlet and triplet excitons in a ratio of 1:3 through hole-electron recombination (4).
To increase the internal quantum efficiency (IQE) to ∼100%, all generated excitons with different spin multiplicities have to be used for electroluminescence. To meet this demand, Adachi and co-workers developed OLEDs based on thermally activated delayed fluorescence (TADF) (7). TADF is a process in which a nonradiative triplet exciton is converted into a radiative singlet exciton via reverse intersystem crossing (RISC), followed by delayed fluorescence (Fig. 1A). TADF materials show marked potential for fully using singlet and triplet excitons for fluorescence; materials have been reported to achieve nearly 100% IQE (8–10). Recently, TADF coupled with the multiple resonance (MR) effect, designated MR-TADF (11), has received considerable attention as synthesized MR-TADF luminophores demonstrate desirable narrowband emissions with high IQE and color purity (12–14).
Fig. 1. The reverse intersystem crossing process in TADF materials.
(A) Conceptual diagram of reverse intersystem crossing (RISC) process in TADF materials. (B) Treatments of RISC process using 1st+Condon, 2nd+Condon, 1st+HT, and 2nd+HT TVCF methods. Compared to the first method, the second approach additionally accounts for nonadiabatic spin-vibronic coupling (NA-SVC). The Herzberg-Teller (HT) treatment considers HT-SVC, which is discarded by the Condon treatment. (C) MR-TADF molecules: ν-DABNA, BOBO-Z (X1=X2=O), BOBS-Z (X1=O, X2=S), and BSBS-Z (X1=X2=S). (D) Arrhenius plots and activation energies obtained using the relation .
A large spin-orbit coupling (SOC) and small adiabatic singlet-triplet (ST) energy difference ΔEST are important physical factors for accelerating the rate-limiting RISC process. Theoretical modeling will assist in characterizing details of the RISC process and realizing chemical fine-tuning. The semiclassical Marcus formula (15) has been conventionally used to estimate the rate constant kRISC and can describe state-transition processes based on ΔEST and SOC (16–19).
Recent studies have shown that the effects of spin-vibronic coupling (SVC) (20–28), which are discarded in Marcus theory, are impactful to modulating kRISC for TADF molecules (24, 25, 29, 30).
Two types of SVC affect the predicted value of kRISC. The first is expressed through Fermi’s golden rule as the second-order vibronic perturbation term (20–25) and is hereafter referred to as nonadiabatic SVC (NA-SVC). It is caused by the interplay between nonadiabatic coupling (NAC) and SOC. The second is the vibrationally induced SOC, which is related to the first derivatives of the SOC (20, 26–28, 30). This effect is called Herzberg-Teller SVC (HT-SVC).
Several previous studies (21–25) have confirmed that the NA-SVC effect can greatly enhance kRISC by several orders of magnitude through indirect spin-flipping mediated by intermediate excited states such as T2 and T3. Moreover, the HT-SVC effect coupled with spin conversion is considered crucial for further influencing kRISC. Marian and co-workers investigated critical effects of HT-SVC on spin-flipping processes (20, 26–28, 30). A recent study revealed that the simulated kRISC for MR-TADF emitters could increase by more than one order of magnitude when incorporating HT-SVC into the rate formula (29). The derived formula was based on the direct sum of the individual rate contributions associated with the NA-SVC and HT-SVC separately. To the best of our knowledge, no existing rate formula considers the fully concerted coupling of both effects. Developing such a formula could help capture additional RISC channels opened by the HT-SVC effects coupled with the indirect spin-flipping process through low-lying intermediate Tn states (Fig. 1B).
The accuracy of ΔEST is crucial because kRISC steeply varies depending on ΔEST, as in Marcus theory. There are two types of experimental approaches for measuring ΔEST: (i) fluorescence and phosphorescence energy difference and (ii) activation energy of the RISC process (31). In the latter, the activation energy is determined from the slope of the Arrhenius plot of the measured kRISC versus temperature T using the Arrhenius equation . is then interpreted as ΔEST (7, 31). Computational quantum chemical approaches to gauge ΔEST should be an effective alternative; however, there are uncontrollable difficulties in reliably predicting small ΔEST for TADF molecules using the time-dependent density functional theory (TDDFT). More computationally demanding treatments involving correlated wave function theories have been shown to mitigate these challenges (32–34).
In this study, we aimed to elucidate in-depth aspects of the theory for simulating kRISC for MR-TADF materials. A simplified excited-state dynamics approach is desirable for MR-TADF molecules with large π-conjugated systems. Thus, a perturbative treatment based on Fermi’s golden rule was considered. In addition, for their construction, the electronic-state potential energy surfaces were assumed to be harmonic in all dimensions. To facilitate formulation and computation, the thermal vibration correlation function (TVCF) formalism (21, 35) was used. Our major interest was to investigate the influence of the treatment of NA-SVC and HT-SVC and their concerted effects on kRISC. This fully concerted SVC effect is denoted 2nd+HT (Fig. 1B). We developed an enhanced TVCF-based method considering the 2nd+HT effect (section S7).
Furthermore, we investigated the choice of ΔEST, which can be determined using experimental data or quantum chemical computations. We also introduced an approach for estimating ΔEST based on fitting the experimental Arrhenius plots. In addition, the reliability of Marcus theory was assessed by contrasting TVCF predictions for 121 MR-TADF molecules. We discussed approximation errors in Marcus theory.
RESULTS
Calculation of RISC rate constant using experimental as ΔEST
ν-DABNA (36) and its element-substituted analogs, BOBO-Z, BOBS-Z, and BSBS-Z (37), are state-of-the-art MR-TADF materials with narrowband blue emissions (Fig. 1C).
We performed kRISC calculations on these compounds as test cases. We assessed the predictions of the four levels of the TVCF theory, termed as 1st+Condon, 2nd+Condon, 1st+HT, and 2nd+HT (Fig. 1B); see the “Formulation of 2nd+HT TVCF Theory” section, and Marcus theory.
The rate equations of all these theories commonly form as a function of ΔEST and T. The ΔEST can be considered a free parameter. We set ΔEST to the measured corresponding to the slope of the experimental Arrhenius plots (Fig. 1D) and T to 300 K. TVCF simulation requires harmonic vibrational wave functions, which were computed at the DFT level. The results are shown in Fig. 2 (A and B) (see also table S1).
Fig. 2. Simulated RISC rate constants for four TADF molecules.
(A and B) kRISC calculated by various rate constant formulas using as ΔEST, and the ratio of the predicted kRISC to the experimental value for BXBX-Z and ν-DABNA. (C) Overview of the Arrhenius plot slope fitting (ARPSfit) method to determine with which the rate constant calculation can reproduce the experimental Arrhenius plot with its slope . (D and E) kRISC calculated by various rate constant formulas using obtained by the ARPSfit method, and the ratio of the predicted kRISC to the experimental value. (F) Contributions of the direct process (T1) and indirect processes mediated by the intermediate states, T2, T3, and T4, to kRISC.
Figure 2 (A and B) shows that, overall, the 2nd+Condon and 1st+HT methods predicted larger kRISC values than the 1st+Condon method, showing good agreement with the experimental values. This result indicates that considering either the NA-SVC or HT-SVC effects substantially improves the accuracy. The 2nd+HT model, considering both effects, provided the most accurate results within one order of magnitude of error relative to the experimental values. A comparison of the 1st+Condon and 1st+HT results revealed that kRISC was increased by the HT-SVC effect itself by a factor of 10 to 1000. The notable differences between the 1st+HT and 2nd+HT predictions confirmed the importance of the NA-SVC effect, which was most pronounced in BOBO-Z and ν-DABNA. The impact was not so large for either BOBS-Z or BSBS-Z. The former feature could be explained by the fact that ΔET1T2 was small (see also section S2).
Although the 2nd+HT model showed promising improvements, it still performed poorly in some cases. For BOBO-Z and ν-DABNA, the ratio of the predicted kRISC at the 2nd+HT level to the experimental value was 0.16 and 0.09, respectively. These ratios were much smaller in an unbalanced manner than those for BOBS-Z and BSBS-Z, which were 1.2 and 0.93, respectively (Fig. 2B). We speculated that these errors arose from the condition that ΔEST was set to . This setup for ΔEST was used in (7, 31).
Calculation of the RISC rate constant using ΔEST from slope fitting method
The observation shown above prompted us to explore an alternative strategy to define the relationship between the experimental and parameter ΔEST in the rate equations. The core of our scheme designated “Arrhenius plot slope fitting (ARPSfit)” is computationally simulating the Arrhenius plots by evaluating the rate equation as a function of 1/T for a given ΔEST. Let the slope of the simulated Arrhenius plots be denoted , which varies depending on the given value of ΔEST (Fig. 2C). In the ARPSfit method, the objective value of ΔEST is determined such that the associated matches the preknowledge . The ΔEST provided by the ARPSfit method is hereafter denoted (Fig. 2C). To obtain a set of Arrhenius plots, the TVCF (or Marcus) formula was evaluated at 11 points of T in the range of 200 to 300 K. Note that ARPSfit coupled with TVCF requires harmonic vibrational analysis.
The resulting for the MR-TADF systems are summarized in table S3. The for BOBO-Z and ν-DABNA were predicted to be 0.035 and 0.037 eV, respectively, at the 2nd+HT TVCF level, which were smaller than the corresponding , 0.102 and 0.087 eV, respectively.
By substituting into the rate equations, we reevaluated kRISC (Fig. 2, D and E, and table S4). A comparison between Fig. 2 (A and D) (or Fig. 2, B and E) enabled us to examine the influence of the redefined ΔEST on kRISC. The use of the ARPSfit-based effectively refined the RISC rate constant predictions for all cases compared to those obtained with the previous setting . For BOBO-Z, BOBS-Z, BSBS-Z, ν-DABNA, the ratio of the 2nd+HT-based kRISC prediction using to the experimental value was 0.93, 1.8, 1.0, and 0.26, respectively, indicating a similar error order (Fig. 2E). Substantial improvements were observed in the Marcus results. These results demonstrated that the ARPSfit method offered a marked improvement in determining ΔEST from .
As observed previously, the NA-SVC contribution appears to be prominent for BOBO-Z and ν-DABNA. To obtain detailed insights, we dissected kRISC into state-wise contributions (Fig. 2F). Details of this analysis are presented in section S2.2. The RISC transition for BOBS-Z and BSBS-Z was 97% or more, characterized as a direct T1 → S1 process and almost unaffected by NA-SVC treatment. Conversely, a strong NA-SVC effect arose in BOBO-Z, in which the intermediate T2 and T3 states made large contributions via NA-SVC. Similarly, for ν-DABNA, the T2-mediated NA-SVC channel accelerated the RISC process by a factor of 6.3, relative to the 1st+Condon/HT prediction. Figure 2E indicates that treatment at the 2nd+HT level was necessary to quantitatively predict kRISC. This indicates that the NA-SVC and HT-SVC effects manifested in the intermediate Tn states are adequately coupled.
Unexpectedly, the Marcus predictions were greatly improved using , closely approaching the experimental values, even for molecules where the NA-SVC effect discarded in Marcus formula was relevant.
Section S3 discusses the first-principles calculations of ΔEST using various levels of theory, PBE0/def2-SV(P) (38, 39), double-hybrid variant B2PLYP (40)/cc-pVDZ, wave function–based SCS-CC2 (41)/cc-pVDZ and its linear transformation (LT) correction method (34). The kRISC was evaluated using these ΔEST.
We showed that the first-principles wave function methods, specifically SCS-CC2 and its LT methods, offered reliable on the predictions of kRISC. The resulting kRISC values reproduced the experimental values within only a 10-fold error and consistent trends across molecules (fig. S3); for example, the 2nd+HT TVCF prediction at the SCS-CC2 LT level for BSBS-Z (4.85 × 105 s−1) was in good agreement with the experimental kRISC (1.60 × 106 s−1). It is suggested that the first-principles calculations can be used for prescreening for molecular development where the experimental ΔEST is inaccessible.
Comparison of 1st+Condon TVCF and semiclassical Marcus formula
The Marcus results provided relatively good accuracy and better agreement with the experimental values than the 1st+Condon predictions (Fig. 2, D and E). This is remarkable because it contradicts the theoretical construction. Marcus formula corresponds to a model downgraded from the 1st+Condon theory. As discussed in the “Comparison of the 1st+Condon TVCF method and Marcus formula” section, the relationship between the two models is clearly formulated through four approximation steps (42): (i) Duschinsky rotation cutoff (Dus-off), (ii) displaced harmonic oscillator (DHO), (iii) short-time (st), and (iv) high-temperature (ht) approximations. These procedures result in five hierarchical levels of rate-constant equations (Fig. 3A).
Fig. 3. Rate constant equations based on 1st+Condon TVCF theory and its approximations.
(A) Five hierarchical levels of the rate constant equations, 1st+Condon TVCF, Duschinsky rotation cutoff (Dus-off), displaced harmonic oscillator (DHO), and short-time (st) approximated models, and Marcus theory from high-temperature (ht) approximation. (B) Plots of the real part of the correlation functions Re[ρ(t)] as a function of the time t evaluated at five levels of the theory for BOBO-Z. (C) Plots of kRISC as a function of ∆EST evaluated at five levels of the theory for BOBO-Z. The gray vertical dashed line indicates the ARPSfit-based determined at 1st+HT TVCF level. The red dot is the experimental value of kRISC.
To determine why the Marcus model yielded such remarkable results, we numerically examined how these approximations affected the kRISC computation. Here, let us focus on BOBO-Z. The results for the other molecules are presented in section S4. Figure 3B shows the real part of the correlation functions ρ(t) (Eq. 10) at individual levels (e.g., Eq. 19B). Applying a Fourier transform to these equations (Eq. 8) yields kRISC (e.g., Eq. 19A), the values of which are displayed as a function of ΔEST in Fig. 3C. In this plot, the kRISC values at a specific (0.045 eV) for ΔEST can be seen. Monitoring these values revealed that the DHO and st approximations substantially affected kRISC prediction.
Figure 3B illustrates that the DHO approximation changes Re[ρ(t)] from the 1st+Condon’s oscillatory damped shape to a monotonically decaying curve. Equation 8 can be reduced to the identity , indicating that it corresponds to the area of the region bounded by the curve Re[ρ(t)] in Fig. 3B (which appears as the intercepts of the curves plotted in Fig. 3C). Moreover, the DHO-approximated attains the largest value and is overestimated compared with the 1st+Condon counterpart. This trend reflects the fact that at the DHO level, Re[ρ(t)] is a positive function, while the 1st+Condon curve involves a positive-negative oscillation, which causes large-scale cancellation in its integration. The above analysis confirms that the DHO treatment spuriously increases kRISC.
The Dus-off and DHO approximations to the 1st+Condon equation result in the following DHO-level expression
| (1A) |
| (1B) |
where the definitions of Sj, , and ωj are given in the “Comparison of the 1st+Condon TVCF method and Marcus formula” section. Applying the st approximation, Eq. 1 is reduced to
| (2A) |
| (2B) |
The above analyses show that the st approximation transforms ρ(t) into a Gaussian shape. Its Fourier transform (Eq. 2) retains a Gaussian structure, as confirmed by the resulting st- and Marcus-level kRISC curves (Fig. 3C). These curves exhibit negative quadratic decay on a logarithmic scale []. With small ΔEST, this shape thus plays a role in artificially enlarging kRISC with a slow decay, compared to a monotonic [O(−ΔEST)] decay. Meanwhile, with larger ΔEST, the decay becomes increasingly steep and causes a near-diminishing kRISC. Therefore, the favorable features of the Marcus results for the four MR-TADF systems are fortuitous owing to error cancellation.
1st+Condon TVCF and Marcus rate constants for 121 MR-TADF molecules
Given that the Marcus predictions coincide closely with the experimental values for the above four molecules, checking the applicability of this error cancellation to other MR-TADF molecules would be interesting. Therefore, we extended the test set to 121 MR-TADF molecules (14), on which kRISC calculations were performed at 1st+Condon and Marcus levels. Figure 4 (A to C) shows scatter plots comparing the calculated values and the experimental data. The rate constant equations were evaluated using , determined by the ARPSfit method. was obtained from the RISC activation energy based on the experimental Arrhenius plots or the difference between the measured fluorescence and phosphorescence photon energies (14) (table S9). The use of the spectroscopic fluorescence-phosphorescence difference for is discussed in section S5.
Fig. 4. Rate constant calculations for 121 MR-TADF molecules.
(A to C) Scatter plots to compare the experimental RISC rate constants kRISC and the predicted rate constants using 1st+Condon and Marcus formulas with their parameter ∆EST set to for 121 MR-TADF molecules. The red dashed line indicates the exact agreement y = x. The two gray dashed lines indicate the relations y = 0.1x and y = 10x. (D to F) Plots of kRISC as a function of ΔEST for AZA-BN (large λ), BOBO-Z (small λ and small ΔEST), and BN-CP2 (small λ and large ΔEST). (G) The types of the MR-TADF molecules characteristics depending on ΔEST and λ: types 1, 2, and 3. The magnitude relation between Marcus and 1st+Condon results depends on the type. (H) Scatter plots of ΔEST versus reorganization energy λ for 121 MR-TADF molecules. The data points are classified into types 1, 2, and 3, based on the ratio of kRISC obtained by Marcus and 1st+Condon theories [kRISC(Marcus)/kRISC(1st + Condon)].
Comparing Fig. 4 (B and C) with their comparative plots in fig. S6, we confirmed that the use of ARPSfit-based showed better agreement with the experimental kRISC than for both models. The 1st+Condon model tended to underestimate kRISC relative to the experimental data, with a Pearson correlation coefficient of 0.51. This value favorably exceeded that of Marcus theory versus the experiments (0.33). However, there were several cases in which Marcus theory predicted a larger kRISC than the 1st+Condon, resulting in a better agreement with the experimental measurements. This trend appeared to be related to the error cancellation observed in the previous analysis. However, in several instances, the Marcus results were largely underestimated. Such large statistical variations in the Marcus results could be due to the Gaussian nature of the formula, which is sensitive to ΔEST alongside vibrational frequencies and reorganization energy. The underestimation of the 1st+Condon predictions may be rectified by the 2nd+HT treatment, which requires additional electronic structure computations.
Figure 4 (D to F) shows the plots of the Marcus and 1st+Condon results as a function of ΔEST for AZA-BN, BOBO-Z, and BN-CP2, respectively, as three typical classes. The reorganization energy λ plays a major role in determining the shape of the Marcus curve. In the large-λ case (e.g., AZA-BN) the ΔEST-dependent decay was as slow as that of the 1st+Condon model (Fig. 4D).
The small-λ cases—BOBO-Z and BN-CP2—led the Marcus prediction to largely depend on ΔEST (Fig. 4, E and F). The suited threshold for λ to be considered as large or small can be ~5 × 102 cm−1 (or 0.06 eV). A small ΔEST yielded a large kRISC (BOBO-Z), and vice versa (BN-CP2). A notable feature could be observed in the Marcus results with small λ, ascribed to small structural changes between S1 and T1; for BN-CP2 (type 3), the RISC transition was characterized as forbidden (Fig. 4C).
The above insights led us to classify the tested MR-TADF molecules into three types, depending on the degrees of ΔEST and λ. Figure 4G summarizes these characteristics. Type 1 is associated with large λ, whereas types 2 and 3 are associated with small λ. The ΔEST of types 2 and 3 were characterized as small and large, respectively. Marcus theory tended to overestimate kRISC for types 1 and 2 and considerably underestimated kRISC for type 3 compared to the 1st+Condon estimations. However, the overestimation was prone to error cancellation in the Marcus predictions, resulting in a near coincidence with the experimental kRISC.
Figure 4H shows scatter plots of λ versus ΔEST for 121 data points. On the basis of the ratio of kRISC—Marcus versus 1st+Condon, these points appeared to fall into three categories, types 1 to 3. This classification partitioned the data points approximately evenly. This implied that 30% of the Marcus predictions could be largely underestimated (type 3). Meanwhile, the other 30% might suffer from erroneous overestimation but resulted in values closer to the experimental data (type 1). This observation could also be confirmed in the cumulative histogram used to quantitatively analyze the distribution of the ratios relative to kExp. (figs. S7 and S8).
DISCUSSION
We presented an in-depth analysis of the spin-flipping process of MR-TADF materials accelerated by two types of spin-vibronic interactions: NA-SVC and HT-SVC. The importance of their concerted effects, 2nd+HT, was identified as a factor that reliably characterized the RISC process. Full incorporation of the 2nd+HT terms into the TVCF framework was developed. This extended approach markedly outperformed the previous perturbative models.
The accuracy of the predicted kRISC is also susceptible to the precision of ΔEST. We developed the APRSfit method for effectively determining ΔEST based on the fitting to experimental Arrhenius plots. This method resulted in lower errors in the calculated kRISC, compared with the cases using .
Furthermore, we examined the extent to which each step of the approximation to the 1st+Condon theory, reduced to Marcus theory, affected kRISC for MR-TADF systems. Our analyses indicate that the DHO and st approximations have a notable impact. These degradations cause fortuitous error cancellation in several Marcus results. By calculating kRISC for 121 MR-TADF molecules, we identified the conditions that could closely align the Marcus predictions with the experimental values. One-third of the tested systems did not meet these conditions, and Marcus theory largely underestimated kRISC for them.
Our extended modeling should benefit the reliable characterization of kRISC for MR-TADF molecules despite its limitation by the harmonic approximation. Future work involves applying it to molecules underestimated by Marcus theory for molecular exploration.
MATERIALS AND METHODS
Computational details
The structural parameters of all the considered molecules were determined by performing TDDFT-level geometry optimizations for the S1 and T1 states separately. The TDDFT-level calculations, including the determination of vibrational states, were performed at the PBE0-D3BJ/def2-SV(P) level using Gaussian 16 (43). This level was compatible with those used in the precedent studies [e.g., (16–18)]. The NAC matrix elements were calculated using Q-Chem version 5.4.2 (44). The SOC matrix elements of the singlet and triplet states were evaluated at the ZORA-PBE0/ZORA-def2-SV(P) level (45) using ORCA version 5.0.2 (46, 47). For these SOC calculations, we used the SARC/J basis (48) for density fitting to the Coulomb term and the chain-of-sphere exchange (RIJCOSX) (49, 50). The derivatives of the SOC matrix elements were calculated via numerical differentiation at the T1 structure. The numerical differentiation is computationally inefficient but cannot be avoided due to the absence of analytic derivative implementation. These costs limited the application of the 2nd+HT method to a very large number of molecules.
All rate constants were calculated using an in-house code. The vibrational modes calculated at the initial and final states were used for the TVCF rate constant calculations, considering the Duschinsky rotation [for details, see (51)]. The correlation function was evaluated over a time range of 0.05 to 10000.15 fs with a time step of 0.1 fs. The correlation function for the negative time was obtained from the complex conjugate of the positive time region. In the second TVCF treatment, the T2, T3, and T4 states were incorporated as intermediate states. The SCS-CC2 calculations were performed using Turbomole version 7.6 (52, 53), and their LT-corrected energies were evaluated using the formula ΔEST(SCS − CC2 LT) = 0.94 ΔEST(SCS − CC2) − 0.01 (34).
Formulation of 2nd+HT TVCF theory
Let us here outline the formulation of the 2nd+HT TVCF rate constant equation. Section S7 shows full details of the derivation. The 2nd+HT theory is built upon the second-order Fermi’s golden rule given by
| (3) |
which provides the transition rate between the initial (i) and final (f) states (42). It is written using the vibronic state energies Eiv and Efu, the Boltzmann distribution with Zi = ∑v e−βEiv. We modeled the spin-flipping process using the perturbation consisting of the SOC term and the NAC term with . The rate constants for the ISC and RISC transitions between S1 and T1 were evaluated as kISC = ∑m=0,±1 kT1,m←S1 and , respectively. As discussed in section S7, Eq. 3 can be partitioned into the first- and second-order components as , where , which accounts for a direct transition process, and the rest, , is associated with an indirect process mediated by the intermediate states.
The SOC term was considered up to the HT level (20), which describes via the first-order Taylor expansion with respect to the vibrational modes {Qfk} as , where the zeroth- and first-order terms are denoted Condon and HT, respectively. This form leads us to decompose and further into the Condon and HT elements as and , respectively. By combining these elements, four levels of the rate-constant equation—the 1st+Condon, 1st+HT, 2nd+Condon, and 2nd+HT—are defined as follows
| (4) |
| (5) |
| (6) |
| (7) |
These equations were evaluated within the framework of the TVCF method (21, 24). In TVCF, the rate equation is expressed as a time-domain propagation of the normal-mode correlation function , as follows
| (8) |
where X refers to the level of the formula: 1st+Condon, 1st+HT, 2nd+Condon, and 2nd+HT. Equation 8 is formed as a function of ΔEST via the following identity
| (9) |
which is based on the relations , , and ΔEST = Ei − Ef, where Ei and Ef are the adiabatic electronic state energies, and and correspond to the vibrational state energies. Note that ΔEST does not affect .
A major approximation in TVCF considers the molecular vibrations of each electronic state to be harmonic oscillations. This enables us to obtain an analytical formula for that can be evaluated in a computationally feasible manner. The resulting can be divided into X-independent and X-dependent factors as follows
| (10) |
The is given as
| (11) |
which is built upon various types of matrices defined as
| (12) |
| (13) |
| (14) |
| (15) |
| (16) |
The a and b are diagonal matrices of the dimension Nvib with their elements and based on the initial state frequency ωj. Their variants denoted and , respectively, are based on the final state frequency in the place of ωj. J is the Duschinsky rotation matrix (54, 55), and K is the displacement vector between the initial and final states.
The integral in Eq. 8 can be evaluated using the Fourier transform of from the time (t) domain to the energy (ΔEST) domain, where (Eq. 10) is evaluated at grid points of t ∈ [−tmax, tmax]. The resulting kf←i is obtained as a function of ΔEST, which is represented on a grid in energy coordinates. The objective kf←i can be determined by substituting ΔEST with the user-input value of the adiabatic energy difference, ΔEfi.
Comparison of the 1st+Condon TVCF method and Marcus formula
In previous studies, the semiclassical Marcus formula (15) was mostly used to calculate the RISC rate constants, as follows
| (17) |
where λRISC is the reorganization energy and is defined as
| (18) |
The Marcus equation can be derived from the 1st+Condon equation k1st+Condon (Eq. 4), via stepwise approximations (42, 56).
Equation 4 is rewritten as
| (19A) |
| (19B) |
The initial step to approximate toward Marcus theory is to cut off the Duschinsky rotation by setting J to an identity matrix (Dus-off). This enables ρ(t) to be factorized into separate entities associated with each vibrational mode, as follows
| (20) |
Furthermore, we consider the DHO model, which uses the common frequency for the initial-state ωj and the final-state , resulting in
| (21) |
where and .
The st approximation, , subject to the assumption ∑j Sj ≫ 1, is applied to Eq. 21, which results in
| (22) |
Finally, the Marcus equation (Eq. 17) can be derived using Eq. 22 by applying the ht approximation that the temperature T is sufficiently large and ℏωj/kT ≪ 1 holds
| (23) |
Acknowledgments
We would like to express our gratitude to S. Yamaguchi for unwavering support and encouragement. We thank Editage for English language editing.
Funding: This study was supported by JSPS KAKENHI (grant nos. 21H01881, JPJSBP120229601, and JPJSBP120237705 to T.Yan.; 20 K05430 to K.J.F.; and 21H04694 to T.Yas.). T.Yas. and K.J.F. thank JST CREST (grant no. JPMJCR21O5). T.Yan. and K.J.F. are grateful for the support from JSPS Fund for the Promotion of Joint International Research (grant no. 22 K21346) and MEXT Promotion of Development of a Joint Usage/Research System Project: Coalition of Universities for Research Excellence Program (CURE) (grant no. JPMXP1323015482) and for grant CPU time from Research Center for Computational Science, Okazaki, Japan (23-IMS-C087).
Author contributions: Conceptualization: M.H., T.Yas., K.J.F.., and T.Yan. Methodology: M.H. and N.I. Investigation: M.H., K.J.F., and T.Yan. Visualization: M.H. Validation: All authors. Supervision: T.Yan. Writing—original draft: M.H. and T.Yan. Writing—review and editing: All authors.
Competing interests: The authors declare that they have no competing interests.
Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials.
Supplementary Materials
This PDF file includes:
Sections S1 to S7
Figs. S1 to S8
Tables S1 to S10
Legend for Supplemental auxiliary files
Other Supplementary Material for this manuscript includes the following:
Supplemental auxiliary files
REFERENCES AND NOTES
- 1.Tang C. W., VanSlyke S. A., Organic electroluminescent diodes. Appl. Phys. Lett. 51, 913–915 (1987). [Google Scholar]
- 2.Burroughes J. H., Bradley D. D. C., Brown A. R., Marks R. N., Mackay K., Friend R. H., Burns P. L., Holmes A. B., Light-emitting diodes based on conjugated polymers. Nature 347, 539–541 (1990). [Google Scholar]
- 3.Friend R. H., Gymer R. W., Holmes A. B., Burroughes J. H., Marks R. N., Taliani C., Bradley D. D. C., Santos D. A. D., Brédas J. L., Lögdlund M., Salaneck W. R., Electroluminescence in conjugated polymers. Nature 397, 121–128 (1999). [Google Scholar]
- 4.Baldo M. A., O'Brien D. F., You Y., Shoustikov A., Sibley S., Thompson M. E., Forrest S. R., Highly efficient phosphorescent emission from organic electroluminescent devices. Nature 395, 151–154 (1998). [Google Scholar]
- 5.Adachi C., Baldo M. A., Thompson M. E., Forrest S. R., Nearly 100% internal phosphorescence efficiency in an organic light-emitting device. J. Appl. Phys. 90, 5048–5051 (2001). [Google Scholar]
- 6.Xiao L., Chen Z., Qu B., Luo J., Kong S., Gong Q., Kido J., Recent progresses on materials for electrophosphorescent organic light-emitting devices. Adv. Mater. 23, 926–952 (2011). [DOI] [PubMed] [Google Scholar]
- 7.Uoyama H., Goushi K., Shizu K., Nomura H., Adachi C., Highly efficient organic light-emitting diodes from delayed fluorescence. Nature 492, 234–238 (2012). [DOI] [PubMed] [Google Scholar]
- 8.Tao Y., Yuan K., Chen T., Xu P., Li H., Chen R., Zheng C., Zhang L., Huang W., Thermally activated delayed fluorescence materials towards the breakthrough of organoelectronics. Adv. Mater. 26, 7931–7958 (2014). [DOI] [PubMed] [Google Scholar]
- 9.Wong M. Y., Zysman-Colman E., Purely organic thermally activated delayed fluorescence materials for organic light-emitting diodes. Adv. Mater. 29, 1605444 (2017). [DOI] [PubMed] [Google Scholar]
- 10.Nakanotani H., Tsuchiya Y., Adachi C., Thermally-activated delayed fluorescence for light-emitting devices. Chem. Lett. 50, 938–948 (2021). [Google Scholar]
- 11.Hatakeyama T., Shiren K., Nakajima K., Nomura S., Nakatsuka S., Kinoshita K., Ni J., Ono Y., Ikuta T., Ultrapure blue thermally activated delayed fluorescence molecules: Efficient HOMO-LUMO separation by the multiple resonance effect. Adv. Mater. 28, 2777–2781 (2016). [DOI] [PubMed] [Google Scholar]
- 12.Kothavale S. S., Lee J. Y., Three- and four-coordinate, boron-based, thermally activated delayed fluorescent emitters. Adv. Opt. Mater. 8, 2000922 (2020). [Google Scholar]
- 13.Madayanad Suresh S., Hall D., Beljonne D., Olivier Y., Zysman-Colman E., Multiresonant thermally activated delayed fluorescence emitters based on heteroatom-doped nanographenes: Recent advances and prospects for organic light-emitting diodes. Adv. Funct. Mater. 30, 1908677 (2020). [Google Scholar]
- 14.Kim H. J., Yasuda T., Narrowband emissive thermally activated delayed fluorescence materials. Adv. Opt. Mater. 10, 2201714 (2022). [Google Scholar]
- 15.Marcus R. A., Electron transfer reactions in chemistry. theory and experiment. Rev. Mod. Phys. 65, 599–610 (1993). [Google Scholar]
- 16.Samanta P. K., Kim D., Coropceanu V., Brédas 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. 139, 4042–4051 (2017). [DOI] [PubMed] [Google Scholar]
- 17.Olivier Y., Yurash B., Muccioli L., D’Avino G., Mikhnenko O., Sancho-García J. C., Adachi C., Nguyen T. Q., Beljonne D., Nature of the singlet and triplet excitations mediating thermally activated delayed fluorescence. Phys. Rev. Mater. 1, 075602 (2017). [Google Scholar]
- 18.Olivier Y., Sancho-Garcia J.-C., Muccioli L., D’Avino G., Beljonne D., Computational design of thermally activated delayed fluorescence materials: The challenges ahead. J. Phys. Chem. Lett. 9, 6149–6163 (2018). [DOI] [PubMed] [Google Scholar]
- 19.Aizawa N., Harabuchi Y., Maeda S., Pu Y.-J., Kinetic prediction of reverse intersystem crossing in organic donor–acceptor molecules. Nat. Commun. 11, 3909 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Penfold T. J., Gindensperger E., Daniel C., Marian C. M., Spin-vibronic mechanism for intersystem crossing. Chem. Rev. 118, 6975–7025 (2018). [DOI] [PubMed] [Google Scholar]
- 21.Peng Q., Niu Y., Shi Q., Gao X., Shuai Z., Correlation function formalism for triplet excited state decay: Combined spin-orbit and nonadiabatic couplings. J. Chem. Theory Comput. 9, 1132–1143 (2013). [DOI] [PubMed] [Google Scholar]
- 22.Northey T., Penfold T. J., The intersystem crossing mechanism of an ultrapure blue organoboron emitter. Org. Electron. 59, 45–48 (2018). [Google Scholar]
- 23.Chen X.-K., Zhang S.-F., Fan J.-X., Ren A.-M., Nature of highly efficient thermally activated delayed fluorescence in organic light-emitting diode emitters: Nonadiabatic effect between excited states. J. Phys. Chem. C 119, 9728–9733 (2015). [Google Scholar]
- 24.Peng Q., Fan D., Duan R., Yi Y., Niu Y., Wang D., Shuai Z., Theoretical study of conversion and decay processes of excited triplet and singlet states in a thermally activated delayed fluorescence molecule. J. Phys. Chem. C 121, 13448–13456 (2017). [Google Scholar]
- 25.Gibson J., Monkman A. P., Penfold T. J., The importance of vibronic coupling for efficient reverse intersystem crossing in thermally activated delayed fluorescence molecules. ChemPhysChem 17, 2956–2961 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Marian C. M., Spin-orbit coupling and intersystem crossing in molecules. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2, 187–203 (2012). [Google Scholar]
- 27.Tatchen J., Gilka N., Marian C. M., Intersystem crossing driven by vibronic spin-orbit coupling: A case study on psoralen. Phys. Chem. Chem. Phys. 9, 5209–5221 (2007). [DOI] [PubMed] [Google Scholar]
- 28.Etinski M., Tatchen J., Marian C. M., Time-dependent approaches for the calculation of intersystem crossing rates. J. Chem. Phys. 134, 154105 (2011). [DOI] [PubMed] [Google Scholar]
- 29.Lin S., Pei Z., Zhang B., Ma H., Liang W., Vibronic coupling effect on the vibrationally resolved electronic spectra and intersystem crossing rates of a TADF emitter: 7-PhQAD. J. Phys. Chem. A 126, 239–248 (2022). [DOI] [PubMed] [Google Scholar]
- 30.Rodriguez-Serrano A., Dinkelbach F., Marian C. M., Intersystem crossing processes in the 2CzPN emitter: A DFT/MRCI study including vibrational spin-orbit interactions. Phys. Chem. Chem. Phys. 23, 3668–3678 (2021). [DOI] [PubMed] [Google Scholar]
- 31.Park I. S., Matsuo K., Aizawa N., Yasuda T., High-performance dibenzoheteraborin-based thermally activated delayed fluorescence emitters: Molecular architectonics for concurrently achieving narrowband emission and efficient triplet-singlet spin conversion. Adv. Funct. Mater. 28, 1802031 (2018). [Google Scholar]
- 32.Sun H., Zhong C., Brédas J.-L., Reliable prediction with tuned range-separated functionals of the singlet-triplet gap in organic emitters for thermally activated delayed fluorescence. J. Chem. Theory Comput. 11, 3851–3858 (2015). [DOI] [PubMed] [Google Scholar]
- 33.Kondo M., Singlet-triplet energy gap of multiresonant molecular systems: A double hybrid time-dependent density functional theory study. Chem. Phys. Lett. 804, 139895 (2022). [Google Scholar]
- 34.Hall D., Sancho-García J. C., Pershin A., Ricci G., Beljonne D., Zysman-Colman E., Olivier Y., Modeling of multiresonant thermally activated delayed fluorescence emitters-properly accounting for electron correlation is key! J. Chem. Theory Comput. 18, 4903–4918 (2022). [DOI] [PubMed] [Google Scholar]
- 35.Niu Y., Peng Q., Deng C., Gao X., Shuai Z., Theory of excited state decays and optical spectra: Application to polyatomic molecules. J. Phys. Chem. A 114, 7817–7831 (2010). [DOI] [PubMed] [Google Scholar]
- 36.Kondo Y., Yoshiura K., Kitera S., Nishi H., Oda S., Gotoh H., Sasada Y., Yanai M., Hatakeyama T., Narrowband deep-blue organic light-emitting diode featuring an organoboron-based emitter. Nat. Photonics 13, 678–682 (2019). [Google Scholar]
- 37.Park I. S., Yang M., Shibata H., Amanokura N., Yasuda T., Achieving ultimate narrowband and ultrapure blue organic light-emitting diodes based on polycyclo-heteraborin multi-resonance delayed-fluorescence emitters. Adv. Mater. 34, e2107951 (2022). [DOI] [PubMed] [Google Scholar]
- 38.Adamo C., Barone V., Toward reliable density functional methods without adjustable parameters: The pbe0 model. J. Chem. Phys. 110, 6158–6170 (1999). [Google Scholar]
- 39.Weigend F., Ahlrichs R., Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy. Phys. Chem. Chem. Phys. 7, 3297–3305 (2005). [DOI] [PubMed] [Google Scholar]
- 40.Grimme S., Semiempirical hybrid density functional with perturbative second-order correlation. J. Chem. Phys. 124, 034108 (2006). [DOI] [PubMed] [Google Scholar]
- 41.Hellweg A., Grün S. A., Hättig C., Benchmarking the performance of spin-component scaled CC2 in ground and electronically excited states. Phys. Chem. Chem. Phys. 10, 4119–4127 (2008). [DOI] [PubMed] [Google Scholar]
- 42.S. H. Lin, Advances in Chemical Physics (John Wiley & Sons Ltd., 2002), pp. 1–88. [Google Scholar]
- 43.M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery, Jr., J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman, D. J. Fox, Gaussian 16, revision B.01 (Gaussian Inc., 2016). [Google Scholar]
- 44.Epifanovsky E., Gilbert A. T. B., Feng X., Lee J., Mao Y., Mardirossian N., Pokhilko P., White A. F., Coons M. P., Dempwolff A. L., Gan Z., Hait D., Horn P. R., Jacobson L. D., Kaliman I., Kussmann J., Lange A. W., Lao K. U., Levine D. S., Liu J., McKenzie S. C., Morrison A. F., Nanda K. D., Plasser F., Rehn D. R., Vidal M. L., You Z. Q., Zhu Y., Alam B., Albrecht B. J., Aldossary A., Alguire E., Andersen J. H., Athavale V., Barton D., Begam K., Behn A., Bellonzi N., Bernard Y. A., Berquist E. J., Burton H. G. A., Carreras A., Carter-Fenk K., Chakraborty R., Chien A. D., Closser K. D., Cofer-Shabica V., Dasgupta S., de Wergifosse M., Deng J., Diedenhofen M., do H., Ehlert S., Fang P. T., Fatehi S., Feng Q., Friedhoff T., Gayvert J., Ge Q., Gidofalvi G., Goldey M., Gomes J., González-Espinoza C. E., Gulania S., Gunina A. O., Hanson-Heine M. W. D., Harbach P. H. P., Hauser A., Herbst M. F., Hernández Vera M., Hodecker M., Holden Z. C., Houck S., Huang X., Hui K., Huynh B. C., Ivanov M., Jász Á., Ji H., Jiang H., Kaduk B., Kähler S., Khistyaev K., Kim J., Kis G., Klunzinger P., Koczor-Benda Z., Koh J. H., Kosenkov D., Koulias L., Kowalczyk T., Krauter C. M., Kue K., Kunitsa A., Kus T., Ladjánszki I., Landau A., Lawler K. V., Lefrancois D., Lehtola S., Li R. R., Li Y. P., Liang J., Liebenthal M., Lin H. H., Lin Y. S., Liu F., Liu K. Y., Loipersberger M., Luenser A., Manjanath A., Manohar P., Mansoor E., Manzer S. F., Mao S. P., Marenich A. V., Markovich T., Mason S., Maurer S. A., McLaughlin P. F., Menger M. F. S. J., Mewes J. M., Mewes S. A., Morgante P., Mullinax J. W., Oosterbaan K. J., Paran G., Paul A. C., Paul S. K., Pavošević F., Pei Z., Prager S., Proynov E. I., Rák Á., Ramos-Cordoba E., Rana B., Rask A. E., Rettig A., Richard R. M., Rob F., Rossomme E., Scheele T., Scheurer M., Schneider M., Sergueev N., Sharada S. M., Skomorowski W., Small D. W., Stein C. J., Su Y. C., Sundstrom E. J., Tao Z., Thirman J., Tornai G. J., Tsuchimochi T., Tubman N. M., Veccham S. P., Vydrov O., Wenzel J., Witte J., Yamada A., Yao K., Yeganeh S., Yost S. R., Zech A., Zhang I. Y., Zhang X., Zhang Y., Zuev D., Aspuru-Guzik A., Bell A. T., Besley N. A., Bravaya K. B., Brooks B. R., Casanova D., Chai J. D., Coriani S., Cramer C. J., Cserey G., DePrince A. E. III, DiStasio R. A. Jr., Dreuw A., Dunietz B. D., Furlani T. R., Goddard W. A. III, Hammes-Schiffer S., Head-Gordon T., Hehre W. J., Hsu C. P., Jagau T. C., Jung Y., Klamt A., Kong J., Lambrecht D. S., Liang W. Z., Mayhall N. J., McCurdy C. W., Neaton J. B., Ochsenfeld C., Parkhill J. A., Peverati R., Rassolov V. A., Shao Y., Slipchenko L. V., Stauch T., Steele R. P., Subotnik J. E., Thom A. J. W., Tkatchenko A., Truhlar D. G., van Voorhis T., Wesolowski T. A., Whaley K. B., Woodcock H. L. III, Zimmerman P. M., Faraji S., Gill P. M. W., Head-Gordon M., Herbert J. M., Krylov A. I., Software for the frontiers of quantum chemistry: An overview of developments in the Q-Chem 5 package. J. Chem. Phys. 155, 084801 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.van Wüllen C., Molecular density functional calculations in the regular relativistic approximation: Method, application to coinage metal diatomics, hydrides, fluorides and chlorides, and comparison with first-order relativistic calculations. J. Chem. Phys. 109, 392–399 (1998). [Google Scholar]
- 46.Neese F., The ORCA program system. WIREs Comput. Mol. Sci. 2, 73–78 (2012). [Google Scholar]
- 47.Neese F., Software update: TheORCAprogram system—Version 5.0. Mol. Sci. 12, e1606 (2022). [Google Scholar]
- 48.Weigend F., Accurate coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 8, 1057–1065 (2006). [DOI] [PubMed] [Google Scholar]
- 49.Neese F., Wennmohs F., Hansen A., Becker U., Efficient, approximate and parallel Hartree–Fock and hybrid DFT calculations. A ‘chain-of-spheres’ algorithm for the Hartree–Fock exchange. Chem. Phys. 356, 98–109 (2009). [Google Scholar]
- 50.Helmich-Paris B., de Souza B., Neese F., Izsák R., An improved chain of spheres for exchange algorithm. J. Chem. Phys. 155, 104109 (2021). [DOI] [PubMed] [Google Scholar]
- 51.Inai N., Yamaguchi S., Yanai T., Theoretical insight into the effect of phosphorus oxygenation on nonradiative decays: Comparative analysis of p-bridged stilbene analogs. ACS Phys. Chem Au 3, 540–552 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.TURBOMOLE V7.6 2021, a development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 1989-2007, TURBOMOLE GmbH, since 2007; www.turbomole.org.
- 53.Balasubramani S. G., Chen G. P., Coriani S., Diedenhofen M., Frank M. S., Franzke Y. J., Furche F., Grotjahn R., Harding M. E., Hättig C., Hellweg A., Helmich-Paris B., Holzer C., Huniar U., Kaupp M., Marefat Khah A., Karbalaei Khani S., Müller T., Mack F., Nguyen B. D., Parker S. M., Perlt E., Rappoport D., Reiter K., Roy S., Rückert M., Schmitz G., Sierka M., Tapavicza E., Tew D. P., van Wüllen C., Voora V. K., Weigend F., Wodyński A., Yu J. M., TURBOMOLE: Modular program suite for ab initio quantum-chemical and condensed-matter simulations. J. Chem. Phys. 152, 184107 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Dushinsky F., The importance of the electron spectrum in multi atomic molecules concerning the franck- condon principle. Acta Physicochim. URSS 7, 551–566 (1937). [Google Scholar]
- 55.Avila Ferrer F. J., Barone V., Cappelli C., Santoro F., Duschinsky, Herzberg-Teller, and multiple electronic resonance interferential effects in resonance raman spectra and excitation profiles. the case of pyrene. J. Chem. Theory Comput. 9, 3597–3611 (2013). [DOI] [PubMed] [Google Scholar]
- 56.Lee M. H., Dunietz B. D., Geva E., Calculation from first principles of intramolecular golden-rule rate constants for photo-induced electron transfer in molecular donor-acceptor systems. J. Phys. Chem. C 117, 23391–23401 (2013). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Sections S1 to S7
Figs. S1 to S8
Tables S1 to S10
Legend for Supplemental auxiliary files
Supplemental auxiliary files




