He
Wang‡
a,
Tianhe
Yang‡
a,
Yuechun
Li
a,
Le
Yu
*a,
Yibo
Lei
*a and
Chaoyuan
Zhu
*bc
aKey Laboratory of Synthetic and Natural Functional Molecule Chemistry of Ministry of Education, College of Chemistry & Materials Science, Shaanxi Key Laboratory of Physico-Inorganic Chemistry, Northwest University, Xi’an, Shaanxi 710127, P. R. China. E-mail: yule@nwu.edu.cn; leiyb@nwu.edu.cn
bDepartment of Applied Chemistry and Institute of Molecular Science, National Yang Ming Chiao-Tung University, Hsinchu 30010, Taiwan. E-mail: cyzhu@mail.nctu.edu.tw
cKey Laboratory of Theoretical Chemistry of Environment, Ministry of Education, School of Environment, South China Normal University, Guangzhou 51006, P. R. China
First published on 1st November 2023
Nonadiabatic molecular dynamics simulations with a global switching algorithm have been performed at the TD-CAM-B3LYP-D3/def2-SVP level of theory for ultrafast photo-induced ring-opening and isomerization reactions upon S1 excitation for 2,2-diphenyl-2H-chromene (DPC). Both DPC-T and DPC-C conformers undergo ring-opening relaxation and isomerization pathways accompanied with pyran conformation conserved and converted on the S1 or S0 states via competition and cooperation between C–O bond dissociation and pyran inversion motions. Upon S1 excitation, the DPC-T mainly relaxes to the T-type conical intersection region and thus yields a higher ring-opening efficiency with a faster S1 decay and intermediate formation than those of the DPC-C mainly relaxing to C-type conical intersection. The simulated ring-opening quantum yield for DPC-T (DPC-C) is 0.91 (0.76), which is in good agreement with the experimental value of 0.7–0.9, and the thermal weight averaged lifetimes are estimated as 182.0 fs, 228.6 fs, and 1262.4 fs for the excited-state decay, intermediate formation, and ring-opening product, respectively. These time constants are in good agreement with the experimentally measured τ1 time constant of 190–450 fs and τ2 time constant of 1000–1800 fs. The present work could be a valuable reference for understanding the nature of the photorelaxation mechanisms of DPC, and could help to develop DPC-based photoresponsive materials.
Numerous experimental studies utilizing time-resolved transient absorption techniques in the visible and IR regions, transient fluorescence, flash photolysis and NMR spectroscopy have been carried out to trace the generation and evolution of CF and OF intermediates for disclosing the microscopic dynamics of DPC and its derivatives.20–28 The presented models basically agreed with a multi-component decay pattern, however, which intermediate and which electronic state accounts for each of the dynamics step is still debated. Kodama et al.22 investigated a series of chromenes with time-resolved absorption spectra and predicted that the C–O bond cleavage occurs on the S1 state within 2 ps and recovers in a ∼ns to ∼ms time domain. The proposed sequential model consists of two photo-triggered steps, CF photolysis only yields a revertible CC intermediate which subsequently converts into other OF photoproducts. Görner and Chibisov20 measured the absorption spectra of naphthopyran in different solvents with continuous irradiation and proposed a parallel model with competing CF → TC and CF → TT processes taking place on the S1 state. This model was not widely adopted and only in a recent spectra study on bromoalkoxyl substitution effects by Chernikova et al.,28 which was suggested to compete with the sequential model and exhibited sensitive dependency on the solution concentration and length of the bromoalkoxyl group. Successive time-resolved spectroscopy studies attempted to perfect the tri-component sequential model progressively. Gentili et al.23 suggested that an S1 state zwitterionic or biradical character C–O cleavage product is generated within 0.23 ± 0.05 ps. Following a barrierless relaxation pathway, internal conversion occurs and gives rise to a ground state CC intermediate at 1.1 ± 0.1 ps, which ultimately transforms to TC within 1.4 ± 0.5 ps. Moine et al.24 addressed that two minima exist on the S1 state in the vicinity of and far away from the Franck–Condon region prior to the decay pathway. Therefore, the τ1 time constant is assigned as the excited-state evolution towards the far minimum, τ2 for the ring-opening process and τ3 for the equilibrium between TC and TT isomers. This sequential model was refined by Herzog et al.25 with τ1 attributed to the excited-state relaxation towards the CF minimum and τ2 to overcoming the insignificant barrier that hinders the conversion to the ground state TC. However, according to the time-resolved vibrational spectroscopy study on DPC in acetonitrile by Strudwick et al.,27 the formation of excited state species was ruled out and the internal conversion to the ground state occurs in less than a few hundred fs with the CC isomer as the exclusive product. The subsequent sequential CC → TC → TT transformation was attributed to the τ2 and τ3 time constants of 7 and 13 ps, respectively. Similar results were proposed by Brazevic et al.26 with naphthopyran as the model compound.
Theoretical studies for the photochemical mechanisms of chromenes are relatively limited and the adopted models are usually truncated.29–32 Celani et al.29 performed CASSCF calculations for the parent benzopyran and a ring-opening reaction coordinate on the S1 state is proposed as a transition state connecting the CF intermediate with the acyclic S1/S0 conical intersection (CI). Migani et al.30 carried out combined CASPT2//CASSCF calculations for the parent and 2,2-diethyl-2H-chromene and proposed the same photochromic mechanism as that of Celani et al.29 Furthermore, the energy profile along the S1 state ring-opening reaction coordinate was refined with the inclusion of a dynamic correlation energy. The barrierless relaxation slope was observed to encounter a flat initial stage, which is inconsistent with experiment results and this process would finish within a few hundred fs. The quantum yield dependence on the excitation of vibronic modes33,34 is interpreted by the fact that the low-frequency ring-puckering mode has the largest projection onto the reactive out-of-plane mode which promotes the C–O fission process. With 2H-pyran used as the model compound, Hewage31 utilized the maximum overlap method to explore the excited state ring opening process. The conversion from the more stable ring-puckered conformation to the planar form is required to activate the ring-opening process. Galvão32 reported a CR-EOMCCSD(T)//CIS study on benzopyran with both CF and CC intermediates optimized on the S1 state and suggested the internal conversion may occur earlier when approaching CI along reaction path. Besides these truncated models, Chuev et al.35 explored the ground and first excited state pyran-ring opening mechanism of 2H-pyridochromene derivatives utilizing the semi-empirical PM3 method with a multielectron configuration interaction. The formation of TC was proposed to follow a similar CI dominated relaxation pathway, as interpreted by ab initio calculations.29,30 Furthermore, TT formation was attributed to a vertical decay from the OF-type S1 minimum. In a recent series of combined time-resolved spectroscopy and theoretical studies by the Burdzinski group aimed at unmasking the substitution control on naphthopyran photocoloration mechanisms,9,12,13,26 the ADC(2) method with a cc-pVDZ basis set was employed to construct minimum energy potential energy curves (PECs) or 2-dimensional potential energy surfaces (PESs) towards specific reaction coordinates. The sequential CF ↔ CF* → CC → TC → TT mechanisms were verified as the energetically favored reaction pathway, however, the S1 PES connecting the CF Franck–Condon and the S1/S0 conical intersection regions was steeply descending without an initially flat topology. Although the electronic structure calculation studies provided qualitative reaction routes based on key intermediates, the time-dependent dynamical properties, such as lifetimes for reactive transients, quantum yields and branching ratios between different pathways, rely on nonadiabatic molecular dynamics simulations. Among the popular methods, the global switching algorithm36 (without calculating the nonadiabatic coupling vector) based trajectory surface hopping dynamics simulations with time dependent density functional theory (TDDFT) on-the-fly potential energy surfaces has been successfully applied to simulate ultrafast photo-triggered processes of realistic systems.37–39 In the present work, the TD-CAM-B3LYP-D3/def2-SV(P) nonadiabatic molecular dynamics simulations based on a global switching algorithm is utilized to investigate the mechanisms for photo-triggered ring-opening and ensuing ring-opening and isomerization reactions of the DPC. A quantitatively dynamic point of view for the competition and cooperation of ring-opening and ring-inversion pathways as well as the effects of intramolecular hydrogen bonds are presented in the present study.
We made the benchmark calculation by using a static-dynamic-static multi-state multi-reference second-order perturbation method (SDSPT2)44,45 at (TD)-CAM-B3LYP-D3 optimized geometries for ten key electronic structures, as shown in Fig. 2. In SDSPT2 calculations, the active space with 12 electrons in 11 orbitals (12e, 11o) is composed of five π and four π* orbitals of benzopyran as well as the σ and σ* orbitals of the C(spiro)–O bond in which the iCAS46 approach is adopted to ensure active spaces for all the optimized geometries maintaining the same atomic orbital character. The energy profiles calculated by (TD)-CAM-B3LYP-D3/def2-SVP are in good agreement with those calculated by SDSPT2/def2-SVP, as shown in Fig. 3 and in Table S1 (ESI†) (the relative energy differences between these two methods are less than 0.3 eV except for S1-C). Actually, the SDSPT2 calculation indicates that the single electron excitation configurations are dominant for the S1 state in all the key intermediates involved in the DPC ring-opening reaction. Therefore, the TD-CAM-B3LYP-D3/def2-SVP is a suitable choice for investigating the photochemical reaction of DPC starting from the S1 state.
![]() | ||
| Fig. 2 Optimized geometry structures of ten critical points for DPC (bond length in angstroms and dihedral angles in degrees) by (TD)-CAM-B3LYP-D3/def2-SVP. | ||
The Grimme's D3 dispersion correction47 is adopted to achieve a better description of the intramolecular hydrogen bonding. In the geometry optimization and single point energy calculation, the def2-SVP basis set48 is used. All local minima on both the S0 and S1 states are optimized, and the linear interpolated internal coordinate (LIIC) PES scans and ground state minimum energy reaction paths are calculated by using the ORCA49 program. Moreover, the conical intersections between the S1 and S0 states are optimized at the TD-CAM-B3LYP-D3/def2-SVP level by a BDF program suite50 and their energies are recalculated by ORCA with the same method. The multistate SDSPT2 calculations are conducted by the Xi’an-CI module51 in the BDF program.
The nonadiabatic molecular dynamics simulations are performed by using the TD-CAM-B3LYP-D3 on-the-fly potential energy surfaces with a global switching based trajectory surface hopping algorithm. The time limit for trajectory propagation is set as 1.5 ps, but for those resonance trajectories without experiencing any attempt at hopping during the evolution, an additional 1.5 ps is added in order to ensure the internal conversion to ground state S0. The step size for both electronic and nuclear motions is chosen as 0.5 fs. Nuclear motion is propagated by numerically integrating the Newtonian equation of motion with the velocity-Verlet method.52 In order to accelerate the trajectory simulation, a slightly small basis set def2-SV(P)48 plus the resolution of identity (RI) approach53 are considered in on-the-fly trajectory surface hopping calculations for both potential energies and energy gradients. Actually, all the optimized structures obtained by def2-SV(P) basis set with RI approximations are basically identical with those calculated with the def2-SVP basis set without RI. A dynamical code from our own group is utilized to do on-the-fly trajectory propagation, and a global switching algorithm36 is used to compute the nonadiabatic hopping probability whenever an avoided crossing point is detected (minimum energy separation between two adjacent adiabatic PESs in three consecutive time steps). The initial geometries and velocities for the sampling trajectories are obtained by random seed sampling with a Boltzmann distribution at a temperature of 300 K. A total of 200 (150) sampling trajectories are run starting from vertically excited DPC-T (DPC-C) to the Franck–Condon region on the S1 state. The TDDFT on-the-fly potential energy surfaces are carried out by using the ORCA49 program.
After a trajectory hops to the ground state S0, we collect metastable cisoid-cis (CC), namely C-type (S0-CC-C) and T-type (S0-CC-T) intermediates, when the rCO distance matches the optimized S0-CC intermediate form. When trajectories propagate to the transoid-cis (TC) conformer, we define the optimized S0-TC as products by detecting two dihedral angles φC2C3C4C5 and φC1C2C3C4 varying in 180 ± 20° and 0 ± 20°, respectively. On the other hand, for nonreactive trajectories back to minima on the S0 state, namely C-type (S0-C) and T-type (S0-T), the terminal condition is set as an rCO distance varying in the 1.40 ± 0.10 Å region.
The energy difference between S0-T and S0-C is 0.07 eV and it is similar to those for the analog molecules of spirobenzopyran.40,41 The conformation difference originates from the pyran ring inversion direction and amplitude of the benzene ring twisting associated with the intramolecular C–H⋯O and C–H⋯π hydrogen bonding. The φO6C5C4C3 values in S0-T and S0-C are 28.4° and −20.9°, respectively. The phenyl ring twisting angles φC13C12C5C4 and φC18C11C5C4 for S0-T (S0-C) are 174.5° (−148.3°) and 90.1° (57.6°), respectively, and this implies that the O6 atom position determines the conformation of the two phenyl rings. Additionally, the bond lengths rH38O6, rH29O6, rH34C12, rH33C4 and rH29C11 in S0-T and S0-C are 2.312 (2.336) Å, 2.660 (3.349) Å, 2.666 (3.025) Å, 2.497 (2.554) Å, and 2.898 (2.533) Å, respectively. The latter three are shorter than the conventional C–H⋯π hydrogen bonding length in T-shape stacked benzene. It is clear that the intramolecular hydrogen bonds between the benzene and pyran rings play an important role in the spatial topology of DPC and would also affect the molecular structure evolution in the photo-induced dynamics processes. The relatively stronger intramolecular hydrogen bonding in S0-T leads to a lower energy in comparison with that of S0-C. The isomerization between S0-T and S0-C could easily take place on the S0 state since this process is basically barrierless, as shown in Fig. S1a (ESI†).
The vertical excitation energy to the S1 state is estimated as 4.67 eV (4.61 eV) from the Franck–Condon region of S0-T (S0-C) and the corresponding oscillator strength is 0.11 (0.12), and this indicates that the S1 state is the lowest “bright” state. The main electronic configuration is the HOMO → LUMO transition for the S1 state. The S1-T minimum has a planar benzopyran moiety which lies in the middle of the S1 state isomerization pathway between the Franck–Condon (FC) regions of DPC-T and DPC-C with the selected internal coordinates close to the average value from S0-T and S0-C. It can be seen from the LIIC PESs of FC-C ↔ S1-T ↔ FC-T in Fig. S1b (ESI†) that the depth of the smooth potential well is about 0.32 eV, and this indicates that the helicity interconversion would take place upon S1 excitation and this plays an important role for the ring-opening process starting from both FC-C and FC-T. On the other hand, the S1-C minimum lies in vicinity of the FC-C region, with a 0.025 Å longer bond length rO6C5 and 0.07 eV lower in energy than S1-T.
The energy difference between FC-C and CI-S1/S0-C (FC-T and CI-S1/S0-T) is 1.62 eV (2.08 eV), in agreement with those reported for benzopyran (1.73 eV by CASPT230) and naphthopyran (1.79 eV by ADC(2)26). Two conical intersections have low energy in comparison with that of the FC regions and thus this could lead to efficient decay processes for the photo-induced ring-opening dynamics of DPC. As depicted in Fig. 2, the geometry structures of CI-S1/S0-C and CI-S1/S0-T are quite different, and this indicates that the decay processes approaching two conical intersections can be quite different from those of its corresponding FC regions. The structure of CI-S1/S0-C is similar to that of S1-C with a relative planar pyran ring moiety, but with a 1.294 Å longer rO6C5 bond and 0.087 Å shorter rO6C1 bond than those of S1-C. This indicates that semi-opened quinoid character in-plane pyran ring expansion of CI-S1/S0-C can be involved in the DPC-C relaxation process on the S1 state. While, in comparison with that of CI-S1/S0-T, it has a 1.12 Å longer rO6C5 bond and 0.076 Å shorter rO6C1 bond than that of S1-T. Comparing with the rO6C5 bond of 2.753 Å (2.564 Å) for CI-S1/S0-C (CI-S1/S0-T), its rH29O6 bond increased to 3.792 Å (3.214 Å). The lengths of other intramolecular hydrogen bonds remain close to those in the corresponding S1 minimum or FC. This indicates that as C5–O6 bond dissociation occurs, the other hydrogen bonds are accompanied with a pyran ring expansion and inversion motion. This can be seen from the fact that the rO6C5 bond of the intermediate local minimum S0-CC-C (S0-CC-T) is increased to 3.161 Å (2.935 Å) in comparison with that of its corresponding CI-S1/S0-C (CI-S1/S0-T). However, the rearrangement of bonds in the pyran ring moiety is completed in S0-CC-C (S0-CC-T), as the rO6C1, rC4C5 and rC2C3 convert into double bonds, while the rC3C4 and rC1C2 become single bonds, and these are opposite to the corresponding CI-S1/S0-C (CI-S1/S0-T). Both S0-CC-C and S0-CC-T approach S0-TC from the opposite direction with two benzene rings exchanging position as seen from the twisting towards φC5C4C3C2. Finally, φC5C4C3C2 for S0-TT (S0-TC) is 173.0° (176.0°), and the relative energies and VEEs for these two isomers are quite similar, which is in agreement with previous experimental reports.20–28
where N is the number of trajectories, and xi is the time constant of each trajectory. We can estimate the trajectory average lifetimes (standard deviation) for reaching CI-S1/S0 hops, to S0-CC intermediates, and to S0-TC products as 141.8 (126.2) fs, 180.0 (135.5) fs and 1271.5 (616.8) fs, respectively, for those starting from the FC-T region, and as 439.8 (355.4) fs, 540.2 (227.2) fs, and 1203.9 (412.6) fs, respectively, for those starting from the FC-C region. We consider the Gibbs free energy difference between S0-T and S0-C to be about 1.1 kcal mol−1 (energy corrected by the def2-TZVP basis set at geometries optimized by def2-SVP). According to the Boltzmann statistical distribution at room temperature (25 °C), the equilibrium weight is estimated as 13.5% at S0-C vs. 86.5% at S0-T, for which the weight averaged lifetimes are estimated for S1 state relaxation to CI-S1/S0 hops, to S0-CC intermediates, and to S0-TC products as 182.0 fs, 228.6 fs, and 1262.4 fs, respectively. The simulation results confirm that the DPC ring opening takes place sequentially, and comparing with the experiments of Gentili et al.,23 Moine et al.,24 Herzog et al.25 and Strudwick et al.,27 who also agree with a sequential model, their predicted τ1 values of 0.23 ± 0.05 ps, 0.45 ps, and 0.20 ± 0.10 ps a and few hundred fs assigned to the S1 intermediate formation or S1/S0 decay can be explained by the first two theoretical averaged lifetimes. Moreover, the simulated S0-TC formation lifetime of 1262.4 fs also agrees with the experimentally fitted time constants τ2 of 1.4 ± 0.5 ps, 1.8 ps and 1.3 ± 0.3 ps. However, the experimentally predicted τ3 time constant of up to tens of ps that is assigned as vibrational cooling or equilibration within S0-TC and S0-TT, which belongs to adiabatic motion on the ground state, is beyond the interest of the present work.
Let us turn to a more detailed analysis about the trajectory hopping spot distributions at CI-S1/S0-T and CI-S1/S0-C in terms of two key internal coordinates, rC5O6 and φC4C3C2C1, as shown in Fig. 5, at which the sampling trajectories are marked as finally ending up with ring closing or opening region. The rC5O6 bond is the main relaxation coordinate that elongates to the hopping region at 2.4–2.9 Å for trajectories starting from both FC-T and FC-C, and this indicates that the pyran ring inversion induced DPC-C ↔ DPC-T isomerization could be taking place on the S1 state. While the hopping region of φC4C3C2C1 is mainly distributed around −30° (+30°), as shown in Fig. 5a (Fig. 5b) for trajectories starting from FC-T (FC-C). Fig. 5 indicates that the DPC-C → DPC-T isomerization ratio is about 18.7% in comparison to the DPC-T → DPC-C isomerization ratio of about 3.0%. This is because the potential well on the S1 state makes trajectories starting from FC-C get more chance to isomerize to DPC-T than the trajectories starting from FC-T, which favors the barrierless relaxation path to CI-S1/S0-T on the S1 state than the potential well hampered DPC-T → DPC-C isomerization. After the sampling trajectories hopping to the ground state, they can form an intermediate state S0-CC, which is described mostly by two key dihedral angles, φC5C4C3C2 and φH24C4C5C12. Trajectories starting from FC-T can form an intermediate state S0-CC-C around φC5C4C3C2 0°–140° and φH24C4C5C12 0°–60°, as shown in Fig. 6a, while trajectories starting from FC-C can form intermediate state S0-CC-T around φC5C4C3C2 −110° to −50° and φH24C4C5C12 ±150° to ±180°, as shown in Fig. 6b. Based on the asynchronous rotation motion of φC5C4C3C2 and φH24C4C5C12, the ground state pyran inversion can be attributed to H24 out of the C5C4C3 plane (HOOP) motion during an aliphatic bridge rearrangement. After a fast H24 twisting motion (finishing at a similar time scale to those that do not experience S0 pyran inversion, ca. 20 fs), the ensuing large amplitude φC5C4C3C2 twisting takes place with a much longer time scale. It has been reported that the HOOP motion plays an important role in the excited state decay of spirobenzopyran,40,41 but for DPC, the influence is less evident. Fig. S2 (ESI†) shows hopping-spot distributions at CI-S1/S0-C and CI-S1/S0-T with respect to the φH24C4C5C3 and φH23C3C4C2 dihedral angles, and HOOP coordinates φH23C3C4C2 and φH24C4C5C3 are distributed from −20° to +20° for trajectories starting from both FC-C and FC-T. HOOP motions of the two typical trajectories starting from FC-T (FC-C) are demonstrated in terms of the rC5O6 bond dissociation against varying of the two HOOP coordinates as shown in Fig. S3 and S4 (ESI†).
![]() | ||
| Fig. 5 Hopping-spot distributions at CI-S1/S0-C and CI-S1/S0-T with respect to the rO6C5 bond and φO6C5C4C3 dihedral angle for trajectories starting from (a) FC-T and (b) FC-C. | ||
![]() | ||
| Fig. 6 Distributions of CI-S1/S0 to S0-CC formations with respect to φH24C4C5C12 and φC5C4C3C2 dihedral angles for trajectories starting from (a) FC-T and (b) FC-C. | ||
Finally, we summarize the relaxation pathways for the sampling trajectories starting from FC-T (FC-C) on the S1 state, as shown in Fig. 7. For the trajectories starting from FC-T, the major S1 state relaxation pathway is C5–O6 bond dissociation followed by hopping via CI-S1/S0-T, which accounts for 97.0% of trajectories, as shown in Fig. 7a. The minor relaxation pathway accounts for 3.0% of trajectories hopping at CI-S1/S0-C, which consists of sequential pyran chirality conversion and C5–O6 bond dissociation. As illustrated in Fig. 7a, once the trajectories decay to the S0 state via CI-S1/S0-T, 9.5% go back to the S0-T reactant, 81.0% evolve directly to the S0-CC-T intermediate, and the rest, 6.5%, yield the S0-CC-C intermediate by means of a ground state HOOP induced aliphatic chain inversion. Similar analysis can be made for relaxation pathways for the sampling trajectories starting from FC-C, as shown in Fig. 7b, in which the major relaxation pathway accounts for 81.3% of trajectories hopping at CI-S1/S0-C, followed by 23.3% back to S0-C, 46.0% converting to S0-CC-C as the C5–O6 elongation and 12.0% transforming to S0-CC-T due to C5C4C3C2 twisting motion. The minor relaxation pathway that inverts the pyran moiety on the S1 state accounts for 18.7% of trajectories hopping at CI-S1/S0-T, followed by 0.7% back to S0-T and the rest go via S0-CC-T before reaching S0-TC. In the next sections, we will demonstrate relaxation mechanisms with time constants for each step as listed in Table S4 (ESI†) based on representative ring-opening trajectories for the sampling trajectories starting from FC-T and FC-C separately.
Footnotes |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3cp04132h |
| ‡ These authors contribute equally to this article. |
| This journal is © the Owner Societies 2023 |