Effect of Nonplanarity on Excited-State Proton ... - ACS Publications

May 31, 2018 - recently, full-dimensional surface-hopping dynamics simula-. Received: .... t t. S. 0. (. )/ 0. 1. 0. (1). This is the simplest functio...
4 downloads 0 Views 6MB Size
Subscriber access provided by University of Massachusetts Amherst Libraries

A: Spectroscopy, Molecular Structure, and Quantum Chemistry

Effect of Non-Planarity on Excited-State Proton Transfer and Internal Conversion in Salicylideneaniline Shiela Pijeau, Donneille Foster, and Edward G. Hohenstein J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.8b02426 • Publication Date (Web): 31 May 2018 Downloaded from http://pubs.acs.org on May 31, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Effect of Non-Planarity on Excited-State Proton Transfer and Internal Conversion in Salicylideneaniline Shiela Pijeau,† Donneille Foster,† and Edward G. Hohenstein∗,†,‡ †Department of Chemistry and Biochemistry, The City College of New York, New York, NY 10031 ‡Ph.D. Program in Chemistry, The Graduate Center of the City University of New York, New York, NY 10016 E-mail: [email protected]

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract Salicylideneaniline (SA) is a prototype for excited-state intramolecular proton transfer (ESIPT) reactions in non-planar molecules. It is generally understood that the dominant photochemical pathway in this molecule is ESIPT followed by nonradiative decay due to twisting about its phenolic bond. However, the presence of a secondary internal conversion pathway resulting from frustrated proton transfer remains a matter of contention. We perform a detailed nonadiabatic dynamics simulation of SA and definitively identify the existence of both reaction pathways, thereby showing the presence of a secondary photochemical pathway and providing insight into the nature of ESIPT dynamics in molecules with nonplanar ground state geometries.

Introduction Excited-state intramolecular proton transfer (ESIPT) is a fundamental photochemical reaction found in a variety of molecular systems across chemistry and biology. Chromophores capable of ESIPT are key design elements in synthetic photoswitches, optical storage devices, sunscreens and photostabilizers. 1–10 Examples of ESIPT also occur naturally in certain dyes and bioflavonoids. 11–17 Many of the prototypical and most well-studied examples of ESIPT contain planar chromophores where this reaction occurs. In these cases, there is often a strong driving force promoting ESIPT and the quantum yield of these reactions approaches unity. In a non-planar molecule, however, there may be additional pathways competing with the ESIPT reaction. Since non-planar molecules exhibiting ESIPT are less common, the photochemistry of these molecules is less well-characterized. A detailed understanding of the effect of nonplanarity is essential to the rational control of ESIPT in non-planar molecules or in sterically constrained environments where non-planar rotamers may exist. Of the nonplanar molecules that exhibit ESIPT, the aromatic Schiff base salicylideneaniline (SA) and its derivatives have perhaps been the most thoroughly studied both experimentally and theoretically. 18–43 The ESIPT in SA is ultrafast and has been observed to complete

2

ACS Paragon Plus Environment

Page 2 of 35

Page 3 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

within 50 fs of photoexcitation. 29,31,34 Following ESIPT, SA undergoes twisting about its phenolic bond; the excited-state lifetime of SA is controlled by this twisting motion and is fairly short. Recent estimates of the excited-state lifetime from time-resolved photoelectron spectroscopy (TRPES) are on the order of 1 ps in the gas phase. 34 However, the origin of the excited-state dynamics leading to these observations remains contentious. Spectroscopic investigations of SA have indicated the presence of two internal conversion pathways. 27,34,39 There is agreement that the dominant deactivation pathway in SA results from twisting about the phenolic carbon-carbon bond that follows ESIPT (the a3 dihedral angle in Figure 1). It has been speculated by some that the rate and quantum yield of the ESIPT reaction is related to the twist of the anilino ring (the a1 dihedral angle in Figure 1). 34 A secondary internal conversion pathway resulting from frustrated ESIPT and twisting about the central carbon-nitrogen bond (the a2 dihedral angle in Figure 1) has also been suggested. 27 Early theoretical work on this system has indicated that both decay pathways are viable, although estimates of the branching ratio were not obtained. 32 This work leveraged one-dimensional quantum dynamics on potential energy surfaces generated with time-dependent density functional theory (TDDFT). More recently, full-dimensional surface-hopping dynamics simulations have been performed using semi-empirical configuration interaction potential surfaces. 35 Surprisingly, these more detailed simulations indicated the presence of only a single photochemical pathway: ESIPT followed by internal conversion mediated by twisting about the phenolic carbon-carbon bond. As of yet, the discrepancy between these dynamics simulations and the experimental observations has not been resolved. Here, we address this question by performing full-dimensional ab initio multiple spawning (AIMS) 44–46 quantum dynamics simulations of SA on potential energy surfaces determined on-the-fly using multireference complete active space configuration interaction (CASCI) wavefunctions with embedding corrections from density functional theory. 47 This approach has been successfully applied to the nonadiabatic dynamics of several molecules that are closely related to SA. 47–49 We observe two distinct deactiva-

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

tion pathways in our simulation and, for the first time, provide a theoretical estimate of the branching ratio between ESIPT and frustrated ESIPT. Further, we identify the deficiencies in the previous simulation 35 that lead to the discrepancy between theory and experiment.

Computational Methods The potential energy surfaces used in this work were obtained on-the-fly with the floating occupation molecular orbital complete active space configuration interaction (FOMO-CASCI) method (using a temperature parameter of 0.35 au and Gaussian broadening of the orbital energy levels). 50,51 A minimal active space consisting of two electrons in two orbitals was used along with a 6-31G∗∗ basis set. To incorporate some of the dynamic electron correlation neglected by this approach, a DFT-embedding correction was applied. 47 This correction was evaluated with the ωPBEh functional. 52 Using this potential energy surface, an AIMS simulation of the excited-state dynamics of SA was performed. 44,45 The energies, analytic gradients 53 and analytic nonadiabatic coupling vectors 54 of the FOMO-CASCI method were computed using the graphical processing unit accelerated implementation in the TeraChem electronic structure package. 55–59 The simulation consisted of 360 initial trajectory basis functions on the S1 electronic state; after considering spawning events, the total simulation included more than 2000 trajectory basis functions. The basis functions were propagated in time for 2 ps or until the population on S1 fell below 0.01; the time integration used 20.0 au time steps. Initial positions and momenta of the trajectory basis functions were sampled from a Wigner distribution of the ground-state harmonic vibrational wavefunction on the S0 electronic state computed at the B3LYP/6-31G∗∗ level of theory.

4

ACS Paragon Plus Environment

Page 4 of 35

Page 5 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Results & Discussion We observe two distinct photochemical pathways in our AIMS simulation. At the Franck-Condon point, SA is in a nonplanar geometry characterized by a twist of approximately 37◦ about the anilic bond (the a1 dihedral angle in Figure 1). Over the first 100 fs following photoexcitation, SA planarizes about the anilic bond. The dominant photochemical pathway is characterized by an ESIPT reaction followed by twisting about the phenolic bond (the a3 dihedral angle); this motion leads to a conical intersection where quenching of the excited state occurs. A secondary pathway resulting from frustrated proton transfer is also present. When the ESIPT reaction is unsuccessful, SA twists about the central carbon-nitrogen bond (the a2 dihedral angle) until a conical intersection is reached and internal conversion occurs. The bifurcation of the initial wavepacket results from competition between the ESIPT reaction and twisting about the central carbon-nitrogen bond; the time-evolution of the wavepacket along these two degrees of freedom is shown in Figure 2. When ESIPT occurs, the twisting motion is impeded and SA becomes trapped in a planar S1 keto minimum. If SA is able to twist sufficiently, the ESIPT reaction is inhibited and the molecule continues to rapidly twist about the central carbon-nitrogen bond (the a2 dihedral angle). This twisting motion leads to a seam of conical intersection where deactivation of the excited state is highly efficient. The portion of the wavepacket trapped in the S1 keto minimum will decay via twisting about the phenolic bond (the a3 dihedral angle) on a somewhat longer timescale. These two pathways exhibit significantly different rates of internal conversion. When proton transfer occurs, the molecule tends to planarize and the wavepacket is trapped in a metastable keto minimum on the S1 electronic state; 80% of the trajectories follow this pathway. The other 20% of the trajectories do not undergo proton transfer and instead decay via the twisted minimum energy conical intersection (MECI). As is evident in Figure 3, the excited state lifetime of the trajectories that undergo excited-state proton transfer is considerably longer than those that do not.

5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 35

To quantify the difference in time scales, we separate the S1 population into the group of trajectories that decay via the keto MECI and those that decay via the twisted enol MECI. This classification is determined by the geometry of the first basis function spawned by each initial trajectory basis function; if the function is spawned at an enol geometry, the parent function and all of its descendants are classified as following the enol decay pathway (and vice versa for basis functions that initially spawn at a keto geometry). This approach to the classification of trajectory basis functions is not entirely unambiguous, since an initial basis function could, in principle, spawn functions at both MECIs. However, that behavior is observed for only 1 of 360 initial trajectory basis functions and less than 0.04% of the overall population is subject to ambiguous classification. To quantify the rate of internal conversion, the change in population is modeled as delayed exponential decay.

NS1 =

   1

if t < t0

  e−(t−t0 )/τ

if t ≥ t0

(1)

This is the simplest functional form that we have found to accurately describe the population dynamics observed in our simulation. We take the excited state lifetime to be the sum of the onset time, t0 , and the rate of exponential decay, τ . Our simulation predicts the trajectories that decay through the keto MECI to have an excited-state lifetime of approximately 800 fs; those that decay via the enol MECI have an excited-state lifetime of approximately 250 fs. The longer lifetime (corresponding to the majority of the trajectories) is in reasonable agreement with recent gas-phase measurements of the lifetime: 0.97 ± 0.1 to 1.17 ± 0.1 ps, depending on the choice of pump wavelength. 34 The shorter time constant resulting from frustrated proton transfer and decay through the enol MECI has not been measured experimentally; however, previous theoretical estimates using a 1-D model potential predict the time constant to be 38 fs. 32 This seems to be unreasonably fast and the present estimate of 250 fs obtained from these full-dimensional computations is expected to be more realistic. To extract the

6

ACS Paragon Plus Environment

Page 7 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

timescale of the proton transfer reaction, exponential functions with and without preexponential factors are fit to the time-dependent population of the S1 enol state (see Fig. 4); time constants of 46 and 52 fs are obtained, respectively. This is in remarkably good agreement with the experimental estimates of the excited-state proton transfer rate in the gas-phase. The experimental rates of 40±20 to 50±20 fs are obtained from by TRPES. 34 The variation of the time constants again results from dependence on the wavelength of the pump pulse. The predicted rates of ESIPT are also in good agreement with earlier experimental measurements that find the time scale of the proton transfer reaction to be roughly 50 fs; 29,31 the present result is also in excellent agreement with a previous theoretical prediction of 50 fs. 32 Competition between proton transfer and twisting motions about the central carbonnitrogen bond leads to the bifurcation of the wavepacket. Following photoexcitation to the S1 state, motion along either the proton transfer or twisting coordinate is favorable. Once significant progress is made in one of these directions, motion in the other direction becomes unfavorable. This can be seen clearly in Fig. 5 where initial progress along the proton transfer coordinate occurs at geometries with 180◦ twist angles and at later times, around 100 fs after excitation, proton transfer occurs at geometries with twist angles between 140◦ and 160◦ . The more rapid proton transfer tends to out-compete the twisting motion and dominates the dynamics; about 80% of the population undergoes proton transfer and decays via the keto MECI (Fig. 3). Analysis of the initial positions of each trajectory basis function suggests that the initial value of the proton transfer coordinate may be correlated with the success of the proton transfer reaction. Perhaps unsurprisingly, the portion of the wavepacket located initially at more negative values of the proton transfer coordinate has a lower ESIPT quantum yield than the portion of the wave packet with more initial progress along the reaction coordinate. Trajectories that follow the keto pathway have initial values of the proton transfer coordinate of -1.34±0.04˚ A, while those that follow the enol pathway have an initial value of -1.42±0.08˚ A. The initial values of the twist angles are less predictive of the outcome of the reaction (177.5±0.6◦ for the keto pathway and 176.9±1.0◦ for

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the enol pathway). A qualitative picture of this process is that SA begins to planarize about its anilic bond and twist about its central carbon-nitrogen bond immediately following excitation (see Fig. 5 and 6). If the proton transfer reaction can complete before the carbon-nitrogen twist becomes too significant, this twisting motion will cease and SA becomes trapped in the planar S1 keto minimum. Since the proton transfer reaction is unfavorable in the twisted molecule, if the reaction does not occur within approximately the first 100 fs, it is unlikely to occur at all (Fig. 5). When SA is trapped in the keto minimum, motion along the anilic and phenolic twisting coordinates eventually leads to a conical intersection (Fig. 6). Although the keto MECI is characterized by a twist about the phenolic angle alone, the seam of intersection can also be reached by simultaneous twisting about both the anilic and phenolic dihedral angles. The enol pathway is can be described entirely as twisting about the central carbon-nitrogen bond; there is very limited motion along the proton transfer coordinate or phenolic twist angles and motion along the anilic dihedral angle ceases once it has become planar (see Fig. 5 and 6). The surface-hopping dynamics simulation of Thiel and coworkers identified only a single photochemical pathway for SA. 35 Their simulation predicts ESIPT to be complete and to occur while SA is planar. This is in obvious disagreement with our AIMS simulation as well as previous work by others. 32,34 The source of the discrepancy seems to be related to the choice of electronic structure method used to determine the potential energy surfaces. Thiel and coworkers apply the semiempirical orthogonalization model 2 with multireference configuration interaction (OM2/MRCI) method. 60–62 The authors report a perfectly planar ground state minimum for SA predicted by this semi-empirical method; all ab initio methods that we have tested predict a non-planar ground state minimum with anil twist angles (angle a1 in Figure 1) of roughly 30◦ to 40◦ (geometric parameters of the optimized ground-state minimum are provided in the Supporting Information). This error severely damages the ability of OM2/MRCI to describe the photophysics of SA since, as demonstrated by Stolow and coworkers, the barrier (or lack thereof) to proton transfer is tied to the anil twist angle indicating that

8

ACS Paragon Plus Environment

Page 8 of 35

Page 9 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

proton transfer is most efficient when SA is in planar geometries. 34 By starting their dynamics simulation at a planar geometry of SA, Thiel and coworkers 35 inadvertently biased their simulation towards successful ESIPT reactions and lost the ability to observe the secondary pathway. It should be noted that this appears to be an error in the OM2/MRCI method itself, rather than an error resulting from a misapplication of the method. This result does, however, suggest a strategy for controlling the quantum yield of ESIPT reactions in derivatives of SA. Structural modifications that reduce the anil twist angle at the Franck-Condon point should promote the ESIPT reaction and suppress the secondary pathway. To verify the accuracy of the potential energy surface used in our dynamics simulation, we compare relative energies of the S1 state at several geometries that are important to the photochemistry of SA. The results of this comparison are shown in Fig. 7. The CASCI method including a DFT-embedding correction is compared to state-averaged complete active space self-consistent field (SA-CASSCF) with a minimal active space (two electrons in two orbitals) as well as a larger active space (12 electrons in 11 orbitals). We compare to the approximate second-order coupled-cluster singles and doubles (CC2) method; CC2 has been shown to provide a robust treatment of ESIPT reactions 63 and a reasonable treatment of electronic states near conical intersections. 64 Finally, we also compare to multistate complete active space secondorder perturbation theory (MS-CASPT2) 65,66 to provide a more rigorous treatment of dynamic electron correlation near regions of intersection. The accuracy of the DFTcorrected CASCI approach compares favorably to these more traditional approaches in its description of relative energies on the S1 potential energy surface. All four methods agree that the ground state minimum has a twisted geometry, while the keto minimum on the S1 state has a planar minimum; this is essential for a qualitatively correct description of the photodynamics of SA. Both our DFT-corrected CASCI approach and CC2 are in good agreement with respect to the ESIPT reaction energy on the S1 state (i.e. the energy difference between the Franck-Condon point and the S1 keto minimum). Also promising is the excellent agreement between SA-CASSCF and our

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

DFT-corrected CASCI in the estimation of the relative energies of the minimum energy conical intersections. This agreement suggests that the minimal active space is indeed sufficient to characterize the region of the S1 potential energy surface accessed in our dynamics simulation. The energetics of the MECI geometries are in reasonable agreement with MS-CASPT2 for the enol intersection, but the DFT-corrected CASCI (and all CASSCF methods) tend to overstabilize the keto intersection. This overstabilization of the keto intersection may account for the slightly underestimated excited-state lifetime we predict. As expected for CASCI-based methods with small active spaces, the vertical excitation energies are significantly overestimated. Although this is certainly not ideal, as has been demonstrated by us and others, 48,49,67 it is the relative energetics of the excited state that are of primary importance to the excited-state dynamics. We expect that the overestimated vertical excitation energies may accelerate the excited-state dynamics to some extent.

Conclusions We have identified two distinct photochemical pathways present in the excited-state dynamics of SA. The major pathway is characterized by an ESIPT reaction and a planar S1 keto photoproduct with an excited-state lifetime of about 1 ps (in the gas phase). The minor pathway results from frustrated proton transfer and has a much shorter excited-state lifetime (250 fs). This resolves the apparent discrepancy between theory and experiment with regard to the existence of the secondary reaction pathway. Deficiencies in the electronic structure method applied in the previous dynamics simulations systematically biased the dynamics towards the ESIPT reaction and generation of the S1 keto product. Although dynamical simulations provide tremendous insight into photochemical processes, their success is inexorably linked to the quality of the potential energy surface (or surfaces) used in the simulation. Since the methods used to determine these potential surfaces often contain quite severe approximations (particularly for larger molecules when the surfaces are generated on-the-fly), it is essential

10

ACS Paragon Plus Environment

Page 10 of 35

Page 11 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

to carefully benchmark these approximate methods against more robust approaches to gauge their suitability for treating the problem at hand. Our results indicate that the secondary pathway due to frustrated proton transfer must be blocked in order to control the photochemistry of the SA molecule. The photochromic isomerization product of SA is produced by the main decay channel that follows successful proton transfer. The yield of the proton transfer reaction could be enhanced by the addition of substituents that lead the molecule to more planar ground state geometries or that lower the barrier to proton transfer at nonplanar geometries. It may also be possible enhance the yield of the proton transfer reaction by introducing SA into sterically constraining environments; perhaps even a viscous solvent could be sufficient to alter the quantum yield of the ESIPT reaction. Finally, we note that due to the presence of multiple photochemical pathways, SA would make an intriguing candidate for coherent control experiments or other techniques aimed at manipulating the intrinsic photochemistry of a molecule.

Acknowledgement Support for this project was provided by the Martin & Michele Cohen Fund for Science and PSC-CUNY Award #60719-00 48, jointly funded by The Professional Staff Congress and The City University of New York. Computational resources were provided through a Research Cluster Grant from Silicon Mechanics: award number SM2015-289297.

Supporting Information Available Energies and optimized geometries of salicylideneaniline are provided. This material is available free of charge via the Internet at http://pubs.acs.org/.

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References (1) Flom, S. R.; Barbara, P. F. The photodynamics of 2-(2-hydroxy-5-methylphenyl)benzotriazole in low-temperature organic glasses. Chem. Phys. Lett. 1983, 94, 488. (2) Wiechmann, M.; Port, H.; Frey, W.; L¨armer, F.; El¨asser, T. Time-resolved spectroscopy on ultrafast proton transfer in 2-(2’-hydroxy-5’-methylphenyl)benzotriazole in liquid and polymer environments. J. Phys. Chem. 1991, 95, 1918. (3) Chudoba, C.; Riedle, E.; Pfeiffer, M.; Elsaesser, T. Vibrational coherence in ultrafast excited state proton transfer. Chem. Phys. Lett. 1996, 263, 622–628. (4) Seo, J.; Kim, S.; Park, S. Y. Strong Solvatochromic Fluorescence from the Intramolecular Charge-Transfer State Created by Excited-State Intramolecular Proton Transfer. J. Am. Chem. Soc. 2004, 126, 11154–11155. (5) Kim, S.; Seo, J.; Jung, H. K.; Kim, J.-J.; Park, S. Y. White Luminescence from Polymer Thin Films Containing Excited-State Intramolecular Proton-Transfer Dyes. Adv. Mater. 2005, 17, 2077–2082. (6) Hsieh, C.-C.; Jiang, C.-M.; Chou, P.-T. Recent Experimental Advances on ExcitedState Intramolecular Proton Coupled Electron Transfer Reaction. Acc. Chem. Res. 2010, 43, 1364–1374. (7) Kwon, J. E.; Park, S. Y. Advanced Organic Optoelectronic Materials: Harnessing Excited-State Intramolecular Proton Transfer (ESIPT) Process. Adv. Mater. 2011, 23, 3615–3642. (8) Zhao, J.; Ji, S.; Chen, Y.; Guo, H.; Yang, P. Excited state intramolecular proton transfer (ESIPT): from principal photophysics to the development of new chromophores and applications in fluorescent molecular probes and luminescent materials. Phys. Chem. Chem. Phys. 2012, 14, 8803–8817. 12

ACS Paragon Plus Environment

Page 12 of 35

Page 13 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(9) Baker, L. A.; Horbury, M. D.; Greenough, S. E.; Coulter, P. M.; Karsili, T. N. V.; Roberts, G. M.; Orr-Ewing, A. J.; Ashfold, M. N. R.; Stavros, V. G. Probing the Ultrafast Energy Dissipation Mechanism of the Sunscreen Oxybenzone after UVA Irradiation. J. Phys. Chem. Lett. 2015, 6, 1363–1368. (10) Padalkar, V. S.; Seki, S. Excited-state intramolecular proton-transfer (ESIPT)-inspired solid state emitters. Chem. Soc. Rev. 2016, 45, 169–202. (11) Sengupta, P. K.; Kasha, M. Excited state proton-transfer spectroscopy of 3hydroxyflavone and quercetin. Chem. Phys. Lett. 1979, 68, 382–385. (12) McMorrow, D.;

Kasha, M. Intramolecular excited-state proton-transfer in 3-

hydroxyflavone - Hydrogen-bonding solvent perturbations. J. Phys. Chem. 1984, 88, 2235–2243. (13) Strandjord, A. J.; Barbara, P. The Proton-Transfer Kinetics of 3-Hydroxyflavone: Solvent Effects. J. Phys. Chem. 1985, 89, 2355–2361. (14) Chatterley, A. S.; Horke, D. A.; Verlet, J. R. R. On the intrinsic photophysics of indigo: a time-resolved photoelectron spectroscopy study of the indigo carmine dianion. Phys. Chem. Chem. Phys. 2012, 14, 16155–16161. (15) Simkovitch, R.; Huppert, D. Excited-State Intramolecular Proton Transfer of the Natural Product Quercetin. J. Phys. Chem. B 2015, 119, 10244–10251. (16) Baker, L. A.; Horbury, M. D.; Greenough, S. E.; Allais, F.; Walsh, P. S.; Habershon, S.; Stavros, V. G. Ultrafast Photoprotecting Sunscreens in Natural Plants. J. Phys. Chem. Lett. 2016, 7, 56–61. (17) Baker, L. A.; Greenough, S. E.; Stavros, V. G. A Perspective on the Ultrafast Photochemistry of Solution-Phase Sunscreen Molecules. J. Phys. Chem. Lett. 2016, 7, 4655–4665. 13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(18) Barbara, P. F.; Rentzepis, P. M.; Brus, L. E. Photochemical Kinetics of Salicylidenaniline. J. Am. Chem. Soc. 1980, 102, 2786–2791. (19) Hadjoudis, E. Photochromism and thermochromism of N-salicylideneanilines and Nsalicylideneamino-pyridines. J. Photochem. 1981, 17, 355–363. (20) Higelin, D.; Sixl, H. Spectroscopic studies of the photochromism of N-salicylideneaniline mixed crystals and glasses. Chem. Phys. 1983, 77, 391–400. (21) Becker, R. S.; Lenoble, C.; Zein, A. A Comprehensive Investigation of the Photophysics and Photochemistry of Salicylideneaniline and Derivatives of Phenylbenzothiazole Including Solvent Effects. J. Phys. Chem. 1987, 91, 3509–3517. (22) Hoshino, N.; Inabe, T.; Mitani, T.; Maruyama, Y. Structure and Optical Properties of a Thermochromic Schiff Base. Thermally Induced Intramolecular Proton transfer in the N,N0 -Bis(salicylidene)-p-phenylenediamine Crystals. Bull. Chem. Soc. Jpn. 1988, 61, 4207–4214. (23) Grabowska, A.; Kownacki, K.; Karpiuk, J.; Dobrin, S.; Kaczmarek, L. Photochromism and proton transfer reaction cycle of new internally H-bonded Schiff bases. Chem. Phys. Lett. 1997, 267, 132–140. (24) Sekikawa, T.; Kobayashi, T.; Inabe, T. Femtosecond Fluorescence Study of the Substitution Effect on the Proton Transfer in Thermochromic Salicylideneaniline Crystals. J. Phys. Chem. A 1997, 101, 644–649. (25) Zgierski, M. Z.; Grabowska, A. Photochromism of salicylideneaniline (SA). How the photochromic transient is created: A theoretical approach. J. Chem. Phys. 2000, 112, 6329. (26) Otsubo, N.; Okabe, C.; Mori, H.; Sakota, K.; Amimoto, K.; Kawato, T.;

14

ACS Paragon Plus Environment

Page 14 of 35

Page 15 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Sekiya, H. Excited-state intramolecular proton transfer in photochromic jet-cooled Nsalicylideneaniline. J. Photochem. Photobiol. A 2002, 154, 33–39. (27) Okabe, C.; Nakabayashi, T.; Inokuchi, Y.; Nishi, N.; Sekiya, H. J. Chem. Phys. 2004, 121, 9436. (28) Vargas, V. Time-Resolved Fluorescence of Salicylideneaniline Compounds in Solution. J. Phys. Chem. A 2004, 108, 281–288. (29) Zi´olek, M.; Kubicki, J.; Maciejewski, A.; Naskrecki, R.; Grabowska, A. An ultrafast excited state intramolecular proton transfer (ESPIT) and photochromism of salicylideneaniline (SA) and its “double” analogue salicylaldehyde azine (SAA). A controversial case. Phys. Chem. Chem. Phys. 2004, 6, 4682–4689. (30) Asahi, T.; Masuhara, H.; Nakatani, K.; Sliwa, M. Photochromic Dynamics of Salicylidene Aniline in Solid State by Using Femtosecond Transient Absorption Spectroscopy. Molecular Crystals and Liquid Crystals 2005, 431, 541–548. (31) Rodr´ıguez-C´ordoba, W.; Zugazagoitia, J. S.; Collado-Fregoso, E.; Peon, J. Excited State Intramolecular Proton Transfer in Schiff Bases. Decay of the Locally Excited Enol State Observed by Femtosecond Resolved Fluorescence. J. Phys. Chem. A 2007, 111, 6241–6247. (32) Ortiz-S´anchez, J. M.; Gelabert, R.; Moreno, M.; Lluch, J. M. Electronic-structure and quantum dynamical study of the photochromism of the aromatic Schiff base salicylideneaniline. J. Chem. Phys. 2008, 129, 214308. (33) Hadjoudis, E.; Chatziefthimiou, S. D.; Mavridis, I. M. Anils: Photochromism by Htransfer. Curr. Org. Chem. 2009, 13, 269–286. (34) Sekikawa, T.; Schalk, O.; Wu, G.; Boguslavskiy, A. E.; Stolow, A. Initial Processes of

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Proton Transfer in Salicylideneaniline Studied by Time-Resolved Photoelectron Spectroscopy. J. Phys. Chem. A 2013, 117, 2971–2979. (35) Sp¨orkel, L.; Cui, G.; Thiel, W. Photodynamics of Schiff Base Salicylideneaniline: Trajectory Surface-Hopping Simulations. J. Phys. Chem. A 2013, 117, 4574–4583. (36) Avadanei, M.; Cozan, V.; Shova, S.; Paix˜ao, J. A. Solid state photochromism and thermochromism of two related N-salicylidene anilines. Chem. Phys. 2014, 444, 43–51. (37) Hutchins, K. M.; Dutta, S.; Loren, B. P.; MacGillivray, L. R. Co-Crystals of a Salicylideneaniline: Photochromism Involving Planar Dihedral Angles. Chem. Mater. 2014, 26, 3042–3044. (38) Jacquemin, P.-L.; Robeyns, K.; Devillers, M.; Garcia, Y. Reversible photochromism of an N-salicylidene aniline anion. Chem. Commun. 2014, 50, 649. (39) Avadanei, M.; Kus, N.; Cozan, V.; Fausto, R. Structure and Photochemistry of NSalicylidene-p-carnoxyaniline. J. Phys. Chem. A 2015, 119, 9121–9132. (40) Shahab, S.; Filippovich, L.; Aharodnikova, M.; Almodarresiyeh, H. A.; Hajikolaee, F. H.; Kumar, R.; Mashayekhi, M. Photochromic properties of the NSalicylideneaniline in Polyvinyl Butyral matrix: Experimental and theoretical investigations. J. Mol. Struct. 2017, 1134, 530–537. (41) Jagadesan, P.; Whittemore, T.; Beirl, T.; Turro, C.; McGrier, P. L. Excited-State Intramolecular Proton-Transfer Properties of Three Tris(N-Salicylideneaniline)-Based Chromophores with Extended Conjugation. Chem. Eur. J. 2017, 23, 917–925. (42) Avadaneia, M.; Tigoianua, R.; Serpab, C.; Pinab, P.; Cozan, V. Conformational aspects of the photochromic reactivity of two N-salicylidene aniline derivatives in a polymer matrix. J. Photochem. Photobiol. A 2017, 332, 475–486.

16

ACS Paragon Plus Environment

Page 16 of 35

Page 17 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(43) Hao, J.; Yang, Y. Insight into the new excited-state intramolecular proton transfer (ESIPT) mechanism of N,N0 -bis(salicylidene)-p-phenylenediamine (p-BSP). Chem. Phys. 2018, 501, 53–59. (44) Mart´ınez, T. J.; Ben-Nun, M.; Levine, R. D. Multi-Electronic-State Molecular Dynamics: A Wave Function Approach with Applications. J. Phys. Chem. 1996, 100, 7884–7895. (45) Ben-Nun, M.; Mart´ınez, T. J. Nonadiabatic molecular dynamics: Validation of the multiple spawning method for a multidimensional problem. J. Chem. Phys. 1998, 108, 7244–7257. (46) Ben-Nun, M.; Quenneville, J.; Mart´ınez, T. J. Ab initio multiple spawning: Photochemistry from first principles molecular dynamics. J. Phys. Chem. A 2000, 104, 5161. (47) Pijeau, S.; Hohenstein, E. G. Improved complete active space configuration interaction energies with a simple correction from density functional theory. J. Chem. Theory Comput. 2017, 13, 1130–1146. (48) Pijeau, S.;

Foster, D.;

Hohenstein, E. G. Excited-State Dynamics of 2-(20 -

Hydroxyphenyl)Benzothiazole: Ultrafast Proton Transfer and Internal Conversion. J. Phys. Chem. A 2017, 121, 4595. (49) Pijeau, S.; Foster, D.; Hohenstein, E. G. Excited-State Dynamics of a Benzotriazole Photostabilizer: 2-(20 -Hydroxy-50 -methylphenyl)benzotriazole. J. Phys. Chem. A 2017, 121, 6377. (50) Granucci, G.; Toniolo, A. Molecular gradients for semiempirical CI wavefunctions with floating occupation molecular orbitals. Chem. Phys. Lett. 2000, 325, 79–85. (51) Slav´ıˇcek, P.; Mart´ınez, T. J. Ab initio floating occupation molecular orbital-complete ac-

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

tive space configuration interaction: An efficient approximation to CASSCF. J. Chem. Phys. 2010, 132, 234102. (52) Rohrdanz, M. A.; Martins, K. M.; Herbert, J. M. A long-range-corrected density functional that performs well for both ground-state properties and time-dependent density functional theory excitation energies, including charge-transfer excited states. J. Chem. Phys. 2009, 130, 054112. (53) Hohenstein, E. G.; Bouduban, M. E. F.; Song, C.; Luehr, N.; Ufimtsev, I. S.; Mart´ınez, T. J. Analytic first derivatives of floating occupation molecular orbitalcomplete active space configuration interaction on graphical processing units. J. Chem. Phys. 2015, 143, 014111. (54) Hohenstein, E. G. Analytic formulation of derivative coupling vectors for complete active space configuration interaction wavefunctions with floating occupation molecular orbitals. J. Chem. Phys. 2016, 145, 174110. (55) Ufimtsev, I. S.; Mart´ınez, T. J. Quantum chemistry on graphical processing units. 1. Strategies for two-electron integral evaluation. J. Chem. Theory Comput. 2008, 4, 222–231. (56) Ufimtsev, I. S.; Mart´ınez, T. J. Quantum chemistry on graphical processing units. 2. Direct self-consistent-field implementation. J. Chem. Theory Comput. 2009, 5, 1004– 1015. (57) Ufimtsev, I. S.; Mart´ınez, T. J. Quantum chemistry on graphical processing units. 3. Analytical energy gradients, geometry optimization, and first principles molecular dynamics. J. Chem. Theory Comput. 2009, 5, 2619–2628. (58) Luehr, N.; Ufimtsev, I. S.; Mart´ınez, T. J. Dynamic Precision for Electron Repulsion Integral Evaluation on Graphical Processing Units (GPUs). J. Chem. Theory Comput. 2011, 7, 949–954. 18

ACS Paragon Plus Environment

Page 18 of 35

Page 19 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(59) Titov, A. V.; Ufimtsev, I. S.; Luehr, N.; Mart´ınez, T. J. Generating Efficient Quantum Chemistry Codes for Novel Architectures. J. Chem. Theory Comput. 2013, 9, 213–221. (60) Weber, W.; Thiel, W. Orthogonalization Corrections for Semiempirical Methods. Theor. Chem. Acc. 2000, 103, 495–506. (61) Koslowski, A.; Beck, M.; Thiel, W. Implementation of a General Multireference Configuration Interaction Procedure with Analytic Gradients in a Semiempirical Context Using the Graphical Unitary Group Approach. J. Comput. Chem. 2003, 24, 714–726. (62) Otte, N.; Scholten, M.; Thiel, W. Looking at Self-Consistent- Charge Density Functional Tight Binding from a Semiempirical Perspective. J. Phys. Chem. A 2007, 111, 5751–5755. (63) Aquino, A. J. A.; Lischka, H.; H¨attig, C. Excited-State Intramolecular Proton Transfer: A Survey of TDDFT and RI-CC2 Excited-State Potential Energy Surfaces. J. Phys. Chem. A 2005, 109, 3201–3208. (64) Tuna, D.; Lefrancois, D.; Wola´ nski, L.; Gozem, S.; Schapiro, I.; Andruni´ow, T.; Dreuw, A.; Olivucci, M. Assessment of Approximate Coupled-Cluster and AlgebraicDiagrammatic-Construction Methods for Ground- and Excited-State Reaction Paths and the Conical-Intersection Seam of a Retinal-Chromophore Model. J. Chem. Theory Comput. 2015, 11, 5758–5781. (65) Finley, J.; Malmqvist, P.-A.; Roos, B. O.; Serrano-Andr´es, L. The multi-state CASPT2 method. Chem. Phys. Lett. 1998, 22, 299–306. (66) Roos, B. O. Theoretical studies of electronically excited states of molecular systems using multiconfigurational perturbation theory. Acc. Chem. Res. 1999, 32, 137–144. (67) Levine, B. G.; Mart´ınez, T. J. Ab initio multiple spawning dynamics of excited butadiene: Role of charge transfer. J. Phys. Chem. A 2009, 113, 12815–12824. 19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 1: Coordinates of SA relevant to its photochemistry. We define three dihedral angles: twisting about the anil carbon-nitrogen bond, a1 , (a), twisting about the central carbonnitrogen bond, a2 , (b) and twisting about the phenol carbon-carbon bond, a3 , (c). In a planar molecule, the value of a1 = 0◦ , a2 = 180◦ , and a3 = 0◦ . We also define two bond lengths relevant to the proton transfer reaction: the rNH and rOH bond lengths in the enol structure (d); the proton transferred keto structure is shown in (e). The proton transfer coordinate is defined as the difference between these bond lengths: rPT = rOH − rNH .

20

ACS Paragon Plus Environment

Page 20 of 35

Page 21 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 2: The AIMS nuclear density on the S1 electronic state as a function of twisting about the central carbon-nitrogen bond, a2 , and the proton transfer coordinate, rPT = rOH − rNH (see Figure 1). Snapshots of the density are shown every 50 fs for the first 350 fs following excitation. The AIMS density is averaged over all 360 trajectories. The horizontal black line represents a dihedral angle of 180◦ .

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 3: The population of the S1 electronic state for the first 2 ps following excitation is shown in green. The total population of S1 is divided into the part that arises from trajectories exhibiting proton transfer (shown in red) and those that do not undergo proton transfer (shown in blue); the population decay of these trajectories is modeled as delayed exponential decay. The decay of the total S1 population is modeled as a sum of those two functions.

22

ACS Paragon Plus Environment

Page 22 of 35

Page 23 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 4: The population of the enol state of SA for the first 150 fs following excitation. The enol state is defined as the set of coordinates with a negative value of the proton transfer coordinate: rPT = rOH − rNH . See Figure 1 for a definition of the proton transfer coordinate. The population is obtained by integrating the AIMS nuclear wavefunction over all negative values of rPT . To isolate the time constant associated with proton transfer, only AIMS trajectories that exhibit proton transfer are included in this analysis.

23

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5: The AIMS nuclear density on the S1 electronic state divided into contributions from from trajectory basis functions that decay via the keto MECI (in red) and those that decay via the enol MECI (in blue). The densities are plotted as a function of the carbonnitrogen dihedral angle, a2 , and the proton transfer coordinate. The horizontal black line represents a dihedral angle of 180◦ and the vertical black line divides the enol and keto structures.

24

ACS Paragon Plus Environment

Page 24 of 35

Page 25 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Figure 6: The AIMS nuclear density on the S1 electronic state divided into contributions from from trajectory basis functions that decay via the keto MECI (in red) and those that decay via the enol MECI (in blue). The densities are plotted as a function of the anilic dihedral angle, a1 , and the phenolic dihedral angle, a3 . The black lines represent dihedral angles of 0◦ .

25

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

0.00 0.00 0.00 0.00 0.00

-2.27 -2.35 -2.27 -1.80 -1.98

S0/S1 CN twist

-0.01 -0.03 -0.06 -0.08 -0.23

Page 26 of 35

-0.93 -1.10 -1.34 -0.90 -1.06

-1.08 -1.45 -1.46 -1.01 -1.32

5.66 5.42 5.12 3.98 3.95

5.47 5.35 4.99 3.84 3.80

3.14 2.67 2.29 2.17 1.91

S0 enol twist

S0 enol planar

S1 keto planar

3.19 2.98 2.36 2.23 1.96

S1 keto twist

-1.93 -2.27 -1.83 -1.32 -1.24

S0/S1 phenol twist

Figure 7: Energies of salicylideneaniline computed with several different methods at geometries relevant to its photochemistry. The excitation energies at certain geometries are shown next to the arrow and reported in eV. Relative energies on the S1 state are given relative to the Franck-Condon point (the S0 enol twist geometry) and are reported in eV. Energies are computed with FOMO-CAS(2/2)-CI/6-31G∗∗ including a DFT correction from ωPBEh (shown in normal print), SA-CAS(2/2)-SCF/6-31G∗∗ (shown in bold print), SACAS(12/11)-SCF/6-31G∗∗ (shown in italicized print), CC2/cc-pVDZ (shown in blue print) and MS-CAS(2/2)-PT2/6-31G∗∗ (shown in red print). Geometries are optimized at the respective level of theory, except in the case of CC2 and MS-CASPT2, where the DFTcorrected CASCI geometries are used.

26

ACS Paragon Plus Environment

Page 27 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Graphical TOC Entry

0

τ

=

25

fs

τ= 800 fs

27

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17

ACS Paragon Plus Environment

Page 28 of 35

Page 29 of 35

The Journal of Physical Chemistry

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40

ACS Paragon Plus Environment

Page 30 of 35

Page 31 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59

ACS Paragon Plus Environment

Page 32 of 35

Page 33 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59

The Journal of Physical Chemistry

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

0.00 0.00 0.00 0.00 0.00

-2.27 -2.35 -2.27 -1.80 -1.98

S0/S1 CN twist

-0.01 -0.03 -0.06 -0.08 -0.23

Page 34 of 35

-0.93 -1.10 -1.34 -0.90 -1.06

-1.08 -1.45 -1.46 -1.01 -1.32

5.66 5.42 5.12 3.98 3.95

5.47 5.35 4.99 3.84 3.80

3.14 2.67 2.29 2.17 1.91

S0 enol twist

S0 enol planar

S1 keto planar

ACS Paragon Plus Environment

3.19 2.98 2.36 2.23 1.96

S1 keto twist

-1.93 -2.27 -1.83 -1.32 -1.24

S0/S1 phenol twist

Page 35 of 35 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

τ

=

0 5

fs

τ= 800 fs

2

ACS Paragon Plus Environment