Neethu
Anand
a,
Munnyon
Kim
a,
Changmin
Lee
b,
Jinhyuk
Ma
a and
Taiha
Joo
*a
aDepartment of Chemistry, Pohang University of Science and Technology, Pohang, 37673, South Korea. E-mail: thjoo@postech.ac.kr
bDepartment of Chemistry, Incheon National University, Incheon 22012, South Korea
First published on 30th September 2025
Internal conversion (IC) dynamics from higher-lying to lower electronic states following photoexcitation are often described as ultrafast, occurring in much less than 1 ps. However, IC processes can exhibit a range of time scales, depending critically on the energetic landscape of excited-state manifolds and the strength of vibronic couplings that drive nonadiabatic transitions. These dynamics play a fundamental role in most photochemical and photophysical applications. A previous time-resolved fluorescence study revealed that two structurally analogous boron-dipyrromethene (BODIPY) molecules, PM650 and PM597, exhibit markedly different IC rates, despite both undergoing ultrafast IC in <100 fs from the S3/S2 to the S1 state. Nuclear wave packets persisting in the S1 state after the IC were also observed by time-resolved fluorescence. To elucidate the origin of these divergent IC rates and the nature of vibronic interactions among excited states, we performed theoretical simulations using the multiconfiguration time-dependent Hartree (MCTDH) method. Our results reveal nonadiabatic decay pathways mediated by vibronically coupled S1, S2, and S3 potential energy surfaces, with multiple conical intersections (CIs) enabling the IC processes. The IC rates obtained from the MCTDH simulations are in good agreement with the experimental observations, including the contrasting rates for PM650 and PM597. Importantly, the proximity of CIs to the Franck–Condon region was found to significantly influence IC efficiency. As more vibrational modes were incorporated into the model, a consistent acceleration of the IC dynamics was observed, underscoring the role of multimode effects in nonadiabatic transitions through CIs. Additionally, coherent vibrational spectra of the S1 state, generated from nuclear densities computed via MCTDH following excitation to higher states, were found to match experimental results closely, further supporting the conclusions of this study. Overall, these findings advance our understanding of the intricate excited-state dynamics and highlight the critical role of vibronic coupling and CIs in ultrafast IC.
Although IC from Sn (n ≥ 2) to S1 is generally considered ultrafast, the actual rates can vary significantly, ranging from a few femtoseconds to several picoseconds, depending on the specific molecular electronic couplings.3,5–16 This process is facilitated by conical intersections (CIs), which provide highly efficient, non-radiative pathways for electronic relaxation.17,18 It is important to note that nonadiabatic processes via CIs in polyatomic molecules are generally not simple one-dimensional phenomena involving a single reaction coordinate. Rather, they require the inclusion of multiple vibrational modes to accurately model the vibronic interactions.19–22 Thus, a detailed understanding of the molecular mechanisms underlying ultrafast IC is crucial for both fundamental photophysics and applications involving light–matter interaction.
The IC process is nonadiabatic and inherently involves vibronic interactions, where electronic and vibrational degrees of freedom are strongly coupled. Upon impulsive photoexcitation to an Sn (n ≥ 2) state, coherent nuclear wave packets (NWPs) are formed in the Franck–Condon (FC) region. These NWPs may survive the IC process and persist in the S1 state.10,23–27 During the passage through CIs, the amplitudes and phases of the NWPs can be modified. Vibrational modes that are strongly coupled to the IC process may exhibit particularly significant modulation. Therefore, measuring NWPs in excited states can yield valuable information about the reaction coordinates of IC and the vibronic couplings between excited electronic states.
Previously, by employing time-resolved fluorescence (TF) with a time resolution high enough to resolve the motion of NWPs, we reported IC rates and the formation and evolution of NWPs in the S1 state following photoexcitation to the S2 and S3 states for two Boron-dipyrromethene (BODIPY) molecules, PM597 and PM650.12 The molecular structures are shown in Fig. 1. Despite the structural similarity of the two molecules, the S2 → S1 IC rates were vastly different; in PM650 the IC rate is nearly instantaneous, ≫(20 fs)−1, whereas in PM597 it occurs at (51 fs)−1. These two molecules, which share similar molecular structures and electronic characters but exhibit contrasting IC dynamics, provide a valuable opportunity to investigate how subtle changes in molecular and electronic structures and delicate vibronic interactions influence nonadiabatic energy relaxation pathways and dynamics. BODIPY is also an excellent model system owing to its favourable spectroscopic properties, including a large extinction coefficient, high fluorescence quantum yield, narrow absorption and emission linewidths in solution, structural tunability, low sensitivity to solvent polarity, and resistance to aggregation.
![]() | ||
| Fig. 1 Molecular structures of (a) PM650 and (b) PM597. The bond lengths and bond angles are indicated. | ||
NWPs in the S1 state following ICs have also been experimentally observed and analysed.12 In that study, we estimated the displacement between the FC region and the S1 state, and compared the calculated values with experimental NWPs. However, our analysis was based on the assumption of an impulsive IC process, in which the transition from S2 to S1 was treated as instantaneous.28 This approach was partially succeeded in reproducing the experimental observations. A more rigorous approach to calculating NWPs and coherent vibrational spectra (CVS) following photoexcitation and subsequent molecular processes was recently introduced using Born–Oppenheimer molecular dynamics (BOMD) simulations.29 In this method, nuclear dynamics are obtained by projecting the BOMD trajectories onto the normal modes of the product state. However, this molecular-dynamics-based method, despite incorporating on-the-fly quantum calculations of structure and forces, still treats nuclear motions classically. Therefore, a more rigorous quantum mechanical treatment of nuclear motion is essential for a complete understanding of quantum dynamics, particularly for processes involving CIs such as ultrafast IC.
In this study, we adopt a quantum mechanical approach to describe ultrafast IC, explicitly accounting for the vibronic coupling effects. We employed the multi-configuration time-dependent Hartree (MCTDH) method to perform quantum wave packet simulations, enabling tracing of IC pathways and calculation of the NWPs that persist in the S1 state after IC.30–35 This quantum nuclear wave packet simulation method not only allows for accurate determination of the IC rates and surviving NWPs, but also provides deeper insight into the role of vibrational modes that are strongly coupled to the IC processes. By capturing the intricate interplay between electronic and nuclear dynamics, our study advances the theoretical framework for understanding ultrafast IC and establishes a foundation for future experimental and computational investigations in photophysics and photochemistry.
| H = H0(Q)I3 + W(1), | (1) |
![]() | (2) |
![]() | (3) |
![]() | (4) |
![]() | (5) |
![]() | (6) |
The stationary points on the PESs and the positions of minimum energy conical intersections (MECIs) are located within the LVC model using standard expressions reported in the literature. The energy and position of MECI along the coupling modes is given by:19
![]() | (7) |
![]() | (8) |
![]() | (9) |
![]() | (10) |
is the SPF for the kth DOF. The SPFs themselves are expanded in terms of time-independent primitive basis functions:![]() | (11) |
For the MCTDH simulations, a subset of 33 vibrational modes was selected out of 114 and 174 modes for PM650 and PM597, respectively, based on their vibronic coupling strength (κ2/2ω2). The initial ground-state vibrational wave packet was vertically promoted to the S2 or S3 electronic state and propagated for 1 ps. To analyse the wave packet dynamics, nuclear density distributions and diabatic electronic state populations were evaluated. Based on the nuclear density projected onto the S1 state following excitation to S2/S3, CVS were computed.
Wave packet simulations were performed using the Heidelberg MCTDH package (version 8.4, Revision 8).38 All simulation parameters were obtained from quantum chemical calculations in the gas phase using the Gaussian 16 software package.39 Ground-state geometry optimizations and frequency calculations for PM650 and PM597 were conducted under Cs symmetry using density functional theory (DFT) with the CAM-B3LYP functional and the 6-311++G(d,p) basis set. Excited-state properties, including vertical excitation energies and oscillator strengths of low-lying electronic states, were computed using time-dependent DFT (TDDFT) with the same functional and basis set. Transition electron density analysis was carried out using Multiwfn software, version 3.8.40,41 LVC parameters and details of the normal modes for the MCTDH simulations are listed in the SI Tables S2–S11.
| States | PM650 (eV) | PM597 (eV) |
|---|---|---|
| S1 (A′′) | 2.65 (0.54) | 2.89 (0.70) |
| S2 (A′′) | 3.41 (0.08) | 3.81 (0.09) |
| S3 (A′) | 3.64 (0.04) | 4.05 (0.05) |
| S4 (A′) | 4.99 (0.26) | 4.93 (0.00) |
The transition density isosurfaces corresponding to these excitations are shown in the SI (Fig. S1). For both dyes, the S0 → S1 transition is strongly delocalized over the BODIPY π-system, indicating a bright π → π* excitation with a large oscillator strength. In contrast, the S0 → S2 transition exhibits more fragmented and localized lobes, suggesting mixed orbital contributions consistent with a less allowed excitation. The S0 → S3 transition shows a similar degree of density localization and is less delocalized than S1. Although both dyes display qualitatively similar features in their transition density isosurfaces, the transition density of PM597 is more strongly aligned and appears more coherent, thereby strengthening the transition dipole and yielding the larger oscillator strength values reported. Because S2 and S3 are closely spaced in energy and have comparable oscillator strengths, both states are expected to be populated upon photoexcitation. Consequently, three excited states, S1 to S3 were included in the MCTDH simulations.
Curve crossings between the S2 and S1 states in PM650 are observed at (−4.68, 4.67 eV) and (5.18, 5.00 eV) along the Q67 and Q75 vibrational modes, respectively. Additionally, an S3/S2 crossing occurs at (−5.90, 6.02 eV) along the Q95 mode. In PM597, S2/S1 curve crossings are located at (6.20, 6.27 eV) and (8.75, 9.68 eV) along the Q96 and Q110 modes, respectively, with an S3/S2 crossing at (−6.80, 8.07 eV) also along Q96. These curve crossings give rise to the well-known multimode effect, which leads to the formation of multidimensional CIs.
A more detailed analysis of CIs was conducted in the multimode context using the LVC potentials. This includes identification of electronic state minima and the locations of MECIs in the space of promoting modes, as shown in Fig. 5. Corresponding values are summarized in Table 2. In PM650, the FC point on the S2 state lies at 3.41 eV with a stabilization energy of 0.44 eV, and the S2/S1 MECI is located at 2.97 eV. The energy difference between the S2 minimum (S2,min) and the MECI is practically zero, indicating that relaxation along the S2 surface inherently involves passing through the MECI. For PM597, the energy difference between the S2,min and the MECI is 0.05 eV (400 cm−1), suggesting that the system should overcome a modest barrier to reach the CI.
![]() | ||
| Fig. 5 Schematic of the energetics of FC point on S2, PES minima and MECI. Location of S2/S1 MECI marked with the dashed line. | ||
| S2,FC | S3,FC | S2/S1 MECI | S3/S2 MECI | S1,min | S2,min | S3,min | |
|---|---|---|---|---|---|---|---|
| PM650 | 3.41 | 3.64 | 2.97 | 3.22 | 2.57 | 2.97 | 3.22 |
| PM597 | 3.81 | 4.05 | 3.47 | 4.43 | 2.82 | 3.42 | 3.71 |
The geometries of the MECIs can be approximated by their positions along the promoting modes; Q67 and Q75 in PM650, and Q96 in PM597. For PM650, the coordinate of the S2/S1 MECI along Q67 (Q21CI,67) is −0.57 and Q21CI,75 = 0.84. In contrast, for PM597, Q21CI,96 = 0.91, indicating that a larger distortion from the FC geometry is required to access the MECI. Thus, the MECI for PM650 is located closer to the FC point, facilitating more efficient IC, whereas in PM597, the MECI lies farther from the FC region, possibly slowing down the IC process.
The geometric parameter QmnCI,i provides insights into the degree of distortion along the normal modes from the FC point to the MECI. To evaluate the associated nuclear configurations, distortions corresponding to the QmnCI,i values were applied to the vibrational modes using the same procedure employed in determining the vibronic coupling constants. For the vibrational modes Q67 and Q75 in PM650 and Q96 in PM597, we constructed a graph in Fig. 6 that shows the percentage difference of relevant structural parameters. The specific bond lengths and angles used in this analysis are indicated in Fig. 1. In PM650, Q21CI,75 shows substantial deviations (>1%) in r2, r8, and a5 with notable changes also observed in r6 and a9. For Q21CI,67 significant deviations are found in r3 and a4. For PM597 Q21CI,96 exhibits prominent structural changes in r2, r3, and a4.
The diabatic electronic populations of the S1–S3 states following excitation to S2 or S3 in PM650 are shown in Fig. 7. Upon excitation to S2, PM650 exhibits ultrafast population transfer to the S1 state in less than 10 fs, accompanied by a minor population in S3. By ∼10 fs, over half of the initial S2 population is redistributed to S1 and S3. Exponential fit of the population of the S1 state gives a rise time constant of 9.6 fs (Fig. S2). Experimental and calculated time constants for the S2 → S1 ICs in PM650 and PM597 upon excitation to S2 are summarized in Table 3. Since the MCTDH method does not incorporate dissipative mechanisms, a dynamic equilibrium is established between S2 and S1 populations, leading to a persistent residual population in S2. In actual systems, however, this remaining S2 population is expected to decay rapidly to S1via dissipation. The proximity of the CI to the FC point and the near-degeneracy between the S2,min and the S2/S1 MECI provide favourable conditions for such ultrafast wave packet transfer well within a single vibrational period along the relevant normal mode. Fig. 7(b) displays the population dynamics following excitation to the S3 state. Due to the negligible energy gap between the S3,min and the S3/S2 MECI, the wave packet on S3 quickly transfers to S2 in less than 10 fs and subsequently to S1 by ∼20 fs. This ultrafast IC from both S2 and S3 to S1 is consistent with the experimental observation.12
| PM650 | PM597 | ||
|---|---|---|---|
| Exp. (fs) | Calc. (fs) | Exp. (fs) | Calc. (fs) |
| <20 | 9.6 | 51 | 85 |
To investigate the collective role of vibrational modes, we performed additional MCTDH simulations with various combinations of vibrational modes. One such case is shown in Fig. 7(c), where nine specific vibrational modes at frequencies of 31.5, 35.5, 40.3, 294.4, 1228.4, 1266.3, 1451.0, 1522.0, and 1629.2 cm−1 are selectively excluded from the simulation. A marked slowdown in population decay was observed compared to that in Fig. 7(a). Notably, the exclusion of the 1522.0 cm−1 mode, along with the others, led to a significantly reduced rate of population transfer. This gradual decay highlights the critical role of the collective contribution of multiple vibrational modes to the ultrafast IC in polyatomic molecules.
The population dynamics shown in Fig. 7(a) and (b) exhibit oscillatory features indicative of coherent nuclear motions. The population traces were fit using a tri-exponential function, and the residual was Fourier transformed to obtain the spectra shown in Fig. 8. Distinct peaks appear at 1242 and 1438 cm−1, corresponding to the vibrational modes of the ν67 and ν75 modes, respectively. This suggests that these modes play a significant role in the IC reaction coordinates and contribute strongly to the vibronic coupling between the S1 and S2 states.
![]() | ||
| Fig. 8 Fourier transform (FT) power spectra of the oscillation part of the populations of (a) S1 and (b) S2 states obtained from the MCTDH simulation following excitation to S2 state. | ||
The analysis of specific vibrational modes involved in the early stages of excited-state dynamics can be performed by tracking the temporal evolution of nuclear density distributions obtained from MCTDH simulations. As an example, Fig. S3 shows the nuclear distributions on the S1–S3 states along the Q67 and Q75 vibrational modes at various time intervals following vertical excitation to the S2 state. As early as 5 fs, substantial nuclear density is already observed in the S1 state, along with some nuclear density in the S3 state. By 20 fs, the nuclear density is distributed mostly in S1 state, as reflected in the population dynamics in Fig. 7. Since the vibrational periods of the Q67 and Q75 modes are 26 and 23 fs, respectively, the initial nuclear wave packet has traversed nearly a full circle in this 20 fs timeframe.
The diabatic electronic populations of the S1–S3 states following excitation to S2 or S3 state in PM597 are shown in Fig. 9. In contrast to the nearly instantaneous depopulation of the S2 state in PM650, the decay of the wave packet in PM597 is substantially slower. The rise time of the S1 population and the decay time of the S2 population are both approximately 85 fs (Fig. S2), which is much slower than that in PM650. As discussed in Section 3.2, this difference may be attributed to variations in the energetics and spatial location of the S2/S1 MECI within the multidimensional coordinate space. The S2/S1 MECI in PM650 lies 0.44 eV below the FC point, while in PM597, it lies 0.3362 eV below. In addition, in PM597, the S2/S1 MECI lies slightly above the S2,min by 0.05 eV, which could also contribute to the slower IC.
When PM597 is excited to the S3 state, the S3 population undergoes ultrafast decay to S2 within ∼10 fs, which is similar to the S3 → S2 IC time in PM650, followed by a much slower transition to S1 on a timescale similar to that of direct S2 excitation. Interestingly, the S3 → S2 IC rates of PM650 and PM597 are comparable, even though the S3/S2 MECI lies above both the S3 FC point and the S3,min, whereas it lies below in PM650.
We also performed additional MCTDH simulations for PM597 with various combinations of vibrational modes. One such example is shown in Fig. 9(c), where five vibrational modes with frequencies of 1007, 1147.8, 1271.9, 1427, and 1442.1 cm−1 are omitted deliberately. As a result, the IC dynamics for both S2 → S1 and S3 → S2 transitions slowed greatly, with little progress observed even after 1 ps.
The temporal evolution of nuclear density distributions for PM597 was also obtained from MCTDH simulations. Fig. S4 presents a series of snapshots of the wave packet density distributions along the Q96 and Q110 vibrational modes following excitation to the S2 state. At 25 fs, the nuclear density in the S2 state shows minimal change, reflecting the much slower IC dynamics in PM597. The density gradually shifts to the S1 state over approximately 100 fs, as indicated by the population dynamics in Fig. 9. Since the vibrational periods of the Q96 and Q110 modes are 26 and 23 fs, respectively, the nuclear wave packet in the S2 state at 25 fs is returned close to its initial position (center) of this two-dimensional space.
We calculated the CVS of the S1 state following excitation to the S2 or S3 states, using nuclear densities obtained from MCTDH simulations. The oscillation amplitude of a vibrational mode in a TF signal is be proportional to the amplitude of the NWP oscillation multiplied by the vibrational reorganization energy of the corresponding mode in S1.28,43 To compare with the experimental CVS from TF spectroscopy, we first calculated the first moment of the nuclear density in the S1 state as a function of time for each normal mode i, denoted as
i. The amplitude of the ith mode in the CVS is then given by:
Ii ∝ ( i,max − i,min)·ωi i = ( i,max − i,min)·κ1i, | (12) |
i is the displacement of the ith mode between the ground and S1 states.
The experimental and theoretical CVS results are shown in Fig. 10 and 11 for PM650 and PM597, respectively, with the vibrational assignments provided in Tables 4 and 5. Calculated CVS up to 1700 cm−1 are shown in Fig. S5 and S6 for PM650 and PM597, respectively. For PM650, all major peaks in the experimental CVS are well reproduced by the MCTDH-calculated spectra, whereas the agreement is less satisfactory for PM597. Nonetheless, the overall good agreement between the calculated and experimental CVS corroborates the MCTDH simulations presented in this work.
| Amplitude | Exp. ω (cm−1) | Calc. ω (cm−1) | Mode |
|---|---|---|---|
| 2.89 | 153.3 | 151.6 | ν 13 |
| 2.62 | 243.2 | 250.8 | ν 18 |
| 2.09 | 292.0 | 292.0 | ν 22 |
| 0.97 | 392.6 | 386.3 | ν 28 |
| 7.06 | 412.2 | 427.5 | ν 29 |
| 2.82 | 547.7 | 548.5 | ν 33 |
| Amplitude | Exp. ω (cm−1) | Calc. ω (cm−1) | Mode |
|---|---|---|---|
| 1.14 | 94.5 | 76.6 | ν 5 |
| 0.88 | 140.5 | 141.1 | ν 9 |
| 7.06 | 257.1 | 271.8 | ν 23 |
| 2.82 | 489.1 | 510.6 | ν 47 |
| Couplings | PM650 (eV) | PM597 (eV) |
|---|---|---|
| κ 1 | 0.0099 ± 0.014 | 0.0069 ± 0.008 |
| κ 2 | 0.0257 ± 0.035 | 0.0214 ± 0.026 |
| κ 3 | 0.0247 ± 0.032 | 0.0202 ± 0.023 |
| λ 12 | 0.0969 ± 0.066 | 0.0907 ± 0.068 |
| λ 13 | 0.0498 ± 0.046 | 0.0364 ± 0.034 |
| λ 23 | 0.0296 ± 0.032 | 0.0193 ± 0.023 |
One of the main reasons for the large difference in IC rates may originate from the different energetic landscapes along the vibrational modes, in particular the promoting modes. The relative energy of S2/S1 MECI with respect to the S2,min in PM650 is nearly degenerate, whereas S2/S1 MECI is located 400 cm−1 higher than S2,min in PM597. For PM650, the negligible energy barrier allows the wave packet to access the MECI shortly after the photoexcitation is initiated. In addition, the S3/S2 MECI and S3,min in PM650 are nearly degenerate, whereas S3/S2 MECI is located 5800 cm−1 higher than S3,min in PM597. In particular, the S3/S2 MECI is located even higher than the S3 FC point. Considering that the diabatic populations in S3 are significant and that S3/S1 interstate couplings are significant, the S3/S2 MECI may also affect the overall IC rates to the S1 state.
Although the coupling constants and the energetics contributes to some extent, the substantial structural deformation inherent to the QmnCI,i configuration should also contribute to the difference in IC rates for the two molecules. In PM650, the S2/S1 MECI is located close to the FC point, while in PM597, it is more distant. PM650 requires minimum structural reorganization to access MECI due to the close positioning of MECI to FC point. In contrast, PM597 requires more extensive distortion as the MECI located farther away from the FC point. It is important to note that in this structural relaxation, the multimode effect is critical to reach the CI in this nonadiabatic process. The multimode effect is clearly demonstrated in the simulation with reduced DOFs, where the IC is decelerated significantly in both molecules.
All vibrational modes in PM650 that show large amplitude of motion in S1 state have strong intrastate coupling to S2 except the 250.8 (ν18) cm−1 mode. This is natural as these modes are excited strongly by the initial photoexcitation to the S2 state. The ν18 mode, however, is not activated by the photoexcitation due to its negligible intrastate coupling, but appear strongly in the experimental CVS indicating that it is activated in the course of the IC process from S2 to S1. Because the IC process is essentially impulsive for all the vibrational modes except the high frequency CH stretching modes, any vibrational modes that has large displacement between the ground (FC) state and S1 states should be excited strongly. The ν18 mode indeed has a large displacement (large κ1) between the ground and S1 states. Therefore, the CVS calculated simply from the displacements matches well with the experiment in PM650. In contrast, for PM597, where the IC process is slow and therefore not impulsive, the CVS calculated from the displacement does not match well, because details of the IC process influence the formation of the NWPs in S1.
Interestingly, the S1 and S2 state populations in PM650 exhibit oscillatory behaviour driven by the in-plane promoting modes ν67 and ν75, which are associated with the S2/S1 CIs. Due to strong vibronic coupling along these coordinates, the wave packets initially populated in the S2 state propagate to the S1 state through the CIs, and a portion of them subsequently return to the S2 state. This clearly indicates that the S1 and S2 PESs are strongly coupled along these two modes, which play a significant role in the ultrafast IC. In contrast, population oscillations are barely observed in PM597, suggesting that there are no dominant vibrational modes facilitating the fast IC. Instead, the IC is likely driven by small contributions from many modes.
The striking difference in IC rates between these structurally similar derivatives is attributed to several computationally identified factors. In PM650, the MECI lies close to the Franck–Condon region, with an almost negligible energy gap between the S2 minimum and the S2/S1 MECI. This allows the nuclear wave packet to rapidly access the MECI following photoexcitation, resulting in the ultrafast IC from S2 to S1. Most of the initial population in S2 transfers to the S1 state within 10 fs. This ultrafast IC process involves only small structural rearrangement. Furthermore, PM650 exhibits strong vibronic coupling along a couple of dominant vibrational modes, providing efficient channels for population transfer. In contrast, PM597 shows a much slower relaxation process from S2 to S1 in approximately 86 fs, substantially slower than that of PM650. This slower decay is attributed to the remote location of the MECI from the FC point and a larger energy gap between the S2 minimum and the S2/S1 MECI. Accessing the MECI in PM597 requires significant structural distortion, and the molecule exhibits weaker vibronic coupling between excited states.
These results extend our understanding of the photophysics of molecules excited to electronic states above S1. Identifying vibrational modes that play significant roles in IC may enable fine-tuning of the IC process by modulating vibronic coupling along these key modes.
Supplementary information (SI): Contains isosurfaces of transition densities for low-lying excited-states, exponential fits of the populations of S1 state following excitation to S2 state, nuclear density contours, CVS calculated from MCTDH simulations, vertical excitation energies of PM650 using different DFT methods, and LVC parameters for the MCTDH simulations. See DOI: https://doi.org/10.1039/d5cp02771c.
| This journal is © the Owner Societies 2025 |