ACS Publications. Most Trusted. Most Cited. Most Read
Revealing Ultrafast Population Transfer between Nearly Degenerate Electronic States
My Activity

Figure 1Loading Img
  • Open Access
Letter

Revealing Ultrafast Population Transfer between Nearly Degenerate Electronic States
Click to copy article linkArticle link copied!

  • Pascal Heim
    Pascal Heim
    Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, Austria
    More by Pascal Heim
  • Sebastian Mai
    Sebastian Mai
    Institute of Theoretical Chemistry, Faculty of Chemistry, University of Vienna, Währinger Str. 17, A-1090 Vienna, Austria
  • Bernhard Thaler
    Bernhard Thaler
    Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, Austria
  • Stefan Cesnik
    Stefan Cesnik
    Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, Austria
  • Davide Avagliano
    Davide Avagliano
    Institute of Theoretical Chemistry, Faculty of Chemistry, University of Vienna, Währinger Str. 17, A-1090 Vienna, Austria
  • Dimitra Bella-Velidou
    Dimitra Bella-Velidou
    Institute of Theoretical Chemistry, Faculty of Chemistry, University of Vienna, Währinger Str. 17, A-1090 Vienna, Austria
  • Wolfgang E. Ernst
    Wolfgang E. Ernst
    Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, Austria
  • Leticia González*
    Leticia González
    Institute of Theoretical Chemistry, Faculty of Chemistry, University of Vienna, Währinger Str. 17, A-1090 Vienna, Austria
    *Email: [email protected]
  • Markus Koch*
    Markus Koch
    Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, Austria
    *Email: [email protected]
    More by Markus Koch
Open PDFSupporting Information (3)

The Journal of Physical Chemistry Letters

Cite this: J. Phys. Chem. Lett. 2020, 11, 4, 1443–1449
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpclett.9b03462
Published January 9, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

Abstract

Click to copy section linkSection link copied!

The response of a molecule to photoexcitation is governed by the coupling of its electronic states. However, if the energetic spacing between the electronically excited states at the Franck–Condon window becomes sufficiently small, it is infeasible to selectively excite and monitor individual states with conventional time-resolved spectroscopy, preventing insight into the energy transfer and relaxation dynamics of the molecule. Here, we demonstrate how the combination of time-resolved spectroscopy and extensive surface hopping dynamics simulations with a global fit approach on individually excited ensembles overcomes this limitation and resolves the dynamics in the n3p Rydberg states in acetone. Photoelectron transients of the three closely spaced states n3px, n3py, and n3pz are used to validate the theoretical results, which in turn allow retrieving a comprehensive kinetic model describing the mutual interactions of these states for the first time.

Copyright © 2020 American Chemical Society
The mechanistic understanding of the initial processes in light-induced molecular excited-state dynamics is both of fundamental interest and essential for the development of various technological applications, for example efficient light-harvesting systems (1) or molecular machines. (2) The first steps of these dynamics, which proceed in the femto- to picosecond range, are often accompanied by nonadiabatic population transfer between the excited states and can therefore be observed in real time with ultrafast time-resolved spectroscopy. (3) Such spectroscopy, in conjunction with state-of-the-art excited-state computational chemistry methods has revealed general concepts for understanding excited-state interactions in molecules (4−7) by using static (8,9) as well as dynamic (10,11) approaches. Nevertheless, following the nonadiabatic dynamics of molecules becomes rather challenging when the energy differences between excited electronic states within the Franck–Condon window becomes small. Although in polyatomic molecules the density of states is generally large, the situation can be particularly demanding in transition metal complexes, (12,13) extended biomolecules, (14) or within the Rydberg manifolds of small molecules, where not only higher-lying states in the Rydberg series are close in energy, (15) but also occasionally low-lying states, (16) especially if they have the same principal quantum number (e.g., n3px, n3py, n3pz). Spectroscopic observation of the dynamical behavior of the individual states is then often buried under the “averaged” population decay of all states within the particular energetic region. These experimental difficulties have two reasons. First, the energetic proximity of the states in combination with the spectral width of sufficiently short laser pulses does not allow exciting single electronic states, but rather creates a complex distribution of population in multiple states. Second, simple sequential or parallel decay models, which can be routinely applied in data analysis via global fitting routines in simpler situations, (17−20) cannot accommodate complex relaxation pathways and state mixing arising from multiple couplings. These two issues obscure the nonadiabatic processes occurring within a dense set of states, requiring specialized approaches to obtain a comprehensive mechanistic understanding of the energy transfer in certain molecules.
In this work, we demonstrate that the combination of time-resolved photoelectron spectroscopy (TRPES) (21,22) and surface hopping simulations with the SHARC method (11) is able to fully characterize the dynamics in the regime of multiple nearly degenerate electronic states. Specifically, we are resolving the population transfer dynamics in the n3p Rydberg manifold of the acetone molecule after two-photon excitation. The photophysical and photochemical processes involving acetone have been in the spotlight since the early days of femtochemistry, (23−25) with a recent renaissance (15,26−30) motivated by the importance of acetone as the simplest aliphatic ketone. The difficulty of resolving light-induced processes in acetone derives from its electronic structure, characterized by series of Rydberg states that are strongly coupled to valence states, with the consequence of complex nonadiabatic relaxation dynamics where the electronic population is cascading down the ladder of Rydberg states. (15,28,30−32) As shown schematically in Figure 1a, the n3p manifold, energetically located at the lower end of the Rydberg series, consists of the three states n3px (A2, 7.34 eV), n3py (A1, 7.40 eV), and n3pz (B2, 7.45 eV), (16) lying within about 100 meV in the Franck–Condon region of the ground state. This manifold of states was investigated in time-resolved experiments and is considered to play an important role in the population transfer to lower states. (26,27,29) The energetic proximity of the three n3p states leads to significant vibrational coupling and—together with the presence of symmetry-dependent couplings with the ππ* (A1) valence state—a complex interaction among all these states. (33−37) As a consequence, the dynamics of the individual n3p states within this dense region has evaded detailed observation to date, despite the importance of the n3p manifold as the bottleneck for the radiationless deactivation pathways of the higher-lying Rydberg states.

Figure 1

Figure 1. (a) Schematic overview of the relevant potential energy surfaces of acetone excited states, including the n3p and ππ* states. (b) Basic kinetic model assumed for the nonadiabatic dynamics among the acetone n3p and ππ* states with the nine considered time constants.

Here, we solve this problem and follow the transient population of the acetone n3p Rydberg states after excitation in terms of the individual nonadiabatic transition time scales between the four most important states: n3px, n3py, n3pz, and ππ*. We note here that it was previously shown in the literature (29,32) that also other dark states (nπ* and n3s) might be involved in the dynamics. Hence, in the kinetic model fits, our label “ππ*” collectively includes also those dark states, although the ππ*, nπ*, and n3s states are explicitly present in both the experiments and the simulations. The transition time scales between the four states are summarized in Figure 1b, which shows the six inter-Rydberg time constants (τxy, τyx, τxz, τzx, τyz, and τzy) and the three Rydberg decay time constants (τ, τ, and τ) considered here. We describe the population dynamics by the following system of differential equations representing a unimolecular first-order kinetic model:
(1)
where is the vector containing the time-dependent state populations. The coupling matrix M contains the nine time constants:
(2)
and g(t)P⃗0 is the source term that populates the states (containing the temporal profile of the laser excitation g(t) and a state-dependent prefactor P⃗0). The main goal of our study is to assign the numeric values of the nine time constants and to investigate their dependence on the excitation energy. This task can be achieved only by a combination of experiment and simulation. The experiment can observe only “effective” time constants τm, which are related to the eigenvalues λm of the coupling matrix M by τm = λm–1 (in the following, m enumerates the eigenvalues of M). These effective time constants correspond to the so-called decay associated spectra (DAS), (17−20) discussed in detail in section S1 in the Supporting Information. As we do not consider the decay of the ππ* state here, one eigenvalue of M is always zero (i.e., one effective time constant is infinite), leaving in principle three time constants that could be obtained from the experiment. In contrast, appropriate nonadiabatic dynamics simulations can directly access the electronic populations of each individual state, what allows obtaining all nine individual time constants from a suitable global fit. The effective time constants derived from the simulated coupling matrix can then be compared to the experimental ones to verify the accuracy of the simulations.
In the TRPES experiments, isolated acetone molecules are excited into the n3p manifold by two-photon pump excitation (320 to 336 nm) and the transient population is probed after a variable time delay via one-photon probe photoionization (402 nm). The time-resolved photoelectron spectrum (see section S2 in the Supporting Information for details on the energy calibration) for 333 nm excitation is exemplarily shown in Figure 2 (see section S3 in the Supporting Information for results at other excitation wavelengths). In the spectrum, the n3px, n3py and n3pz states are observed as separated photoelectron bands. The photoelectron kinetic energies (probe photon energy plus state energy minus ionization potential, see section S2 in the Supporting Information) of the three bands are centered at approximately 0.75, 0.80, and 0.84 eV, respectively. The narrow Franck–Condon envelope of the three bands indicates Rydberg state ionization to the ionic ground state with a Δν = 0 propensity rule. The n3py and n3pz states are simultaneously populated at t = 0 fs and show similar signal increase and decrease. In contrast, the n3px signal behaves differently, building up later and decreasing much more slowly.

Figure 2

Figure 2. Overview of the experimental results. Measured transient photoelectron spectra (a), the 2D global fit of the spectrum (b) with corresponding DAS (c) for 333 nm pump wavelength (Eexc = 7.44 eV). Panel d shows the time constant dependence over excitation energy, including three data points from Hüter and Temps. (27) The red lines in panels a and b indicate a change of the vertical axis scale. PE, photoelectron kinetic energy; Eexc, two-photon excitation energy.

We fit the measured photoelectron spectrum S(E, t) (Figure 2a) with a 2D global fit routine to the following function: (17−20)
(3)
where Θ(t) is the Heaviside step function; τm = λm–1 are the effective time constants, and DASm(E) are the decay-associated spectra (see section S1 in the Supporting Information). By varying the number of decay-associated spectra and time constants contained in the global fit, we find that m = {1, 2} (fit shown in Figure 2b) is sufficient and m = {1, 2, 3} does not increase the fit quality. This finding reflects the apparent similarity of the n3py and n3pz transient signals. Hence, it appears that two of the three relevant effective time constants are very similar and that two time constants are sufficient to describe the time-resolved photoelectron spectrum. The two corresponding DAS (Figure 2c) confirm that n3py and n3pz behave very similar, as given by the two peaks of the DAS corresponding to τ2. Consequently, we can assign the larger constant τ1 to the decay of n3px and the smaller constant τ2 to the parallel decay of the n3py and n3pz states. Unfortunately, the DAS does not allow uncovering more details of the complex population transfer between the different n3p states other than a population transfer from n3py and n3pz to n3px, which is indicated by the negative DAS amplitude of the fast time constant.
The dependence of the effective time constants on the excitation energy is plotted in Figure 2d. Whereas τ2 seems to remain constant at about 150 fs, τ1 is strongly energy-dependent and decreases from about 1800 to 700 fs upon increasing the pump energy. At the highest excitation energy (7.75 eV), our values of τ1 show very good agreement with the results of Hüter and Temps. (27) Moreover, time constants reported by these authors for higher energies show a consistent behavior in the energy dependence. Unfortunately, their energy resolution did not allow separating the different n3p states, so that they did not report decay constants that could be compared to our τ2 values.
As the experiment cannot obtain the individual matrix elements of M, we performed nonadiabatic dynamics simulations using the surface hopping approach SHARC. (11,38) The potential energy surfaces were described employing a linear vibronic coupling (LVC) model (39) composed of 49 diabatic electronic states, including all states from the S0 ground state to the ππ* state, which is the S48 at the Franck–Condon geometry. We propagated (see section S4 in the Supporting Information) three independent ensembles of about 1000 trajectories each, in which the initial electronic wave function was the pure n3px, n3py, or n3pz state. For each ensemble we monitored the population of the n3px, n3py, n3pz, and ππ* states, resulting in a set of 12 population transients. This strategy of simulating the decay of each initial state separately is essential, because there are nine independent time constants to fit and employing only four population transients (from a single ensemble) would lead to severe overfitting.
Fitted population transients are presented in Figure 3. Panels a–c show the temporal evolution of the 12 population transients together with the fitting functions; the kinetic model fits the data excellently, with almost no systematic deviations. The obtained time constants, which are shown in Figure 3d together with their error estimates (obtained by bootstrapping), are quite interesting for the following three reasons: (i) The six inter-Rydberg time constants are relatively similar, with all constants ranging from 390 to 530 fs except τxy (700 fs). (ii) Population transfer toward energetically lower states proceeds faster (τzx, τzy, and τyx) than vice versa (τxz, τyz, and τxy). (iii) According to the three Rydberg decay constants, the n3py state decays the fastest among the n3p states (5 and 13 times faster than n3pz and n3px, respectively), as previously suggested in the literature. (33−37) The time constants for the n3pz and n3px states have large errors because the simulation time (1000 fs) is smaller than these estimated time constants. Especially the time constant for n3px is extremely large and imprecise, which indicates that this time constant is not necessary to describe the population transients. As expected, the fastest decay occurs from the n3py state to the ππ* state, because both states have the same symmetry (A1) and it is sufficient to accumulate energy in one of the totally symmetric modes that lead to an avoided crossing of these states. This situation is different from the decay between states of different symmetry, which requires accumulated energy in at least two normal modes: a totally symmetric tuning mode that makes the diabatic energies equal and a nonsymmetric coupling mode of the correct irreducible representation that induces some coupling between two states. The requirement of two activated normal modes explains why all population transfer processes between states of different symmetry are slower than the same-symmetry n3pyππ* transfer. In general, the magnitude of the time constants suggests that in acetone the n3p Rydberg states readily interconvert among each other because of their energetic closeness and the presence of coupling modes of all required symmetries.

Figure 3

Figure 3. Overview of the theoretical results obtained with the SHARC-LVC method. (a–c) Temporal evolution of the diabatic populations (thin lines) and kinetic model fits (thick lines), for the ensemble starting in (a) n3px, (b) n3py, and (c) n3pz. (d) Scheme showing the fitted time constants and associated errors (see Figure S8 in section S4 in the Supporting Information for a similar fit enforcing detailed balance, showing that τxy might be overestimated and τxz underestimated but otherwise the time constants are consistent with detailed balance). (e) Plot of the energy dependence of the eigenvalue-derived effective time constants and comparison to the experimental and literature (27) values.

In order to obtain information about the energy dependence of the relaxation time constants, we group all trajectories into 11 subensembles based on their individual excitation energies. All fitted time constants are presented in section S5 in the Supporting Information. The energy intervals and the corresponding effective time constants obtained are shown in Figure 3e. The agreement of the energy dependence of the computed effective time constants with its experimental counterpart is very good. The simulated values shown in Figure 3e confirm that there is one strongly energy-dependent effective time constant (τ1), being about 1600 fs for the lowest energies and about 500 fs for the highest ones. Additionally, there are two almost identical effective time constants (τ2 and τ3) in the range of approximately 80–200 fs, only weakly depending on energy.
In conclusion, we demonstrate how a combination of TRPES with extensive surface hopping simulations is able to reveal the details of nonadiabatic dynamics in energetically dense sets of states. Our synergistic approach allows disentangling the interconversion time scales among the three n3p states (n3px, n3py, and n3pz) and the decay time scales to the dark ππ* state in acetone. This work constitutes the first study that resolves the highly nonsequential population flow among these nearly degenerate states. An important key of accessing the individual time constants was the simulation of multiple independent ensembles with complementary initial conditions, as otherwise the global fit is severely under-determined. Despite the known limitations of trajectory surface-hopping methods, (40−42) the results of the simulations—validated through their good agreement with the experimental effective time constants and energy dependence, see also a simulated TRPES spectrum in Figure S20 in the Supporting Information—show that the n3py is most strongly coupled to the ππ* state and exhibits the fastest decay, as expected from symmetry arguments. The n3py also acts as the gateway state for the deactivation of the n3px and n3pz states, as indicated by the fact that the n3px/z→ n3pyππ* route is faster than the direct n3px/zππ* routes. The present results are expected to be helpful in understanding the light-driven dynamics of other aliphatic ketones and designing similar experiments. Further, they also evidence the power of using a linear vibronic coupling Hamiltonian within surface-hopping dynamics, as it allows considering large ensembles of trajectories propagating over a large number of electronic states.

Methods

Click to copy section linkSection link copied!

Experimental Methods. Pump–probe experiments were performed with femtosecond laser pulses from a commercial Ti:sapphire laser system with 800 nm central wavelength (for details, see ref (28)). Pump pulses were frequency up-converted by combining optical parametric amplification and subsequent frequency quadrupling to a wavelength range from 320 to 336 nm (6 nm, 15 meV full width at half-maximum, fwhm). Probe pulses of 402 nm (3 nm, 60 meV fwhm) were obtained by frequency doubling with a BBO crystal. Dichroic mirrors were used in both beam paths to remove undesired wavelengths from the up-conversion process. The laser pulses were focused into the extraction region of a 0.5 m long, linear magnetic bottle time-of-flight spectrometer (43) where a small retarding field was used to increase the electron kinetic energy resolution. High-purity acetone was introduced into the chamber as background gas with a partial pressure of about 4 × 10–6 mbar.
Computational Methods. Potential energy surfaces of all electronic states of acetone were represented with a linear vibronic coupling (LVC) model (39,44) including all 24 vibrational degrees of freedom and 49 electronic states. This large number of states is necessary in order to include the important ππ* state at the reference geometry (S48). The reference harmonic potential was obtained from an optimization plus frequency calculation for the n3s state at the SOS-ADC(2) level of theory. (45) Parameters for the description of the excited-state potentials were obtained from calculations at the same level of theory. The basis set was a combination of cc-pVTZ (46) for C and O, cc-pVDZ (46) for H, and an additional 10s8p6d4f Rydberg basis set (47) at O. All electronic structure calculations were carried out with Turbomole 7.0. (48)
Initial conditions were sampled from the Wigner distribution of the SOS-ADC(2) harmonic oscillator of the ground state, generating 1000 initial geometries and velocities. Three independent sets of trajectories were simulated from these, starting in the n3px, n3py, or n3pz diabatic state. The trajectories were propagated with SHARC2.0 (11,38) for 1000 fs using a 0.5 fs time step; the electronic coefficients were propagated with a 0.02 fs step using the local diabatization approach. (49) We applied an energy-based decoherence correction (50) and rescaled the full momentum vector during a hop.
The results were analyzed in terms of the diabatic state populations, to which the kinetic model was fitted. Additional analysis was carried out by dividing the set of trajectories into different energy windows and fitting their populations independently.
See the Supporting Information for additional computational details.

Supporting Information

Click to copy section linkSection link copied!

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpclett.9b03462.

  • Discussion and derivation of the fitting functions, experimental details regarding energy calibration, additional experimental results, additional computational details, and computational results (PDF)

  • V0 data in SHARC format (TXT)

  • LVC parameters in SHARC format (TXT)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

Click to copy section linkSection link copied!

  • Corresponding Authors
  • Authors
    • Pascal Heim - Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, Austria
    • Sebastian Mai - Institute of Theoretical Chemistry, Faculty of Chemistry, University of Vienna, Währinger Str. 17, A-1090 Vienna, AustriaPresent Address: S.M.: Vienna University of Technology, Vienna, Austria
    • Bernhard Thaler - Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, AustriaOrcidhttp://orcid.org/0000-0001-5412-1519
    • Stefan Cesnik - Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, Austria
    • Davide Avagliano - Institute of Theoretical Chemistry, Faculty of Chemistry, University of Vienna, Währinger Str. 17, A-1090 Vienna, Austria
    • Dimitra Bella-Velidou - Institute of Theoretical Chemistry, Faculty of Chemistry, University of Vienna, Währinger Str. 17, A-1090 Vienna, Austria
    • Wolfgang E. Ernst - Institute of Experimental Physics, Graz University of Technology, Petersgasse 16, A-8010 Graz, AustriaOrcidhttp://orcid.org/0000-0001-8849-5658
  • Author Contributions

    P.H. and S.M. contributed equally to this work.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

Click to copy section linkSection link copied!

P.H., B.T., S.C., W.E.E., and M.K. thank the Austrian Science Fund (FWF) for Grant P29369-N36 and NAWI Graz for financial support. S.M., D.A., and L.G. thank the FWF (Grant I2883) and the University of Vienna for financial support. The authors also thank the European Cost Action Attosecond Chemistry (CA18222). The presented calculations were partly carried out at the Vienna Scientific Cluster (VSC3).

References

Click to copy section linkSection link copied!

This article references 50 other publications.

  1. 1
    Herek, J. L.; Wohlleben, W.; Cogdell, R. J.; Zeidler, D.; Motzkus, M. Quantum control of energy flow in light harvesting. Nature 2002, 417, 533535,  DOI: 10.1038/417533a
  2. 2
    Balzani, V.; Credi, A.; Venturi, M. Light powered molecular machines. Chem. Soc. Rev. 2009, 38, 15421550,  DOI: 10.1039/b806328c
  3. 3
    Weinacht, T., Pearson, B. J., Eds. Time-resolved spectroscopy: An experimental perspective; CRC Press, 2019.
  4. 4
    Blanchet, V.; Zgierski, M. Z.; Seideman, T.; Stolow, A. Discerning vibronic molecular dynamics using time-resolved photoelectron spectroscopy. Nature 1999, 401, 5254,  DOI: 10.1038/43410
  5. 5
    Polli, D.; Altoe, P.; Weingart, O.; Spillane, K. M.; Manzoni, C.; Brida, D.; Tomasello, G.; Orlandi, G.; Kukura, P.; Mathies, R. A. Conical intersection dynamics of the primary photoisomerization event in vision. Nature 2010, 467, 440443,  DOI: 10.1038/nature09346
  6. 6
    Schuurman, M. S.; Stolow, A. Dynamics at conical intersections. Annu. Rev. Phys. Chem. 2018, 69, 427450,  DOI: 10.1146/annurev-physchem-052516-050721
  7. 7
    Timmers, H.; Zhu, X.; Li, Z.; Kobayashi, Y.; Sabbar, M.; Hollstein, M.; Reduzzi, M.; Martínez, T. J.; Neumark, D. M.; Leone, S. R. Disentangling conical intersection and coherent molecular dynamics in methyl bromide with attosecond transient absorption spectroscopy. Nat. Commun. 2019, 10, 3133,  DOI: 10.1038/s41467-019-10789-7
  8. 8
    Garavelli, M. Computational organic photochemistry: Strategy, achievements and perspectives. Theor. Chem. Acc. 2006, 116, 87105,  DOI: 10.1007/s00214-005-0030-z
  9. 9
    Robb, M. A. Theoretical chemistry for electronic excited states; Theoretical and Computational Chemistry Series; The Royal Society of Chemistry, 2018; pp P001225.
  10. 10
    Lasorne, B.; Worth, G. A.; Robb, M. A. Excited-state dynamics. WIREs Comput. Mol. Sci. 2011, 1, 460475,  DOI: 10.1002/wcms.26
  11. 11
    Mai, S.; Marquetand, P.; González, L. Nonadiabatic dynamics: The SHARC approach. WIREs Comput. Mol. Sci. 2018, 8, e1370,  DOI: 10.1002/wcms.1370
  12. 12
    Chergui, M. On the interplay between charge, spin and structural dynamics in transition metal complexes. Dalton Trans 2012, 41, 1302213029,  DOI: 10.1039/c2dt30764b
  13. 13
    Atkins, A. J.; González, L. Trajectory surface-hopping dynamics including intersystem crossing in [Ru(bpy)3]2+. J. Phys. Chem. Lett. 2017, 8, 38403845,  DOI: 10.1021/acs.jpclett.7b01479
  14. 14
    Nogueira, J. J.; Plasser, F.; González, L. Electronic delocalization, charge transfer and hypochromism in the UV absorption spectrum of polyadenine unravelled by multiscale computations and quantitative wavefunction analysis. Chem. Sci. 2017, 8, 56825691,  DOI: 10.1039/C7SC01600J
  15. 15
    Koch, M.; Thaler, B.; Heim, P.; Ernst, W. E. The role of Rydberg-valence coupling in the ultrafast relaxation dynamics of acetone. J. Phys. Chem. A 2017, 121, 63986404,  DOI: 10.1021/acs.jpca.7b05012
  16. 16
    Nobre, M.; Fernandes, A.; da Silva, F. F.; Antunes, R.; Almeida, D.; Kokhan, V.; Hoffmann, S. V.; Mason, N.; Eden, S.; Limão-Vieira, P. The VUV electronic spectroscopy of acetone studied by synchrotron radiation. Phys. Chem. Chem. Phys. 2008, 10, 550560,  DOI: 10.1039/B708580J
  17. 17
    Holzapfel, W.; Finkele, U.; Kaiser, W.; Oesterhelt, D.; Scheer, H.; Stilz, H. U.; Zinth, W. Initial electron-transfer in the reaction center from Rhodobacter sphaeroides. Proc. Natl. Acad. Sci. U. S. A. 1990, 87, 51685172,  DOI: 10.1073/pnas.87.13.5168
  18. 18
    van Stokkum, I. H.; Larsen, D. S.; van Grondelle, R. Global and target analysis of time-resolved spectra. Biochim. Biophys. Acta, Bioenerg. 2004, 1657, 82104,  DOI: 10.1016/j.bbabio.2004.04.011
  19. 19
    Wu, G.; Boguslavskiy, A. E.; Schalk, O.; Schuurman, M. S.; Stolow, A. Ultrafast non-adiabatic dynamics of methyl substituted ethylenes: The π3s Rydberg state. J. Chem. Phys. 2011, 135, 164309,  DOI: 10.1063/1.3652966
  20. 20
    Schalk, O.; Boguslavskiy, A. E.; Stolow, A. Substituent effects on dynamics at conical intersections: Cyclopentadienes. J. Phys. Chem. A 2010, 114, 40584064,  DOI: 10.1021/jp911286s
  21. 21
    Stolow, A.; Bragg, A. E.; Neumark, D. M. Femtosecond time-resolved photoelectron spectroscopy. Chem. Rev. 2004, 104, 17191758,  DOI: 10.1021/cr020683w
  22. 22
    Hertel, I. V.; Radloff, W. Ultrafast dynamics in isolated molecules and molecular clusters. Rep. Prog. Phys. 2006, 69, 1897,  DOI: 10.1088/0034-4885/69/6/R06
  23. 23
    Kim, S. K.; Pedersen, S.; Zewail, A. H. Direct femtosecond observation of the transient intermediate in the α-cleavage reaction of (CH3)2CO to 2CH3+CO: Resolving the issue of concertedness. J. Chem. Phys. 1995, 103, 477,  DOI: 10.1063/1.469614
  24. 24
    Shibata, T.; Suzuki, T. Photofragment ion imaging with femtosecond laser pulses. Chem. Phys. Lett. 1996, 262, 115119,  DOI: 10.1016/0009-2614(96)01024-X
  25. 25
    Buzza, S. A.; Snyder, E. M.; Castleman, A. W. Further direct evidence for stepwise dissociation of acetone and acetone clusters. J. Chem. Phys. 1996, 104, 5040,  DOI: 10.1063/1.471133
  26. 26
    Rusteika, N.; Møller, K. B.; Sølling, T. I. New insights on the photodynamics of acetone excited with 253–288nm femtosecond pulses. Chem. Phys. Lett. 2008, 461, 193197,  DOI: 10.1016/j.cplett.2008.06.079
  27. 27
    Hüter, O.; Temps, F. Ultrafast α–CC bond cleavage of acetone upon excitation to 3p and 3d Rydberg states by femtosecond time-resolved photoelectron imaging. J. Chem. Phys. 2016, 145, 214312,  DOI: 10.1063/1.4971243
  28. 28
    Maierhofer, P.; Bainschab, M.; Thaler, B.; Heim, P.; Ernst, W. E.; Koch, M. Disentangling multichannel photodissociation dynamics in acetone by time-resolved photoelectron-photoion coincidence spectroscopy. J. Phys. Chem. A 2016, 120, 64186423,  DOI: 10.1021/acs.jpca.6b07238
  29. 29
    Couch, D. E.; Kapteyn, H. C.; Murnane, M. M.; Peters, W. K. Uncovering highly-excited state mixing in acetone using ultrafast VUV pulses and coincidence imaging techniques. J. Phys. Chem. A 2017, 121, 23612366,  DOI: 10.1021/acs.jpca.7b01112
  30. 30
    Uenishi, R.; Horio, T.; Suzuki, T. Time-resolved photoelectron imaging of acetone with 9.3 eV photoexcitation. J. Phys. Chem. A 2019, 123, 68486853,  DOI: 10.1021/acs.jpca.9b05179
  31. 31
    Koch, M.; Heim, P.; Thaler, B.; Kitzler, M.; Ernst, W. E. Direct observation of a photochemical activation energy: a case study of acetone photodissociation. J. Phys. B: At., Mol. Opt. Phys. 2017, 50, 125102,  DOI: 10.1088/1361-6455/aa6a71
  32. 32
    Sølling, T. I.; Diau, E. W.-G.; Kötting, C.; De Feyter, S.; Zewail, A. H. Femtochemistry of Norrish Type I reactions: IV. Highly excited ketones—Experimental. ChemPhysChem 2002, 3, 7997,  DOI: 10.1002/1439-7641(20020118)3:1<79::AID-CPHC79>3.0.CO;2-#
  33. 33
    Thakur, S. N.; Guo, D.; Kundu, T.; Goodman, L. Two-photon photoacoustic spectroscopy of acetone 3p Rydberg states. Chem. Phys. Lett. 1992, 199, 335340,  DOI: 10.1016/0009-2614(92)80128-X
  34. 34
    Xing, X.; McDiarmid, R.; Philis, J. G.; Goodman, L. Vibrational assignments in the 3p Rydberg states of acetone. J. Chem. Phys. 1993, 99, 75657573,  DOI: 10.1063/1.465686
  35. 35
    Merchán, M.; Roos, B. O.; McDiarmid, R.; Xing, X. A combined theoretical and experimental determination of the electronic spectrum of acetone. J. Chem. Phys. 1996, 104, 1791,  DOI: 10.1063/1.470976
  36. 36
    ter Steege, D. H. A.; Wirtz, A. C.; Buma, W. J. Vibronic coupling in excited states of acetone. J. Chem. Phys. 2002, 116, 547560,  DOI: 10.1063/1.1423946
  37. 37
    McDiarmid, R.; Xing, X. Nonadiabatic coupling of the 3p Rydberg and ππ* valence states of acetone. J. Chem. Phys. 1997, 107, 675679,  DOI: 10.1063/1.475152
  38. 38
    Mai, S.; Richter, M.; Heindl, M.; Menger, M. F. S. J.; Atkins, A. J.; Ruckenbauer, M.; Plasser, F.; Oppel, M.; Marquetand, P.; González, L. SHARC2.0: Surface Hopping Including Arbitrary Couplings – program package for non-adiabatic dynamics; sharc-md.org, 2018.
  39. 39
    Plasser, F.; Gómez, S.; Mai, S.; González, L. Highly efficient surface hopping dynamics using a linear vibronic coupling model. Phys. Chem. Chem. Phys. 2019, 21, 5769,  DOI: 10.1039/C8CP05662E
  40. 40
    Subotnik, J. E.; Jain, A.; Landry, B.; Petit, A.; Ouyang, W.; Bellonzi, N. Understanding the surface hopping view of electronic transitions and decoherence. Annu. Rev. Phys. Chem. 2016, 67, 387417,  DOI: 10.1146/annurev-physchem-040215-112245
  41. 41
    Wang, L.; Akimov, A.; Prezhdo, O. V. Recent progress in surface hopping: 2011–2015. J. Phys. Chem. Lett. 2016, 7, 21002112,  DOI: 10.1021/acs.jpclett.6b00710
  42. 42
    Plasser, F.; Mai, S.; Fumanal, M.; Gindensperger, E.; Daniel, C.; González, L. Strong influence of decoherence corrections and momentum rescaling in surface hopping dynamics of transition metal complexes. J. Chem. Theory Comput. 2019, 15, 50315045,  DOI: 10.1021/acs.jctc.9b00525
  43. 43
    Kruit, P.; Read, F. Magnetic field paralleliser for 2π electron-spectrometer and electron-image magnifier. J. Phys. E: Sci. Instrum. 1983, 16, 313,  DOI: 10.1088/0022-3735/16/4/016
  44. 44
    Fumanal, M.; Plasser, F.; Mai, S.; Daniel, C.; Gindensperger, E. Interstate vibronic coupling constants between electronic excited states for complex molecules. J. Chem. Phys. 2018, 148, 124119,  DOI: 10.1063/1.5022760
  45. 45
    Dreuw, A.; Wormit, M. The algebraic diagrammatic construction scheme for the polarization propagator for the calculation of excited states. WIREs Comput. Mol. Sci. 2015, 5, 8295,  DOI: 10.1002/wcms.1206
  46. 46
    Dunning, T. H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 10071023,  DOI: 10.1063/1.456153
  47. 47
    Kaufmann, K.; Baumeister, W.; Jungen, M. Universal Gaussian basis sets for an optimum representation of Rydberg and continuum wavefunctions. J. Phys. B: At., Mol. Opt. Phys. 1989, 22, 2223,  DOI: 10.1088/0953-4075/22/14/007
  48. 48
    TURBOMOLE V7.0, A development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 2015.
  49. 49
    Granucci, G.; Persico, M.; Toniolo, A. Direct semiclassical simulation of photochemical processes with semiempirical wave functions. J. Chem. Phys. 2001, 114, 1060810615,  DOI: 10.1063/1.1376633
  50. 50
    Granucci, G.; Persico, M. Critical appraisal of the fewest switches algorithm for surface hopping. J. Chem. Phys. 2007, 126, 134114,  DOI: 10.1063/1.2715585

Cited By

Click to copy section linkSection link copied!
Citation Statements
Explore this article's citation statements on scite.ai

This article is cited by 5 publications.

  1. J. Patrick Zobel, Moritz Heindl, Felix Plasser, Sebastian Mai, Leticia González. Surface Hopping Dynamics on Vibronic Coupling Models. Accounts of Chemical Research 2021, 54 (20) , 3760-3771. https://doi.org/10.1021/acs.accounts.1c00485
  2. Ruaridh Forbes, Simon P. Neville, Martin A. B. Larsen, Anja Röder, Andrey E. Boguslavskiy, Rune Lausten, Michael S. Schuurman, Albert Stolow. Vacuum Ultraviolet Excited State Dynamics of the Smallest Ketone: Acetone. The Journal of Physical Chemistry Letters 2021, 12 (35) , 8541-8547. https://doi.org/10.1021/acs.jpclett.1c02612
  3. Andreas W. Hauser, Martina Havenith, Markus Koch, Martin Sterrer. Festschrift for Wolfgang E. Ernst – electronic and nuclear dynamics and their interplay in molecules, clusters and on surfaces. Physical Chemistry Chemical Physics 2023, 25 (17) , 11880-11882. https://doi.org/10.1039/D3CP90052E
  4. Kaoru Yamazaki, Katsumi Midorikawa. Carbon 1s Edge Induced Femtosecond Nonradiative Decays in Tropone Dication (C7H7O2+). 2022, W4A.8. https://doi.org/10.1364/UP.2022.W4A.8
  5. Julia Westermayr, Philipp Marquetand. Machine learning and excited-state molecular dynamics. Machine Learning: Science and Technology 2020, 1 (4) , 043001. https://doi.org/10.1088/2632-2153/ab9c3e

The Journal of Physical Chemistry Letters

Cite this: J. Phys. Chem. Lett. 2020, 11, 4, 1443–1449
Click to copy citationCitation copied!
https://doi.org/10.1021/acs.jpclett.9b03462
Published January 9, 2020

Copyright © 2020 American Chemical Society. This publication is licensed under CC-BY.

Article Views

2336

Altmetric

-

Citations

Learn about these metrics

Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.

Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.

The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.

  • Abstract

    Figure 1

    Figure 1. (a) Schematic overview of the relevant potential energy surfaces of acetone excited states, including the n3p and ππ* states. (b) Basic kinetic model assumed for the nonadiabatic dynamics among the acetone n3p and ππ* states with the nine considered time constants.

    Figure 2

    Figure 2. Overview of the experimental results. Measured transient photoelectron spectra (a), the 2D global fit of the spectrum (b) with corresponding DAS (c) for 333 nm pump wavelength (Eexc = 7.44 eV). Panel d shows the time constant dependence over excitation energy, including three data points from Hüter and Temps. (27) The red lines in panels a and b indicate a change of the vertical axis scale. PE, photoelectron kinetic energy; Eexc, two-photon excitation energy.

    Figure 3

    Figure 3. Overview of the theoretical results obtained with the SHARC-LVC method. (a–c) Temporal evolution of the diabatic populations (thin lines) and kinetic model fits (thick lines), for the ensemble starting in (a) n3px, (b) n3py, and (c) n3pz. (d) Scheme showing the fitted time constants and associated errors (see Figure S8 in section S4 in the Supporting Information for a similar fit enforcing detailed balance, showing that τxy might be overestimated and τxz underestimated but otherwise the time constants are consistent with detailed balance). (e) Plot of the energy dependence of the eigenvalue-derived effective time constants and comparison to the experimental and literature (27) values.

  • References


    This article references 50 other publications.

    1. 1
      Herek, J. L.; Wohlleben, W.; Cogdell, R. J.; Zeidler, D.; Motzkus, M. Quantum control of energy flow in light harvesting. Nature 2002, 417, 533535,  DOI: 10.1038/417533a
    2. 2
      Balzani, V.; Credi, A.; Venturi, M. Light powered molecular machines. Chem. Soc. Rev. 2009, 38, 15421550,  DOI: 10.1039/b806328c
    3. 3
      Weinacht, T., Pearson, B. J., Eds. Time-resolved spectroscopy: An experimental perspective; CRC Press, 2019.
    4. 4
      Blanchet, V.; Zgierski, M. Z.; Seideman, T.; Stolow, A. Discerning vibronic molecular dynamics using time-resolved photoelectron spectroscopy. Nature 1999, 401, 5254,  DOI: 10.1038/43410
    5. 5
      Polli, D.; Altoe, P.; Weingart, O.; Spillane, K. M.; Manzoni, C.; Brida, D.; Tomasello, G.; Orlandi, G.; Kukura, P.; Mathies, R. A. Conical intersection dynamics of the primary photoisomerization event in vision. Nature 2010, 467, 440443,  DOI: 10.1038/nature09346
    6. 6
      Schuurman, M. S.; Stolow, A. Dynamics at conical intersections. Annu. Rev. Phys. Chem. 2018, 69, 427450,  DOI: 10.1146/annurev-physchem-052516-050721
    7. 7
      Timmers, H.; Zhu, X.; Li, Z.; Kobayashi, Y.; Sabbar, M.; Hollstein, M.; Reduzzi, M.; Martínez, T. J.; Neumark, D. M.; Leone, S. R. Disentangling conical intersection and coherent molecular dynamics in methyl bromide with attosecond transient absorption spectroscopy. Nat. Commun. 2019, 10, 3133,  DOI: 10.1038/s41467-019-10789-7
    8. 8
      Garavelli, M. Computational organic photochemistry: Strategy, achievements and perspectives. Theor. Chem. Acc. 2006, 116, 87105,  DOI: 10.1007/s00214-005-0030-z
    9. 9
      Robb, M. A. Theoretical chemistry for electronic excited states; Theoretical and Computational Chemistry Series; The Royal Society of Chemistry, 2018; pp P001225.
    10. 10
      Lasorne, B.; Worth, G. A.; Robb, M. A. Excited-state dynamics. WIREs Comput. Mol. Sci. 2011, 1, 460475,  DOI: 10.1002/wcms.26
    11. 11
      Mai, S.; Marquetand, P.; González, L. Nonadiabatic dynamics: The SHARC approach. WIREs Comput. Mol. Sci. 2018, 8, e1370,  DOI: 10.1002/wcms.1370
    12. 12
      Chergui, M. On the interplay between charge, spin and structural dynamics in transition metal complexes. Dalton Trans 2012, 41, 1302213029,  DOI: 10.1039/c2dt30764b
    13. 13
      Atkins, A. J.; González, L. Trajectory surface-hopping dynamics including intersystem crossing in [Ru(bpy)3]2+. J. Phys. Chem. Lett. 2017, 8, 38403845,  DOI: 10.1021/acs.jpclett.7b01479
    14. 14
      Nogueira, J. J.; Plasser, F.; González, L. Electronic delocalization, charge transfer and hypochromism in the UV absorption spectrum of polyadenine unravelled by multiscale computations and quantitative wavefunction analysis. Chem. Sci. 2017, 8, 56825691,  DOI: 10.1039/C7SC01600J
    15. 15
      Koch, M.; Thaler, B.; Heim, P.; Ernst, W. E. The role of Rydberg-valence coupling in the ultrafast relaxation dynamics of acetone. J. Phys. Chem. A 2017, 121, 63986404,  DOI: 10.1021/acs.jpca.7b05012
    16. 16
      Nobre, M.; Fernandes, A.; da Silva, F. F.; Antunes, R.; Almeida, D.; Kokhan, V.; Hoffmann, S. V.; Mason, N.; Eden, S.; Limão-Vieira, P. The VUV electronic spectroscopy of acetone studied by synchrotron radiation. Phys. Chem. Chem. Phys. 2008, 10, 550560,  DOI: 10.1039/B708580J
    17. 17
      Holzapfel, W.; Finkele, U.; Kaiser, W.; Oesterhelt, D.; Scheer, H.; Stilz, H. U.; Zinth, W. Initial electron-transfer in the reaction center from Rhodobacter sphaeroides. Proc. Natl. Acad. Sci. U. S. A. 1990, 87, 51685172,  DOI: 10.1073/pnas.87.13.5168
    18. 18
      van Stokkum, I. H.; Larsen, D. S.; van Grondelle, R. Global and target analysis of time-resolved spectra. Biochim. Biophys. Acta, Bioenerg. 2004, 1657, 82104,  DOI: 10.1016/j.bbabio.2004.04.011
    19. 19
      Wu, G.; Boguslavskiy, A. E.; Schalk, O.; Schuurman, M. S.; Stolow, A. Ultrafast non-adiabatic dynamics of methyl substituted ethylenes: The π3s Rydberg state. J. Chem. Phys. 2011, 135, 164309,  DOI: 10.1063/1.3652966
    20. 20
      Schalk, O.; Boguslavskiy, A. E.; Stolow, A. Substituent effects on dynamics at conical intersections: Cyclopentadienes. J. Phys. Chem. A 2010, 114, 40584064,  DOI: 10.1021/jp911286s
    21. 21
      Stolow, A.; Bragg, A. E.; Neumark, D. M. Femtosecond time-resolved photoelectron spectroscopy. Chem. Rev. 2004, 104, 17191758,  DOI: 10.1021/cr020683w
    22. 22
      Hertel, I. V.; Radloff, W. Ultrafast dynamics in isolated molecules and molecular clusters. Rep. Prog. Phys. 2006, 69, 1897,  DOI: 10.1088/0034-4885/69/6/R06
    23. 23
      Kim, S. K.; Pedersen, S.; Zewail, A. H. Direct femtosecond observation of the transient intermediate in the α-cleavage reaction of (CH3)2CO to 2CH3+CO: Resolving the issue of concertedness. J. Chem. Phys. 1995, 103, 477,  DOI: 10.1063/1.469614
    24. 24
      Shibata, T.; Suzuki, T. Photofragment ion imaging with femtosecond laser pulses. Chem. Phys. Lett. 1996, 262, 115119,  DOI: 10.1016/0009-2614(96)01024-X
    25. 25
      Buzza, S. A.; Snyder, E. M.; Castleman, A. W. Further direct evidence for stepwise dissociation of acetone and acetone clusters. J. Chem. Phys. 1996, 104, 5040,  DOI: 10.1063/1.471133
    26. 26
      Rusteika, N.; Møller, K. B.; Sølling, T. I. New insights on the photodynamics of acetone excited with 253–288nm femtosecond pulses. Chem. Phys. Lett. 2008, 461, 193197,  DOI: 10.1016/j.cplett.2008.06.079
    27. 27
      Hüter, O.; Temps, F. Ultrafast α–CC bond cleavage of acetone upon excitation to 3p and 3d Rydberg states by femtosecond time-resolved photoelectron imaging. J. Chem. Phys. 2016, 145, 214312,  DOI: 10.1063/1.4971243
    28. 28
      Maierhofer, P.; Bainschab, M.; Thaler, B.; Heim, P.; Ernst, W. E.; Koch, M. Disentangling multichannel photodissociation dynamics in acetone by time-resolved photoelectron-photoion coincidence spectroscopy. J. Phys. Chem. A 2016, 120, 64186423,  DOI: 10.1021/acs.jpca.6b07238
    29. 29
      Couch, D. E.; Kapteyn, H. C.; Murnane, M. M.; Peters, W. K. Uncovering highly-excited state mixing in acetone using ultrafast VUV pulses and coincidence imaging techniques. J. Phys. Chem. A 2017, 121, 23612366,  DOI: 10.1021/acs.jpca.7b01112
    30. 30
      Uenishi, R.; Horio, T.; Suzuki, T. Time-resolved photoelectron imaging of acetone with 9.3 eV photoexcitation. J. Phys. Chem. A 2019, 123, 68486853,  DOI: 10.1021/acs.jpca.9b05179
    31. 31
      Koch, M.; Heim, P.; Thaler, B.; Kitzler, M.; Ernst, W. E. Direct observation of a photochemical activation energy: a case study of acetone photodissociation. J. Phys. B: At., Mol. Opt. Phys. 2017, 50, 125102,  DOI: 10.1088/1361-6455/aa6a71
    32. 32
      Sølling, T. I.; Diau, E. W.-G.; Kötting, C.; De Feyter, S.; Zewail, A. H. Femtochemistry of Norrish Type I reactions: IV. Highly excited ketones—Experimental. ChemPhysChem 2002, 3, 7997,  DOI: 10.1002/1439-7641(20020118)3:1<79::AID-CPHC79>3.0.CO;2-#
    33. 33
      Thakur, S. N.; Guo, D.; Kundu, T.; Goodman, L. Two-photon photoacoustic spectroscopy of acetone 3p Rydberg states. Chem. Phys. Lett. 1992, 199, 335340,  DOI: 10.1016/0009-2614(92)80128-X
    34. 34
      Xing, X.; McDiarmid, R.; Philis, J. G.; Goodman, L. Vibrational assignments in the 3p Rydberg states of acetone. J. Chem. Phys. 1993, 99, 75657573,  DOI: 10.1063/1.465686
    35. 35
      Merchán, M.; Roos, B. O.; McDiarmid, R.; Xing, X. A combined theoretical and experimental determination of the electronic spectrum of acetone. J. Chem. Phys. 1996, 104, 1791,  DOI: 10.1063/1.470976
    36. 36
      ter Steege, D. H. A.; Wirtz, A. C.; Buma, W. J. Vibronic coupling in excited states of acetone. J. Chem. Phys. 2002, 116, 547560,  DOI: 10.1063/1.1423946
    37. 37
      McDiarmid, R.; Xing, X. Nonadiabatic coupling of the 3p Rydberg and ππ* valence states of acetone. J. Chem. Phys. 1997, 107, 675679,  DOI: 10.1063/1.475152
    38. 38
      Mai, S.; Richter, M.; Heindl, M.; Menger, M. F. S. J.; Atkins, A. J.; Ruckenbauer, M.; Plasser, F.; Oppel, M.; Marquetand, P.; González, L. SHARC2.0: Surface Hopping Including Arbitrary Couplings – program package for non-adiabatic dynamics; sharc-md.org, 2018.
    39. 39
      Plasser, F.; Gómez, S.; Mai, S.; González, L. Highly efficient surface hopping dynamics using a linear vibronic coupling model. Phys. Chem. Chem. Phys. 2019, 21, 5769,  DOI: 10.1039/C8CP05662E
    40. 40
      Subotnik, J. E.; Jain, A.; Landry, B.; Petit, A.; Ouyang, W.; Bellonzi, N. Understanding the surface hopping view of electronic transitions and decoherence. Annu. Rev. Phys. Chem. 2016, 67, 387417,  DOI: 10.1146/annurev-physchem-040215-112245
    41. 41
      Wang, L.; Akimov, A.; Prezhdo, O. V. Recent progress in surface hopping: 2011–2015. J. Phys. Chem. Lett. 2016, 7, 21002112,  DOI: 10.1021/acs.jpclett.6b00710
    42. 42
      Plasser, F.; Mai, S.; Fumanal, M.; Gindensperger, E.; Daniel, C.; González, L. Strong influence of decoherence corrections and momentum rescaling in surface hopping dynamics of transition metal complexes. J. Chem. Theory Comput. 2019, 15, 50315045,  DOI: 10.1021/acs.jctc.9b00525
    43. 43
      Kruit, P.; Read, F. Magnetic field paralleliser for 2π electron-spectrometer and electron-image magnifier. J. Phys. E: Sci. Instrum. 1983, 16, 313,  DOI: 10.1088/0022-3735/16/4/016
    44. 44
      Fumanal, M.; Plasser, F.; Mai, S.; Daniel, C.; Gindensperger, E. Interstate vibronic coupling constants between electronic excited states for complex molecules. J. Chem. Phys. 2018, 148, 124119,  DOI: 10.1063/1.5022760
    45. 45
      Dreuw, A.; Wormit, M. The algebraic diagrammatic construction scheme for the polarization propagator for the calculation of excited states. WIREs Comput. Mol. Sci. 2015, 5, 8295,  DOI: 10.1002/wcms.1206
    46. 46
      Dunning, T. H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 10071023,  DOI: 10.1063/1.456153
    47. 47
      Kaufmann, K.; Baumeister, W.; Jungen, M. Universal Gaussian basis sets for an optimum representation of Rydberg and continuum wavefunctions. J. Phys. B: At., Mol. Opt. Phys. 1989, 22, 2223,  DOI: 10.1088/0953-4075/22/14/007
    48. 48
      TURBOMOLE V7.0, A development of University of Karlsruhe and Forschungszentrum Karlsruhe GmbH, 2015.
    49. 49
      Granucci, G.; Persico, M.; Toniolo, A. Direct semiclassical simulation of photochemical processes with semiempirical wave functions. J. Chem. Phys. 2001, 114, 1060810615,  DOI: 10.1063/1.1376633
    50. 50
      Granucci, G.; Persico, M. Critical appraisal of the fewest switches algorithm for surface hopping. J. Chem. Phys. 2007, 126, 134114,  DOI: 10.1063/1.2715585
  • Supporting Information

    Supporting Information


    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jpclett.9b03462.

    • Discussion and derivation of the fitting functions, experimental details regarding energy calibration, additional experimental results, additional computational details, and computational results (PDF)

    • V0 data in SHARC format (TXT)

    • LVC parameters in SHARC format (TXT)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.