Using the recently developed multistate mapping approach to surface hopping (multistate MASH) method combined with SA(3)-CASSCF(12,12)/aug-cc-pVDZ electronic structure calculations, the gas-phase isotropic ultrafast electron diffraction (UED) of cyclobutanone is predicted and analyzed. After excitation into the n-3s Rydberg state (S2), cyclobutanone can relax through two S2/S1 conical intersections, one characterized by compression of the CO bond and the other by dissociation of the α–CC bond. Subsequent transfer into the ground state (S0) is then achieved via two additional S1/S0 conical intersections that lead to three reaction pathways: α ring-opening, ethene/ketene production, and CO liberation. The isotropic gas-phase UED signal is predicted from the multistate MASH simulations, allowing for a direct comparison to the experimental data. This work, which is a contribution to the cyclobutanone prediction challenge, facilitates the identification of the main photoproducts in the UED signal and thereby emphasizes the importance of dynamics simulations for the interpretation of ultrafast experiments.

The photochemistry of cyclic ketones has been studied extensively.1–6 Despite their small size, they have a rich photochemistry with multiple competing pathways that include dissociation, fluorescence, and intersystem crossing.7,8 The relative importance of the different pathways is closely linked to the size of the organic ring and the associated ring strain.7 The photodissociation of cyclic ketones has been of particular interest ever since the pioneering work of Norrish.9 

In this class of molecules, a target of recurring interest is cyclobutanone, (CH2)3CO.1,2,7,10–12 Early experimental studies identified the photoproducts of cyclobutanone as propylene (cyclopropane) and carbon monoxide (C3H6 + CO) or ethylene and ketene (CH2CH2 + CH2CO). The corresponding quantum yields were found to be 40% and 60%,1 although these were later shown to have a strong dependence on the excitation wavelength.8 Recent experimental work includes ultrafast transient absorption spectroscopy on cyclobutanone in solution.7 Upon excitation to the S1 (nπ*) state by UV pulses in the range 255–312 nm, Kao et al. found that singlet dissociation pathways are dominant at shorter wavelengths, predominantly via Norrish Type-1 α cleavage, with minor contributions from ketene formation. At higher excitation energies, such as 200 nm, the n-3s Rydberg state comes into play.13–15 In this regime, using time-resolved mass spectrometry and photoelectron spectroscopy, Kuhlman et al. found that a ring puckering mode effectively couples the S2 and S1 states, allowing for rapid internal conversion.16 

The experimental studies have been accompanied by theoretical work. Electronic structure calculations using complete active space self-consistent field, CASSCF(10,8), and multistate complete active space second-order perturbation theory, MS-CASPT2(10,8), predicted a small barrier for ring-opening on the S1 state, with the possibility of minor contributions from barrierless dissociation in the T1 state.17 Building on this, simulations of cyclobutanone using the ab initio multiple spawning (AIMS) method found a ring-opening mechanism to be dominant with a small portion of trajectories yielding CH2=CH2 + CH2CO and CH2CH2 + CH2=CO fragmentation.18 We also note that Kuhlman et al. constructed a five-dimensional linear-vibronic Hamiltonian (LVH) model to simulate the S2/S1 decay in cyclobutanone using the multiconfigurational time-dependent Hartree (MCTDH) method, which succeeded in replicating the short time constants of 0.95 ps.19 

This paper is motivated by the prediction challenge based on an ultrafast electron diffraction (UED) experiment carried out at the SLAC MeV-UED facility.20 UED has emerged as a powerful experimental technique for observing molecular dynamics in the recent decade,21–25 alongside ultrafast x-ray scattering.26–29 Notably, both techniques have made advances in extracting information that extends beyond structural dynamics.30,31 We use this challenge as an opportunity to benchmark the electronic structure calculations for cyclobutanone. This comparison includes non-standard selected-configuration interaction (SCI) electronic structure methods that overcome the issue of large active space selection, constitute a black-box alternative to complete active space (CAS) methods, and provide complementary benchmarks.32,33 The nonadiabatic simulations are carried out using the newly developed multistate mapping approach to surface hopping (multistate MASH) method.34 Furthermore, we use the simulations to predict the UED signals and analyze these to determine the main contributions to the scattering.

Many computational methods have been developed to simulate nonadiabatic dynamics over the past decades. Some, such as MCTDH35,36 and multi-layer MCTDH,37 can provide numerically exact results. However, these methods require high-accuracy precomputed potential energy surfaces (PESs), which, combined with the exponential scaling of quantum mechanics, imposes severe limitations on the number of degrees of freedom that can be treated. To circumvent this problem, the nuclear wavefunction can be represented as a linear combination of traveling Gaussian basis functions or classically guided trajectories, and the required electronic structure quantities can be calculated on-the-fly, also known as direct dynamics. Such approaches can significantly reduce the phase space explored and, therefore, the computational cost per degree of freedom. On-the-fly methods include examples where the nuclei are propagated using equations of motion derived from the variational principle, notably variational multiconfigurational Gaussian (vMCG),38–40 or semiclassically, which includes methods such as full or ab initio multiple spawning (FMS/AIMS)41–44 and multiconfigurational Ehrenfest (MCE),45–47 or classically, such as trajectory surface hopping (TSH).48 In addition to the above-mentioned methods, mapping methods treat the electronic and nuclear degrees of freedom in an equal manner. Examples include the Meyer–Miller–Stock–Thoss (MMST) mapping Hamiltonian, symmetrical quasi-classical windowing, and generalized spin mapping.49–52 Ideas from this last category informed the development of the surface hopping variant used in this paper (see Sec. II A 2 below). For an in-depth overview of different methods for nonadiabatic dynamics, the reader is directed toward the edited volume in Ref. 53.

1. Trajectory surface hopping

For nonadiabatic molecular dynamics, TSH is perhaps the most commonly employed approach to support the interpretation of experiments.54–61 In TSH, an ensemble of individual trajectories is used to represent the propagation of the nuclear wavepacket. In each trajectory, the nuclei are treated classically and the electronic amplitudes are used to predict hops to other electronic states. For a full discussion of TSH, please see Ref. 62, while we continue by highlighting the key aspects of the fewest switches surface hopping (FSSH) variant of TSH.

Each trajectory in FSSH is initialized by projecting a ground-state vibrational distribution into the excited state of interest. We consider pure excitation, i.e., ca = 1, where ca is the electronic wavefunction coefficient for the active state a. The nuclear coordinates, R, are propagated using classical equations of motion,
(1)
for each atom α, with t being the time, Mα is the mass of atom α, and ∇αEa is the gradient (force) on the active adiabatic electronic state. The integration is typically carried out using a standard velocity-Verlet scheme, although other propagators may be able to achieve better convergence.63 The forces are normally obtained from electronic structure calculations of adiabatic states. In parallel with the nuclear propagation, one must integrate the electronic coefficients, c = (c1, …, cN), for the N electronic states considered. In the adiabatic representation, the probability PabFSSH of a hop from state a to state b is evaluated according to
(2)
where dab is the nonadiabatic coupling vector between states a and b, v is the classical nuclear velocity, ca and cb are the electronic coefficients for states a and b, respectively, and Δt is the time step.64 In some cases, an overlap scheme is employed to avoid the expensive calculations of nonadiabatic couplings.65,66 At each time step, PabFSSH is evaluated and compared to a random number, resulting in either acceptance or rejection of the hop. If a hop is accepted but the energy in the system is insufficient for it to occur, the hop is rejected and the nuclear velocities are reflected. This scenario is referred to as a frustrated hop. For hops that do take place, the active state is updated (ab), the velocities are modified to conserve total energy, and the trajectories continue to be propagated using the forces on the new active state.

The algorithm presented above is robust but also causes one of the most well-known shortcomings of TSH, the so-called overcoherence issue. A number of, more or less ad hoc, decoherence correction schemes have been developed in response.67,68 The method presented next aims to alleviate the need for such ad hoc corrections.

2. Multistate mapping approach to surface hopping

Recently, a new version of TSH was devised by Mannouch and Richardson, known as “mapping approach to surface hopping” (MASH).69 MASH combines elements of mapping methods with TSH to attempt to get the advantages of both. In contrast to the stochastic algorithm in FSSH, MASH uses a deterministic algorithm to evaluate the active state,
(3)
In other words, MASH defines the active state to be the state with the largest value of |ci|2, alleviating the need for ad hoc decoherence corrections and improving accuracy when compared to the original FSSH procedure.69 Originally, MASH was formulated for two-state systems, but MASH was then adapted to any number of states by Runeson and Manolopoulos, giving rise to what we hereon refer to as multistate MASH.34,70

To start trajectories, multistate MASH determines the nuclear initial conditions in the same manner as TSH, mapping the ground-state wavefunction onto a classical phase space. However, as in other mapping approaches, MASH also considers the electronic coefficients c as a phase space variable. Therefore, rather than describing the initial state as pure, such as done in FSSH, multistate MASH represents the initial state as a distribution over c = (c1, …, cN). In practice, this is done by random assignment of the complex-valued coefficients, ci = xi + iyi, for each electronic state from Gaussian distributions with zero mean and standard deviation one for both the real and imaginary components. The resulting set of coefficients is normalized such that 1=i=1N|ci|2. Coefficients for which the initial state of interest does not have the largest absolute value are resampled until suitable sets are found. It is worth noting that other choices of initial distributions are possible, but in most cases, have been found to yield similar results.34 Therefore, we use this method of selecting initial conditions due to its ease of implementation.

To measure populations, several different estimators have been proposed in the literature; more information can be found in the supplementary material. In this paper, we measure the active state, corresponding to an estimator of the form,
(4)
With this choice of estimator and the initial distribution described above, the ensemble-averaged population transfer from state i to state n is
(5)

In the two-state case, Mannouch and Richardson have shown that this prescription overestimates the rate of transfer compared to the Landau–Zener theory. It does, however, maintain a physical (non-negative) population estimate with a low number of trajectories. Note that this approach differs from what was previously used in both MASH and multistate MASH. Further information on alternative population estimators is given in the supplementary material.

One potential side-effect of multistate MASH is hopping between states that are not coupled. It is possible to nevertheless accept all hops, regardless of the coupling between the states, which was the approach taken in Ref. 34. However, due to the risk of unphysical behavior, we introduce an energy criterion for all hops, such that |EaEb| < 0.055 Hartree (≈1.5 eV) to prevent erroneous hops between uncoupled states. A value of ≈1.5 eV was selected based on previous dynamics with TSH; however, further testing of this value is still required. Hops that do not meet this criterion are treated on the same footing as reflected hops, with the total velocity of all atoms reflected. Note that this direction of rescaling differs from the previous work with MASH and is chosen due to simplicity.

We simulate electron diffraction using the independent atom model (IAM), originally devised by Debye.71–74 This method, extensively used in the analysis of scattering and crystallography experiments, approximates the electron scattering probability, Z(s,R)2, as a coherent sum of tabulated form factors,75  fNe(s),
(6)
where s is the momentum transfer vector, R the molecular geometry, RAB the distance vector between atoms A and B, and Nat the number of atoms in the molecule. s = |s| is the norm of the momentum transfer vector. Note that Z(s,R)2 is given in units of the Rutherford cross section (e.g., see Ref. 74 for further details). The electron scattering atomic form factors, fAe(s), are defined as
(7)
where fAx(s) are the tabulated form factors for x-rays and ZA the nuclear charges.
We calculate the rotationally averaged signal accounting for all possible molecular orientations in a thermal ensemble. This yields the isotropic elastic electron scattering probability, S0(s,R), which takes the form,
(8)
where RAB = |RAB| is the distance between atoms A and B, and where we have introduced the zeroth-order spherical Bessel function j0(sRAB) defined as
(9)
In the experiment, not only the elastic component of electron scattering is measured but also the inelastic component. While the IAM cannot describe individual inelastic transitions, the total inelastic component can be approximated by an incoherent sum of atomic inelastic scattering functions SA(s). This yields the total isotropic scattering, St(s,R), as the sum of the elastic and inelastic contributions,
(10)

We note that further improvements in the scattering signal are straightforward. For instance, the alignment effects resulting from the linear polarization of the pump laser can be accounted for Ref. 76 and corrections to the form factors due to relativistic effects can be made;77 however, for the latter, we note that this effect is comparatively minor for the electron energies at the SLAC MeV-UED source. It is also possible to calculate the scattering cross sections from ab initio electronic wavefunctions, which makes it possible to accurately account for the effects on scattering from chemical bonding, electron correlation, and change in the electronic state due to excitation by the pump pulse.78–80 

A ground-state minimum energy structure of cyclobutanone was obtained using state-averaged CASSCF,81 SA(3)-CASSCF(12,12)/aug-cc-pVDZ, i.e., state-averaging over the three lowest singlet states, an active space with 12 electrons in 12 orbitals with the aug-cc-pVDZ Cartesian basis set.82 Further details on the active space are given in the supplementary material. All electronic structure calculations were performed in OpenMolcas version v23.02.83,84 The ground-state minimum energy structure was subjected to a frequency calculation yielding no imaginary frequencies, confirming the structure to be a true minimum. Further excited state structures were optimized using SA(3)-CASSCF(12,12) to obtain the S2 minimum, two S2/S1 minimum energy conical intersections (MECIs), and two S1/S0 MECIs. The selected structures were subjected to linear interpolation in internal coordinates (LIIC) to yield a progression of intermediate structures. Potential energy curves along the LIICs were calculated using SA(3)-CASSCF(12,12) and extended multistate complete active space second-order perturbation theory, XMS-CASPT2, with the same (12,12) active space as for the CASSCF calculations (see the supplementary material).85 For stability, an imaginary shift of 0.5 a.u. was applied for all XMS-CASPT2 calculations.86 

For further comparison, the potential energy curves were also calculated using a modified version of adaptive configuration interaction (ACI) and Monte Carlo configuration interaction (MCCI).87–89 The state-averaged MCCI calculations were performed using MCCI V489,90 with the SA error-controlled calculations performed using a locally modified development version of GeneralSCI 1.0.91 Further details are given in the supplementary material.

The calculated ground-state frequencies were used to generate a Wigner distribution from which 1000 geometries were sampled. At each geometry, excitation energies and oscillator strengths were calculated using SA(3)-CASSCF(12,12). The photoabsorption cross section was constructed using the nuclear ensemble approach (NEA), convoluted by Lorentzian functions with a phenomenological broadening parameter of 0.05 eV.92 Initial conditions for the multistate MASH trajectories were sampled from this Wigner distribution using an implicit laser pulse of width 0.2 eV, centered at 6.2 eV. All trajectories were initiated from the bright S2 (n-3s Rydberg) state using multistate MASH (see Sec. II A 2 for more details)34 implemented in a developmental version of SHARC.93 The electronic structure calculations were carried out using SA(3)-CASSCF(12,12)/aug-cc-pVDZ with all three electronic states available in the dynamics. The trajectories were propagated using a 0.5 fs time step, with 25 substeps, for 300 fs using a local diabatization scheme. Previous applications of multistate MASH have employed time steps in the region of 0.25 fs, and due to the similarities of the algorithm to TSH, similar time steps are expected to be adequate.34 For trajectories in which the total energy changes by more than 0.25 eV, a value selected based on previous studies,57,94–96 the propagation is stopped, and only time points with the correct total energy are used in the analysis. In addition, we note that a subset of trajectories becomes “trapped.” This occurs when the trajectories enter a region where a reflected hop occurs but the reflected hop fails to remove the trajectory from the “forbidden” region. Such trapped trajectories are only included in the analysis until the point that they get trapped. For a summary of the trajectories removed from the analysis, see the supplementary material. Populations of the ensemble of trajectories are estimated according to Sec. II A 2. Finally, rotationally averaged total scattering patterns were calculated using the geometries from the multistate MASH trajectories and the independent atom model (see Sec. II B) for the initial 200 fs of the dynamics.

The character of the two lowest-lying S1 and S2 electronic states of cyclobutanone has previously been identified as nπ* and n-3s, respectively.13,16 The next set of excited states are the higher-lying 3p Rydberg states, which appear a full 0.7 eV above the n-3s Rydberg state.14 Triplet states have been shown to be important for the photodynamics upon excitation with long wavelengths.7 However, due to the short excited state lifetime of cyclobutanone following excitation into the S2 state, triplet states are less likely to make a significant contribution to the current photodynamics. For these reasons, we have chosen to focus our attention on an accurate description of the two lowest-lying singlet states S1 and S2.

A summary of the low-lying excited singlet states available to cyclobutanone at the equilibrium geometry is presented in Table I. Both SA(3)-CASSCF(12,12) and XMS-CASPT2(12,12) yield an S1 state with nπ* character at the Franck–Condon (FC) geometry and an S2 state with the n-3s Rydberg character. Furthermore, the excitation energies for SA(3)-CASSCF(12,12) are in excellent agreement with XMS-CASPT2(12,12), with SA(3)-CASSCF(12,12) energies showing a constant difference of 0.04 eV for both the excited states. Excitation energies for SA(3)-ACI and SA(3)-MCCI are also presented in Table I. These methods give a good description of the low-lying singlet states of cyclobutanone with excitation energies deviating by ∼0.05 eV for the valence nπ* state and ∼0.1 eV for the n-3s Rydberg state compared to the XMS-CASPT2(12,12) benchmark. In the ACI simulation, a large orbital window, corresponding to CAS-CI(12,36), has also been tested. The large orbital window gives results analogous to the ones obtained with CASSCF(12,12), providing further support for this choice of active space.

TABLE I.

Table of excitation energies in eV for the lowest two excited electronic states of cyclobutanone at the S0 minimum energy geometry optimized using SA(3)-CASSCF(12,12)/aug-cc-pVDZ. The state characters are given in parentheses along with excitation energies.

MethodS1S2
SA(3)-CASSCF(12,12) 4.44 (nπ*) 6.24 (n-3s) 
XMS-CASPT2(12,12) 4.40 (nπ*) 6.20 (n-3s) 
SA(3)-ACI 4.43 (nπ*) 6.31 (n-3s) 
SA(3)-MCCI 4.45 (nπ*) 6.30 (n-3s) 
MethodS1S2
SA(3)-CASSCF(12,12) 4.44 (nπ*) 6.24 (n-3s) 
XMS-CASPT2(12,12) 4.40 (nπ*) 6.20 (n-3s) 
SA(3)-ACI 4.43 (nπ*) 6.31 (n-3s) 
SA(3)-MCCI 4.45 (nπ*) 6.30 (n-3s) 

A theoretical photoabsorption cross section of cyclobutanone calculated using SA(3)-CASSCF(12,12) is shown in Fig. 1, allowing for a direct comparison with an experimental absorption spectrum. Good agreement can be observed between the two spectra, further validating our electronic structure method along with confirming the assignment of states using the insets shown in Fig. 1.

FIG. 1.

Photoabsorption cross section of gas-phase cyclobutanone calculated using the SA(3)-CASSCF(12,12)/aug-cc-pVDZ, the NEA, and Wigner sampling of 1000 geometries (purple line). The experimental data are reproduced from Ref. 97 (blue line). Orbitals for the S1 and S2 states are shown as an inset.

FIG. 1.

Photoabsorption cross section of gas-phase cyclobutanone calculated using the SA(3)-CASSCF(12,12)/aug-cc-pVDZ, the NEA, and Wigner sampling of 1000 geometries (purple line). The experimental data are reproduced from Ref. 97 (blue line). Orbitals for the S1 and S2 states are shown as an inset.

Close modal

After the photophysics of cyclobutanone has been studied in the FC region, the different deactivation pathways available upon photoexcitation can be characterized. A series of potentially important optimized geometries for these pathways are shown in Fig. 2, optimized using SA(3)-CASSCF(12,12)/aug-cc-pVDZ. The S0 minimum energy geometry (S0, upper left) can be characterized by a single carbon atom breaking the planar symmetry of the molecule, along with a C=O bond length of 1.19 Å and an α–CC bond of 1.58 Å. When the molecule is excited by a 200 nm pulse, the S2 excited electronic state is populated at the FC region and can potentially relax into the S2 minimum. The S2 minimum energy geometry (S2, upper center) displays subtle changes in bond lengths, along with a planarization of the carbon ring. Two S2/S1 MECIs were successfully optimized, with the S2/S1 MECI (S2/S1, lower left) already reported previously.19 This MECI displays a rather severe compression of the C=O bond. The second S2/S1 MECI (S2/S1, upper right) exhibits a dissociated α–CC bond alongside a compressed C=O bond. Two further MECIs are shown in Fig. 2, corresponding to MECIs between the ground and first excited state. Both these S1/S0 MECIs have been reported previously by Liu et al.18 The first of these MECIs (S1/S0, lower center), termed CI-3 in Ref. 18, breaks the ring structure to produce CH2=CH2 and O=CCH2. The second S1/S0 MECI (S1/S0, lower right), termed CI-1 in Ref. 18, is closely related to the second S2/S1 MECI (S2/S1, upper right), with a dissociated α–CC bond. However, in contrast to the S2/S1 MECI, the S1/S0 MECI contains a slightly stretched C=O bond relative to the S0 geometry. It is obvious from this static perspective of excited state geometries, that cyclobutanone possesses a rich photochemistry. However, the relative importance of each of these structures is still unknown, as are the final products of each conical intersection (CI).

FIG. 2.

Six critical geometries are shown for gas-phase cyclobutanone. The S0 minimum (upper left), S2 minimum (upper center), a dissociative S2/S1 MECI (upper right), another S2/S1 MECI (lower left), a ring-breaking S1/S0 MECI (lower center), and a ring-opening S1/S0 (lower right). All the structures shown are optimized using SA(3)-CASSCF(12,12)/aug-cc-pVDZ.

FIG. 2.

Six critical geometries are shown for gas-phase cyclobutanone. The S0 minimum (upper left), S2 minimum (upper center), a dissociative S2/S1 MECI (upper right), another S2/S1 MECI (lower left), a ring-breaking S1/S0 MECI (lower center), and a ring-opening S1/S0 (lower right). All the structures shown are optimized using SA(3)-CASSCF(12,12)/aug-cc-pVDZ.

Close modal

To gain a better understanding of the PESs of cyclobutanone and to benchmark SA(3)-CASSCF(12,12) away from the FC region, an LIIC was performed on some of the key geometries shown in Fig. 2. Four structures were selected: the S0 minimum (Fig. 2, S0, upper left), the S2 minimum (Fig. 2, S2, upper center), the dissociative S2/S1 MECI (Fig. 2, S2/S1, upper right), and finally, the α–CC bond dissociation S1/S0 MECI (Fig. 2, S1/S0, lower right). Here, the two CIs were selected due to being the most commonly encountered CIs during the dynamics presented in Sec. IV C.

Upon photoexcitation to the S2 state, cyclobutanone in the S0 minimum structure can relax to the closeby S2 minimum via planarization of the carbon ring (the green line shown in panel 1 of Fig. 3). Here, cyclobutanone undergoes a dissociation of an α–CC bond, resulting in a barrier height of 0.21 eV for SA(3)-CASSCF(12,12), somewhat below the 0.49 eV predicted by XMS-CASPT(12,12), with the barrier heights given as energy differences relative to the S2 minimum. Note that the absolute values for transition state barriers in LIICs should be treated with care as they are known to be overestimated. Regardless, as a result of this smaller reaction barrier, SA(3)-CASSCF(12,12) is likely to estimate a faster rate of decay from S2 to S1. A steep drop in energy is observed for both the methods after the transition barrier leading down to the S2/S1 MECI (see Fig. 2, upper right and inset of Fig. 3). This S2/S1 MECI is peaked and has a single pathway according to the criteria set out in Ref. 98. This allows for efficient funneling onto the S1 state (the blue line shown in Fig. 3) where a barrier-less decay is observed to the structurally similar S1/S0 MECI (see Fig. 2) and where the S1/S0 MECI is observed with a peaked bifurcating branching space.

FIG. 3.

Linear interpolation in internal coordinates from the S0 minimum to S2 minimum to the dissociative S2/S1 MECI, ending with the ring opening S1/S0 MECI. The inset structures correspond to the optimized structures used along the LIIC coordinate. The excitation energies for the three lowest-lying singlet states are given using both SA(3)-CASSSCF(12,12)/aug-cc-pVDZ (dotted line) and XMS-CASPT2(12,12)/aug-cc-pVDZ (solid line).

FIG. 3.

Linear interpolation in internal coordinates from the S0 minimum to S2 minimum to the dissociative S2/S1 MECI, ending with the ring opening S1/S0 MECI. The inset structures correspond to the optimized structures used along the LIIC coordinate. The excitation energies for the three lowest-lying singlet states are given using both SA(3)-CASSSCF(12,12)/aug-cc-pVDZ (dotted line) and XMS-CASPT2(12,12)/aug-cc-pVDZ (solid line).

Close modal

It can be deduced from Fig. 3 that there is excellent agreement between SA(3)-CASSCF(12,12) and XMS-CASPT2(12,12), with slight differences observed for the approximate transition barrier height. Nevertheless, overall the SA(3)-CASSCF(12,12) results offer an excellent comprise between speed and accuracy for the on-the-fly nonadiabatic dynamics of gas-phase cyclobutanone. Additional LIICs for the same geometries are given in the supplementary material (Figs. 3 and 4) for both SA(3)-ACI and SA(3)-MCCI. These further support our assessment. We also note that this indicates that good agreement can be obtained with methods that potentially offer a more black-box approach to the calculation of multireference wavefunctions than the CAS family of methods.

FIG. 4.

Adiabatic populations for the three lowest-lying singlet states, S0–2, given by the population estimator in Eq. (5) as a function of time.

FIG. 4.

Adiabatic populations for the three lowest-lying singlet states, S0–2, given by the population estimator in Eq. (5) as a function of time.

Close modal

A total of 229 trajectories were initiated on the S2 state. Trajectories were removed from the statistics if either a change of 0.25 eV or more was observed in the total energy or if consecutive reflected hops indicated that the trajectory was trapped in a “forbidden” region of the PES. This resulted in a significant number of discarded trajectories. For a full summary of discarded trajectories, see the supplementary material (Fig. 2). To briefly summarize, after 50 fs, the number of trajectories had dropped to 146, and by 200 fs, only 18 trajectories of the initial 229 trajectories had survived, while at 300 fs, this was down to 10 trajectories. At 300 fs, a total of 71 trajectories had been discarded due to consecutive reflected hops, while 148 trajectories had been discarded due to issues with the energy conservation. We acknowledge that this could potentially cause artifacts in the simulations, especially if certain processes are systematically removed or enhanced by the removal of trajectories.

The adiabatic populations for the three electronic states S0–2 used in our simulations are shown in Fig. 4 for the initial 200 fs of the dynamics. A rapid decay of the S2 state (green line) can be observed after ∼8 fs. This decay is correlated with a population increase in S1 (teal line), indicating that the fast initial dynamics on S2 rapidly reaches an S2/S1 CI. For the early time dynamics (t < 30 fs), the ground state can be seen to be inaccessible from the S2 state (panel 1 shown in Fig. 3). After 30 fs, there is a steady increase in the ground-state S0 (purple line) population, showing passage through the S1/S0 CI and, therefore, formation of the photoproducts linked to each CI (Fig. 2). After 100 fs, we observe a decreased rate of population transfer from S2 to S1, resulting in depletion of S1 and an overall decrease in the population of the two excited states matched by a continued increase in the population of the ground state. After 175 fs, the population transfer is largely over, although errors in this region will be high due to the small number of trajectories. The populations shown in Fig. 4 indicate ultrafast non-radiative deactivation pathways for cyclobutanone, there is the likely possibility of further non-equilibrium ground-state chemistry as a result of very hot molecules in the ground state following the decay dynamics.

As shown previously, the deactivation pathways available to cyclobutanone imply the cleavage of one or two bonds in the ring structure. To identify these different pathways, the bond lengths for all multistate MASH trajectories are plotted, separated into α- and β–CC bonds according to the inset shown in Fig. 5. Four distinct outcomes are observed, as follows:

  • There is no bond breaking (purple), and the molecule remains in the FC region. These are characterized by short α and β–CC bond lengths (R < 2.2 Å). It is important to note that the trajectories included in this group do not progress over 90 fs.

  • There exists a single α–CC bond breaking (teal). This pathway gives rise to a ring-opened structure, i.e., CH2CH2CH2CO. This is indicated by an elongation of an α–CC bond, but due to the carbon backbone still being intact, the distance between the two carbon atoms is limited to R ∼ 4.5 Å. This pathway is a result of passing through the S1/S0 CI (lower right, Fig. 2).

  • Two α–CC bonds break to liberate CO, where there is a large increase in α–CC bonds to values above 4.5 Å (blue). Attempts to optimize an MECI using SA(3)-CASSCF(12,12) involving CO dissociation were unsuccessful; however, trajectories were observed with CO dissociation after having decayed to the ground state via the α–CC dissociation S1/S0 CI (lower right, Fig. 2). This means the CO dissociation takes place on the ground state rapidly after passing through a CI. It is worth noting that for option (iii), a single trajectory also displays ground state dissociation of CH2CH2CH2 into CH2CH2 and CH2, which is shown in Fig. 5 as the blue line displaying increases in both α and β–CC bonds.

  • An α–CC bond and β–CC bond breaking (pale green) showing the stepwise ring-breaking mechanism where ethene and ketene are produced. Here, an α–CC bond breaks first, followed by a β–CC bond breaking. This can be linked to the S1/S0 MECI shown in Fig. 2 (S1/S0, lower center).

FIG. 5.

2D projection of all CC bond lengths across all trajectories. The CC bonds are partitioned according to the labeling on the inset. Each trajectory is clustered based on the final geometry of the trajectory, resulting in four channels: (i) FC, (ii) ring opening, (iii) CO production, and (iv) ethene production. See the main text for more details.

FIG. 5.

2D projection of all CC bond lengths across all trajectories. The CC bonds are partitioned according to the labeling on the inset. Each trajectory is clustered based on the final geometry of the trajectory, resulting in four channels: (i) FC, (ii) ring opening, (iii) CO production, and (iv) ethene production. See the main text for more details.

Close modal
The nonadiabatic dynamics of cyclobutanone presented in Sec. IV C has served as the basis for calculating the rotationally averaged total electron scattering signal as a function of time, using the IAM as described in the theory section. To correctly account for the time evolution of the molecular geometries, we use a percent difference in the form,
(11)
where St(s, t) is the total electron scattering probability at time t and St(s, 0) is the total electron scattering probability at t = 0 (i.e., before the pump). In Fig. 6, the time-resolved UED pattern for cyclobutanone is shown as a function of the amplitude of the momentum transfer vector s. Due to the rapid loss of trajectories during the simulations, discussed in Sec. IV C, we only show the UED signal for the first 200 fs; however, caution should be exercised for the signal after ∼150 fs due to a limited number of trajectories. The total UED signal was re-calculated from the multistate MASH trajectories using a tighter total energy cutoff of 0.1 eV, with the result almost indistinguishable from the results shown in Fig. 6 (see the supplementary material).
FIG. 6.

Gas-phase UED pattern shown with percent difference as a function of time for cyclobutanone. All active trajectories are taken into account via the process described in Sec. III. The dashed horizontal lines represent the key features of the UED pattern.

FIG. 6.

Gas-phase UED pattern shown with percent difference as a function of time for cyclobutanone. All active trajectories are taken into account via the process described in Sec. III. The dashed horizontal lines represent the key features of the UED pattern.

Close modal

UED cross section percent differences shown in Fig. 6 present four distinct features at s ≈ 1.25, 2.4, 4.6, and 7.4 Å−1. These features constitute a fingerprint of the structural dynamics taking place in cyclobutanone; therefore, they can be interpreted with the help of additional static signals for the key geometries discussed in Sec. IV B (purple line shown in Fig. 7). A static signal for a fixed time [80 fs for panels (a) and (b) and 150 fs for panels (c) and (f)] is also shown in Fig. 7 for each pathway (green lines), showing some broadening of the peaks due to vibrational dispersion, along with subtle shifts of peak maxima. In addition, the electron scattering cross sections are calculated for the four individual outcomes using groups of trajectories in the dynamics simulation [(i), (ii), (iii), (iv) in Sec. IV C] and shown in Fig. 8 to aid the interpretation of the results. Due to significant similarities between the photoproducts and the high degree of symmetry in cyclobutanone, there is a large number of overlapping features in all static and individual cluster signals (see Figs. 7 and 8). Based on these similarities, we postulate that the strong negative signal at ∼1.25 Å−1 and the strong positive feature at ∼2.4 Å−1, common to all photoproducts is the breaking of one or two α-bonds to produce a ring-opened structure, CO and CH2CH2 CH2, or ethene and ketene. These two features, which reach a maximum at t ∼ 50 fs, are then related to the electronic state population transferring through the S2/S1 CI and subsequently through the S1/S0 CI [see Figs. 7(d), 7(e) and 7(f)]. This timescale is consistent with what is observed in the populations of S1 and the S0 ground state, as shown in Fig. 4, but significantly faster than that observed in Ref. 19. It is possible that our simulations underestimate the time in S2 due to a smaller barrier, as discussed in the context of Fig. 3 for SA(3)-CASSCF(12,12) compared to the matching XMS-CASPT2, resulting in somewhat fast decay and product formation.

FIG. 7.

Six static UED percent difference signals for key optimized geometries of gas-phase cyclobutanone and the static signal from an example structure obtained from our dynamics simulations showing CO dissociation (blue) and an average of the trajectories where the considered structure is the final product (green). The insets show the structure used to calculate the static signal: (a) S2 minimum, (b) S2/S1 compression, (c) S2/S1 α cleavage, (d) S1/S0 α cleavage, (e) S2/S1 ethene production, and finally (f) CO production. The vertical dashed lines match the horizontal dashed lines shown in Fig. 6, showing key features in the overall UED.

FIG. 7.

Six static UED percent difference signals for key optimized geometries of gas-phase cyclobutanone and the static signal from an example structure obtained from our dynamics simulations showing CO dissociation (blue) and an average of the trajectories where the considered structure is the final product (green). The insets show the structure used to calculate the static signal: (a) S2 minimum, (b) S2/S1 compression, (c) S2/S1 α cleavage, (d) S1/S0 α cleavage, (e) S2/S1 ethene production, and finally (f) CO production. The vertical dashed lines match the horizontal dashed lines shown in Fig. 6, showing key features in the overall UED.

Close modal
FIG. 8.

Percent difference UED patterns for each cluster of trajectories shown in Fig. 5. The subplots show the clustered UED for (a) α–CC cleavage, (b) ethene production, (c) CO production, and (d) FC. Clusters that do not contain data until the target 200 fs endpoint are truncated at the final data point.

FIG. 8.

Percent difference UED patterns for each cluster of trajectories shown in Fig. 5. The subplots show the clustered UED for (a) α–CC cleavage, (b) ethene production, (c) CO production, and (d) FC. Clusters that do not contain data until the target 200 fs endpoint are truncated at the final data point.

Close modal

The other UED feature common to the total signal (Fig. 6) and all individual pathways (Fig. 8) is an oscillatory positive signal at s ≈ 7.4 Å−1. For the initial 50 fs of simulation, in this s region, one can see strong coherent motions with an oscillatory period of 15 fs, indicating a possible coherent vibrational motion. All the bonds between heavy atoms (C and O) are plotted for an exemplary trajectory in the supplementary material (Fig. 6). Only the CO bond length has a period that matches these oscillations, suggesting that the carbonyl stretch may be responsible for this feature.

An additional feature can be observed at s ∼ 4.6 Å−1 with weak positive intensity for the initial 80 fs of the time-resolved UED becoming weakly negative after 80 fs. This positive weak feature (4%–6%) only shows in those pathways where β-bond breaking does not occur [Figs. 8(a) and 8(c)] and it is negative in the one where this happens [Fig. 8(b)]. When looking at the static signals shown in Fig. 7, one can see that the CO dissociation channel [Fig. 7(f)] yields a positive signal at s ∼ 4.6 Å−1 and that this feature is also present in the ring-open structures [Figs. 7(c) and 7(d)] at s ∼ 4.9 Å−1. Therefore, one can relate the sign and strength of this feature to the separation between the two carbons forming the β bonds in the molecule (see the inset in Fig. 5). The change of sign for this feature at t ≈ 80 fs, together with the great influence path (iv) has on the dynamics, suggests that ethene and ketene formation can be directly observed by looking at this region of the time-resolved UED signal.

It is important to note that the FC group presents oscillatory behavior in all the aforementioned UED features. These oscillations can be explained by taking into account that the molecule is vibrationally excited (i.e., hot) when the FC region is accessed, giving rise to high-amplitude motions. This behavior is not seen in the averaged signal due to the small contribution of this feature to the overall dynamics.

In summary, motivated by an experiment at the SLAC MeV-UED facility and the associated prediction challenge, we have applied multistate MASH and SA(3)-CASSCF(12,12)/aug-cc-pVDZ to identify the gas-phase photodynamics of cyclobutanone after photoexcitation by 200 nm pump pulses into the n-3s Rydberg state. Mechanistic details have been identified using both static analysis and nonadiabatic dynamics, finding a fast decay from the S2 state into the S1 via two CIs. A further two CIs allow for a similarly fast access to the ground state. This results in the same set of photoproducts previously observed, namely, α ring-opening and ethene + ketene production, the former of which can further dissociate on the ground electronic state to liberate CO.

For direct comparison with the experimental results, we have also predicted the gas-phase isotropic UED signal of cyclobutanone (see Fig. 6). We find that there is significant overlap between several of the products, meaning that the UED signal contains multiple features. However, several distinct features correspond to the formation of reaction products and loss of cyclobutanone in the equilibrium geometry. The timescale of the reaction can also be inferred from the UED signals, which, unsurprisingly, closely matches that of our simulated reaction dynamics. However, we do note that the reaction, in our simulations, proceeds significantly faster than previous observations.16,19

Significant caveats remain for our predictions. The instability of the active space means that we could potentially be blind to certain reaction channels. In addition, the value obtained for each channel’s quantum yield is susceptible to significant errors from the choice of the electronic structure method and the associated instabilities. Furthermore, the loss of trajectories means that the predictions for later times suffer from poor statistics and should be treated with caution. Notably, the convergence of properties such as quantum yields can be quite slow with respect to the number of trajectories.99 Nevertheless, experimental signals, especially for scattering, can converge for comparatively small numbers of representative trajectories.26,100–103 This, combined with our reasonably accurate electronic structure (known to be important for the validity of the trajectories96,99,104) lends credence to our predictions. In a broader context, this highlights the need for robust “black-box” electronic structure methods compatible with on-the-fly nonadiabatic dynamics simulations. Methods such as ADC(2) or TDDFT already exist but are not universally applicable as they have well-documented flaws. A potential way forward could be the selected CI family of methods,89,90 some benchmarking of which has been already been included in our paper. Further development and benchmarking of these methods is required to confirm whether they indeed offer an accurate electronic structure in a more robust, black-box fashion. In addition, it is important to add that the truncation of trajectories due to uncoupled hops affects the statistics of the dynamics and, therefore, needs to be refined in future work. Finally, the multistate MASH used in the simulations is still being developed, and we anticipate that refinement of the algorithm will improve the results further.

The supplementary material describes the active space employed in our calculations and presents the LIICs for the ACI and MCCI calculations along with further computational details for these methods. It also includes the total isotropic UED signal calculated with a tighter total energy threshold together with a description of the trajectories discarded due to the energy threshold. Finally, a summary of population estimators and a CO vibrational analysis is given.

The authors thank the organizers of the cyclobutanone prediction challenge for arranging this special edition. A.K., M.J.P., L.H., and A.P. acknowledge funding from the Leverhulme Trust (Grant No. RPG-2020-208), and A.K., A.M.C., and M.J.P. further acknowledge EPSRC EP/V006819 and EP/T021675/1. A.K. also acknowledges EPSRC EP/V049240 (together with M.S.), COSMOS Programme Grant (No. EP/X026973/1), and Grant No. DE-SC0020276 from the U.S. Department of Energy. J.E.R. was funded by a mobility fellowship from the Swiss National Science Foundation under Award No. P500PN_206641/1.

The authors have no conflicts to disclose.

Lewis Hutton: Data curation (equal); Formal analysis (equal); Investigation (equal); Writing – original draft (equal); Writing – review & editing (equal). Andrés Moreno Carrascosa: Conceptualization (equal); Data curation (supporting); Formal analysis (equal); Funding acquisition (equal); Writing – original draft (equal); Writing – review & editing (lead). Andrew W. Prentice: Data curation (supporting); Formal analysis (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Mats Simmermacher: Data curation (supporting); Formal analysis (supporting); Writing – original draft (supporting); Writing – review & editing (supporting). Johan E. Runeson: Data curation (supporting); Formal analysis (supporting); Writing – review & editing (supporting). Martin J. Paterson: Conceptualization (supporting); Data curation (supporting); Formal analysis (supporting); Writing – review & editing (supporting). Adam Kirrander: Conceptualization (equal); Formal analysis (supporting); Funding acquisition (lead); Writing – review & editing (equal).

All data and codes are available at reasonable request.

1.
S. W.
Benson
and
G. B.
Kistiakowsky
, “
The photochemical decomposition of cyclic ketones
,”
J. Am. Chem. Soc.
64
,
80
86
(
1942
).
2.
E. K. C.
Lee
,
R. G.
Shortridge
, and
C. F.
Rusbult
, “
Fluorescence excitation study of cyclobutanone, cyclopentanone, and cyclohexanone in the gas phase
,”
J. Am. Chem. Soc.
93
,
1863
1867
(
1971
).
3.
M.
Baba
and
I.
Hanazaki
, “
The S1(n ,π*) states of cyclopentanone and cyclobutanone in a supersonic nozzle beam
,”
J. Chem. Phys.
81
,
5426
5433
(
1984
).
4.
K. Y.
Tang
and
E. K. C.
Lee
, “
Laser photolysis of cyclobutanone. Photodecomposition from selected vibronic levels at long wavelengths
,”
J. Phys. Chem.
80
,
1833
1836
(
1976
).
5.
R.
Harrison
,
H.
Hawkins
,
R.
Leo
, and
P.
John
, “
The absorption of pulsed infrared radiation by cyclobutanone and its subsequent decomposition
,”
Chem. Phys. Lett.
70
,
555
559
(
1980
).
6.
G. C.
Causley
and
B. R.
Russell
, “
Electric dichroism spectroscopy in the vacuum ultraviolet. I. Cyclobutanone, cyclopentanone, cyclohexanone, and cycloheptanone
,”
J. Chem. Phys.
72
,
2623
2631
(
1980
).
7.
M. H.
Kao
,
R. K.
Venkatraman
,
M. N.
Ashfold
, and
A. J.
Orr-Ewing
, “
Effects of ring-strain on the ultrafast photochemistry of cyclic ketones
,”
Chem. Sci.
11
,
1991
2000
(
2020
).
8.
J. C.
Hemminger
and
E. K. C.
Lee
, “
Fluorescence excitation and photodecomposition of the first excited singlet cyclobutanone (1A2): A study of predissociation of and collisional energy transfer from the vibronically selected species
,”
J. Chem. Phys.
56
,
5284
5295
(
1972
).
9.
O. D.
Saltmarsh
and
R. G. W.
Norrish
, “
98. Primary photochemical reactions. Part VI. The photochemical decomposition of certain cyclic ketones
,”
J. Chem. Soc.
1935
,
455
459
.
10.
H. O.
Denschlag
and
E. K. C.
Lee
, “
Benzene photosensitization and direct photolysis of cyclobutanone and cyclobutanone-2-t in the gas phase
,”
J. Am. Chem. Soc.
90
,
3628
3638
(
1968
).
11.
N. E.
Lee
and
E. K. C.
Lee
, “
Tracer study of photochemically excited cyclobutanone-2-t and cyclobutanone. II. Detailed mechanism, energetics, unimolecular decomposition rates, and intermolecular vibrational energy transfer
,”
J. Chem. Phys.
50
,
2094
2107
(
1969
).
12.
E. W.
Diau
,
C.
Kötting
, and
A. H.
Zewail
, “
Femtochemistry of norrish type-1 reactions: II. The anomalous predissociation dynamics of cyclobutanone on the S1 surface
,”
ChemPhysChem
2
,
294
309
(
2001
).
13.
C. R.
Drury-Lessard
and
D. C.
Moule
, “
Ring puckering in the 1B2(n, 3s) Rydberg electronic state of cyclobutanone
,”
J. Chem. Phys.
68
,
5392
5395
(
1978
).
14.
L.
O’Toole
,
P.
Brint
,
C.
Kosmidis
,
G.
Boulakis
, and
P.
Tsekeris
, “
Vacuum-ultraviolet absorption spectra of propanone, butanone and the cyclic ketones CnH2n−2O (n = 4, 5, 6, 7)
,”
J. Chem. Soc., Faraday Trans.
87
,
3343
3351
(
1991
).
15.
R. F.
Whitlock
and
A. B.
Duncan
, “
Electronic spectrum of cyclobutanone
,”
J. Chem. Phys.
55
,
218
224
(
1971
).
16.
T. S.
Kuhlman
,
T. I.
Sølling
, and
K. B.
Møller
, “
Coherent motion reveals non-ergodic nature of internal conversion between excited states
,”
ChemPhysChem
13
,
820
827
(
2012
).
17.
S. H.
Xia
,
X. Y.
Liu
,
Q.
Fang
, and
G.
Cui
, “
Excited-state ring-opening mechanism of cyclic ketones: A MS-CASPT2//CASSCF study
,”
J. Phys. Chem. A
119
,
3569
3576
(
2015
).
18.
L.
Liu
and
W. H.
Fang
, “
New insights into photodissociation dynamics of cyclobutanone from the AIMS dynamic simulation
,”
J. Chem. Phys.
144
,
144317
(
2016
).
19.
T. S.
Kuhlman
,
S. P. A.
Sauer
,
T. I.
Sølling
, and
K. B.
Møller
, “
Symmetry, vibrational energy redistribution and vibronic coupling: The internal conversion processes of cycloketones
,”
J. Chem. Phys.
137
,
22A522
(
2012
).
20.
T. J.
Martínez
,
T.
Wolf
,
Slavíček
,
G.
Worth
,
M.
Barbatti
,
B.
Curchod
, and
S.
Bonella
, “
Prediction challenge: Cyclobutanone photochemistry
,” https://publishing.aip.org/publications/journals/special-topics/jcp/prediction-challenge-cyclobutanone-photochemistry/, accessed 02 July 2024.
21.
T.
Ishikawa
,
S. A.
Hayes
,
S.
Keskin
,
G.
Corthey
,
M.
Hada
,
K.
Pichugin
,
A.
Marx
,
J.
Hirscht
,
K.
Shionuma
,
K.
Onda
,
Y.
Okimoto
,
S.-y.
Koshihara
,
T.
Yamamoto
,
H.
Cui
,
M.
Nomura
,
Y.
Oshima
,
M.
Abdel-Jawad
,
R.
Kato
, and
R. J. D.
Miller
, “
Direct observation of collective modes coupled to molecular orbital-driven charge transfer
,”
Science
350
,
1501
1505
(
2015
).
22.
J.
Yang
,
J.
Beck
,
C. J.
Uiterwaal
, and
M.
Centurion
, “
Imaging of alignment and structural changes of carbon disulfide molecules using ultrafast electron diffraction
,”
Nat. Commun.
6
,
8172
(
2015
).
23.
J.
Yang
,
X.
Zhu
,
T. J.
Wolf
,
Z.
Li
,
J. P. F.
Nunes
,
R.
Coffee
,
J. P.
Cryan
,
M.
Gühr
,
K.
Hegazy
,
T. F.
Heinz
et al, “
Imaging CF3I conical intersection and photodissociation dynamics with ultrafast electron diffraction
,”
Science
361
,
64
67
(
2018
).
24.
T. J. A.
Wolf
,
D. M.
Sanchez
,
J.
Yang
,
R. M.
Parrish
,
J. P. F.
Nunes
,
M.
Centurion
,
R.
Coffee
,
J. P.
Cryan
,
M.
Gühr
,
K.
Hegazy
,
A.
Kirrander
,
R. K.
Li
,
J.
Ruddock
,
X.
Shen
,
T.
Vecchione
,
S. P.
Weathersby
,
P. M.
Weber
,
K.
Wilkin
,
H.
Yong
,
Q.
Zheng
,
X. J.
Wang
,
M. P.
Minitti
, and
T. J.
Martínez
, “
The photochemical ring-opening of 1,3-cyclohexadiene imaged by ultrafast electron diffraction
,”
Nat. Chem.
11
,
504
509
(
2019
).
25.
W. O.
Razmus
,
K.
Acheson
,
P.
Bucksbaum
,
M.
Centurion
,
E.
Champenois
,
I.
Gabalski
,
M. C.
Hoffman
,
A.
Howard
,
M.-F.
Lin
,
Y.
Liu
,
P.
Nunes
,
S.
Saha
,
X.
Shen
,
M.
Ware
,
E. M.
Warne
,
T.
Weinacht
,
K.
Wilkin
,
J.
Yang
,
T. J. A.
Wolf
,
A.
Kirrander
,
R. S.
Minns
, and
R.
Forbes
, “
Multichannel photodissociation dynamics in CS2 studied by ultrafast electron diffraction
,”
Phys. Chem. Chem. Phys.
24
,
15416
15427
(
2022
).
26.
M. P.
Minitti
,
J. M.
Budarz
,
A.
Kirrander
,
J. S.
Robinson
,
D.
Ratner
,
T. J.
Lane
,
D.
Zhu
,
J. M.
Glownia
,
M.
Kozina
,
H. T.
Lemke
,
M.
Sikorski
,
Y.
Feng
,
S.
Nelson
,
K.
Saita
,
B.
Stankus
,
T.
Northey
,
J. B.
Hastings
, and
P. M.
Weber
, “
Imaging molecular motion: Femtosecond x-ray scattering of an electrocyclic chemical reaction
,”
Phys. Rev. Lett.
114
,
255501
(
2015
).
27.
J. M.
Ruddock
,
N.
Zotev
,
B.
Stankus
,
H.-W.
Yong
,
D.
Bellshaw
,
S.
Boutet
,
T. J.
Lane
,
M.
Liang
,
S.
Carbajo
,
W.
Du
,
A.
Kirrander
,
M. P.
Minitti
, and
P. M.
Weber
, “
Simplicity beneath complexity: Counting molecular electrons reveals transients and kinetics of photodissociation reactions
,”
Angew. Chem., Int. Ed.
58
,
6371
6375
(
2019
).
28.
J. M.
Ruddock
,
H.
Yong
,
B.
Stankus
,
W.
Du
,
N.
Goff
,
Y.
Chang
,
A.
Odate
,
A. M.
Carrascosa
,
D.
Bellshaw
,
N.
Zotev
,
M.
Liang
,
S.
Carbajo
,
J.
Koglin
,
J. S.
Robinson
,
S.
Boutet
,
A.
Kirrander
,
M. P.
Minitti
, and
P. M.
Weber
, “
A deep UV trigger for ground-state ring-opening dynamics of 1,3-cyclohexadiene
,”
Sci. Adv.
5
,
eaax6625
(
2019
).
29.
I.
Gabalski
,
M.
Sere
,
K.
Acheson
,
F.
Allum
,
S.
Boutet
,
G.
Dixit
,
R.
Forbes
,
J. M.
Glownia
,
N.
Goff
,
K.
Hegazy
,
A. J.
Howard
,
M.
Liang
,
M. P.
Minitti
,
R. S.
Minns
,
A.
Natan
,
N.
Peard
,
W. O.
Rasmus
,
R. J.
Sension
,
M. R.
Ware
,
P. M.
Weber
,
N.
Werby
,
T. J. A.
Wolf
,
A.
Kirrander
, and
P. H.
Bucksbaum
, “
Transient vibration and product formation of photoexcited CS2 measured by time-resolved x-ray scattering
,”
J. Chem. Phys.
157
,
164305
(
2022
).
30.
J.
Yang
,
X.
Zhu
,
J. P.
F Nunes
,
J. K.
Yu
,
R. M.
Parrish
,
T. J. A.
Wolf
,
M.
Centurion
,
M.
Gühr
,
R.
Li
,
Y.
Liu
,
B.
Moore
,
M.
Niebuhr
,
S.
Park
,
X.
Shen
,
S.
Weathersby
,
T.
Weinacht
,
T. J.
Martinez
, and
X.
Wang
, “
Simultaneous observation of nuclear and electronic dynamics by ultrafast electron diffraction
,”
Science
368
,
885
889
(
2020
).
31.
H.
Yong
,
N.
Zotev
,
J. M.
Ruddock
,
B.
Stankus
,
M.
Simmermacher
,
A. M.
Carrascosa
,
W.
Du
,
N.
Goff
,
Y.
Chang
,
D.
Bellshaw
,
M.
Liang
,
S.
Carbajo
,
J. E.
Koglin
,
J. S.
Robinson
,
S.
Boutet
,
M. P.
Minitti
,
A.
Kirrander
, and
P. M.
Weber
, “
Observation of the molecular response to light upon photoexcitation
,”
Nat. Commun.
11
,
2157
(
2020
).
32.
A. W.
Prentice
,
J. P.
Coe
, and
M. J.
Paterson
, “
Modular approach to selected configuration interaction in an arbitrary spin basis: Implementation and comparison of approaches
,”
J. Chem. Theory Comput.
19
,
9161
9176
(
2023
).
33.
D.
Jacquemin
,
F.
Kossoski
,
F.
Gam
,
M.
Boggio-Pasqua
, and
P. F.
Loos
, “
Reference vertical excitation energies for transition metal compounds
,”
J. Chem. Theory Comput.
19
,
8782
8800
(
2023
);arXiv:2309.17311.
34.
J. E.
Runeson
and
D. E.
Manolopoulos
, “
A multi-state mapping approach to surface hopping
,”
J. Chem. Phys.
159
,
094115
(
2023
).
35.
H.-D.
Meyer
,
U.
Manthe
, and
L.
Cederbaum
, “
The multi-configurational time-dependent Hartree approach
,”
Chem. Phys. Lett.
165
,
73
78
(
1990
).
36.
M. H.
Beck
,
A.
Jäckle
,
G. A.
Worth
, and
H. D.
Meyer
, “
The multiconfiguration time-dependent Hartree (MCTDH) method: A highly efficient algorithm for propagating wavepackets
,”
Phys. Rep.
324
,
1
105
(
2000
).
37.
H.
Wang
and
M.
Thoss
, “
Multilayer formulation of the multiconfiguration time-dependent Hartree theory
,”
J. Chem. Phys.
119
(
3
),
1289
1299
(
2003
).
38.
G. A.
Worth
,
M. A.
Robb
, and
I.
Burghardt
, “
A novel algorithm for non-adiabatic direct dynamics using variational Gaussian wavepackets
,”
Faraday Discuss.
127
,
307
(
2004
).
39.
B.
Lasorne
,
M. J.
Bearpark
,
M. A.
Robb
, and
G. A.
Worth
, “
Direct quantum dynamics using variational multi-configuration Gaussian wavepackets
,”
Chem. Phys. Lett.
432
,
604
609
(
2006
).
40.
G.
Richings
,
I.
Polyak
,
K.
Spinlove
,
G.
Worth
,
I.
Burghardt
, and
B.
Lasorne
, “
Quantum dynamics simulations using Gaussian wavepackets: The vMCG method
,”
Int. Rev. Phys. Chem.
34
,
269
308
(
2015
).
41.
B. F. E.
Curchod
and
T. J.
Martínez
, “
Ab initio nonadiabatic quantum molecular dynamics
,”
Chem. Rev.
118
,
3305
3336
(
2018
).
42.
T. J.
Martínez
,
M.
Ben-Nun
, and
R. D.
Levine
, “
Molecular collision dynamics on several electronic states
,”
J. Phys. Chem. A
101
,
6389
6402
(
1997
).
43.
T. J.
Martinez
,
M.
Ben-Nun
, and
R. D.
Levine
, “
Multi-electronic-state molecular dynamics: A wave function approach with applications
,”
J. Phys. Chem.
100
,
7884
7895
(
1996
).
44.
T. J.
Martínez
and
R. D.
Levine
, “
Non-adiabatic molecular dynamics: Split-operator multiple spawning with applications to photodissociation
,”
J. Chem. Soc., Faraday Trans.
93
,
941
947
(
1997
).
45.
D. V.
Shalashilin
, “
Quantum mechanics with the basis set guided by Ehrenfest trajectories: Theory and application to spin-boson model
,”
J. Chem. Phys.
130
,
244101
(
2009
).
46.
K.
Saita
and
D. V.
Shalashilin
, “
On-the-fly ab initio molecular dynamics with multiconfigurational Ehrenfest method
,”
J. Chem. Phys.
137
,
22A506
(
2012
).
47.
D. V.
Makhov
,
C.
Symonds
,
S.
Fernandez-Alberti
, and
D. V.
Shalashilin
, “
Ab initio quantum direct dynamics simulations of ultrafast photochemistry with multiconfigurational Ehrenfest approach
,”
Chem. Phys.
493
,
200
218
(
2017
).
48.
J. C.
Tully
, “
Molecular dynamics with electronic transitions
,”
J. Chem. Phys.
93
,
1061
1071
(
1990
).
49.
H. D.
Meyera
and
W. H.
Miller
, “
A classical analog for electronic degrees of freedom in nonadiabatic collision processes
,”
J. Chem. Phys.
70
,
3214
3223
(
1979
).
50.
S. J.
Cotton
and
W. H.
Miller
, “
Symmetrical windowing for quantum states in quasi-classical trajectory simulations
,”
J. Phys. Chem. A
117
,
7190
7194
(
2013
).
51.
W. H.
Miller
and
S. J.
Cotton
, “
Classical molecular dynamics simulation of electronically non-adiabatic processes
,”
Faraday Discuss.
195
,
9
30
(
2016
).
52.
J. E.
Runeson
and
J. O.
Richardson
, “
Generalized spin mapping for quantum-classical dynamics
,”
J. Chem. Phys.
152
,
084110
(
2020
).
53.
Quantum Chemistry and Dynamics of Excited States: Methods and Applications
, edited by
L.
González
and
R.
Lindh
, 1st ed. (
John Wiley and Sons
,
United Kingdom
,
2021
), pp.
1
662
.
54.
M.
Ruckenbauer
,
S.
Mai
,
P.
Marquetand
, and
L.
González
, “
Revealing deactivation pathways hidden in time-resolved photoelectron spectra
,”
Sci. Rep.
6
,
35522
(
2016
).
55.
D.
Bellshaw
,
D. A.
Horke
,
A. D.
Smith
,
H. M.
Watts
,
E.
Jager
,
E.
Springate
,
O.
Alexander
,
C.
Cacho
,
R. T.
Chapman
,
A.
Kirrander
, and
R. S.
Minns
, “
Ab-initio surface hopping and multiphoton ionisation study of the photodissociation dynamics of CS2
,”
Chem. Phys. Lett.
683
,
383
388
(
2017
).
56.
R. J.
Squibb
,
M.
Sapunar
,
A.
Ponzi
,
R.
Richter
,
A.
Kivimäki
,
O.
Plekan
,
P.
Finetti
,
N.
Sisourat
,
V.
Zhaunerchyk
,
T.
Marchenko
,
L.
Journel
,
R.
Guillemin
,
R.
Cucini
,
M.
Coreno
,
C.
Grazioli
,
M.
Di Fraia
,
C.
Callegari
,
K. C.
Prince
,
P.
Decleva
,
M.
Simon
,
J. H. D.
Eland
,
N.
Doslić
,
R.
Feifel
, and
M. N.
Piancastelli
, “
Acetylacetone photodynamics at a seeded free-electron laser
,”
Nat. Commun.
9
,
63
(
2018
).
57.
I.
Polyak
,
L.
Hutton
,
R.
Crespo-Otero
,
M.
Barbatti
, and
P. J.
Knowles
, “
Ultrafast Photoinduced dynamics of 1,3-cyclohexadiene using XMS-CASPT2 surface hopping
,”
J. Chem. Theory Comput.
15
,
3929
3940
(
2019
).
58.
P.
Chakraborty
,
Y.
Liu
,
T.
Weinacht
, and
S.
Matsika
, “
Excited state dynamics of cis,cis-1,3-cyclooctadiene: Non-adiabatic trajectory surface hopping
,”
J. Chem. Phys.
152
,
174302
(
2020
).
59.
I. C. D.
Merritt
,
D.
Jacquemin
, and
M.
Vacher
, “
Cistrans photoisomerisation of azobenzene: a fresh theoretical look
,”
Phys. Chem. Chem. Phys.
23
,
19155
19165
(
2021
).
60.
L.
Hutton
and
B. F. E.
Curchod
, “
Photodynamics of gas-phase pyruvic acid following light absorption in the actinic region
,”
ChemPhotoChem
6
,
e202200151
(
2022
).
61.
K. D.
Borne
,
J. C.
Cooper
,
M. N. R.
Ashfold
,
J.
Bachmann
,
S.
Bhattacharyya
,
R.
Boll
,
M.
Bonanomi
,
M.
Bosch
,
C.
Callegari
,
M.
Centurion
,
M.
Coreno
,
B. F. E.
Curchod
,
M. B.
Danailov
,
A.
Demidovich
,
M.
Di Fraia
,
B.
Erk
,
D.
Faccialà
,
R.
Feifel
,
R. J. G.
Forbes
,
C. S.
Hansen
,
D. M. P.
Holland
,
R. A.
Ingle
,
R.
Lindh
,
L.
Ma
,
H. G.
McGhee
,
S. B.
Muvva
,
J. P. F.
Nunes
,
A.
Odate
,
S.
Pathak
,
O.
Plekan
,
K. C.
Prince
,
P.
Rebernik
,
A.
Rouzée
,
A.
Rudenko
,
A.
Simoncig
,
R. J.
Squibb
,
A. S.
Venkatachalam
,
C.
Vozzi
,
P. M.
Weber
,
A.
Kirrander
, and
D.
Rolles
, “
Ultrafast electronic relaxation pathways of the molecular photoswitch quadricyclane
,”
Nat. Chem.
16
,
499
(
2024
).
62.
M.
Barbatti
, “
Nonadiabatic dynamics with trajectory surface hopping method
,”
WIREs Comput. Mol. Sci.
1
,
620
633
(
2011
).
63.
S.
Blanes
and
P.
Moan
, “
Practical symplectic partitioned Runge–Kutta and Runge–Kutta–Nyström methods
,”
J. Comput. Appl. Math.
142
,
313
330
(
2002
).
64.
R.
Crespo-Otero
and
M.
Barbatti
, “
Recent advances and perspectives on nonadiabatic mixed quantum–Classical dynamics
,”
Chem. Rev.
118
,
7026
7068
(
2018
).
65.
S.
Hammes-Schiffer
and
J. C.
Tully
, “
Proton transfer in solution: Molecular dynamics with quantum transitions
,”
J. Chem. Phys.
101
,
4657
4667
(
1994
).
66.
F.
Plasser
,
M.
Ruckenbauer
,
S.
Mai
,
M.
Oppel
,
P.
Marquetand
, and
L.
González
, “
Efficient and flexible computation of many-electron wave function overlaps
,”
J. Chem. Theory Comput.
12
,
1207
1219
(
2016
).
67.
G.
Granucci
,
M.
Persico
, and
A.
Zoccante
, “
Including quantum decoherence in surface hopping
,”
J. Chem. Phys.
133
,
134111
(
2010
).
68.
J. E.
Subotnik
,
A.
Jain
,
B.
Landry
,
A.
Petit
,
W.
Ouyang
, and
N.
Bellonzi
, “
Understanding the surface hopping view of electronic transitions and decoherence
,”
Annu. Rev. Phys. Chem.
67
,
387
417
(
2016
).
69.
J. R.
Mannouch
and
J. O.
Richardson
, “
A mapping approach to surface hopping
,”
J. Chem. Phys.
158
,
104111
(
2023
).
70.
J. E.
Runeson
,
T. P.
Fay
, and
D. E.
Manolopoulos
, “
Exciton dynamics from the mapping approach to surface hopping: Comparison with Förster and Redfield theories
,”
Phys. Chem. Chem. Phys.
26
,
4929
(
2024
).
71.
P.
Debye
,
Phys. Z.
31
,
419
428
(
1930
).
72.
L.
Bewilogua
,
Phys. Z.
33
,
688
692
(
1932
).
73.
A.
Kirrander
and
P. M.
Weber
, “
Fundamental limits on spatial resolution in ultrafast X-ray diffraction
,”
Appl. Sci.
7
,
534
(
2017
).
74.
M.
Simmermacher
,
P. M.
Weber
, and
A.
Kirrander
, “
Theory of time-dependent scattering
,” in
Structural Dynamics with X-Ray and Electron Scattering
,
Theoretical and Computational Chemistry Series
, 251st ed., edited by
K.
Amini
,
A.
Rouzée
, and
M. J. J.
Vrakking
(
Royal Society of Chemistry
,
United Kingdom
,
2023
), Chap. 3, p.
85
.
75.
International Tables for Crystallography Volume C: Mathematical, Physical and Chemical Tables
, edited by
E.
Prince
, 2006th ed. (
Wiley
,
2006
), ISBN: 978-1-4020-1900-5.
76.
M.
Simmermacher
,
A.
Kirrander
, and
N. E.
Henriksen
, “
Time-resolved x-ray scattering from impulsively aligned or oriented molecules
,”
Phys. Rev. A
102
,
052825
(
2020
).
77.
F.
Salvat
,
A.
Jablonski
, and
C. J.
Powell
, “
elsepa—Dirac partial-wave calculation of elastic scattering of electrons and positrons by atoms, positive ions and molecules
,”
Comput. Phys. Commun.
165
,
157
190
(
2005
).
78.
A. M.
Carrascosa
,
H.
Yong
,
D. L.
Crittenden
,
P. M.
Weber
, and
A.
Kirrander
, “
Ab-initio calculation of total x-ray scattering from molecules
,”
J. Chem. Theory Comput.
15
,
2836
2846
(
2019
).
79.
N.
Zotev
,
A.
Moreno Carrascosa
,
M.
Simmermacher
, and
A.
Kirrander
, “
Excited electronic states in total isotropic scattering from molecules
,”
J. Chem. Theory Comput.
16
,
2594
2605
(
2020
).
80.
A.
Moreno Carrascosa
,
J. P.
Coe
,
M.
Simmermacher
,
M. J.
Paterson
, and
A.
Kirrander
, “
Towards high-resolution X-ray scattering as a probe of electron correlation
,”
Phys. Chem. Chem. Phys.
24
,
24542
24552
(
2022
).
81.
B. O.
Roos
,
P. R.
Taylor
, and
P. E.
Sigbahn
, “
A complete active space SCF method (CASSCF) using a density matrix formulated super-CI approach
,”
Chem. Phys.
48
,
157
173
(
1980
).
82.
D. E.
Woon
and
T. H.
Dunning
, “
Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon
,”
J. Chem. Phys.
98
,
1358
1371
(
1993
).
83.
F.
Aquilante
,
J.
Autschbach
,
A.
Baiardi
,
S.
Battaglia
,
V. A.
Borin
,
L. F.
Chibotaru
,
I.
Conti
,
L.
De Vico
,
M.
Delcey
,
I.
Fdez Galván
,
N.
Ferré
,
L.
Freitag
,
M.
Garavelli
,
X.
Gong
,
S.
Knecht
,
E. D.
Larsson
,
R.
Lindh
,
M.
Lundberg
,
P. Å.
Malmqvist
,
A.
Nenov
,
J.
Norell
,
M.
Odelius
,
M.
Olivucci
,
T. B.
Pedersen
,
L.
Pedraza-González
,
Q. M.
Phung
,
K.
Pierloot
,
M.
Reiher
,
I.
Schapiro
,
J.
Segarra-Martí
,
F.
Segatta
,
L.
Seijo
,
S.
Sen
,
D.-C.
Sergentu
,
C. J.
Stein
,
L.
Ungur
,
M.
Vacher
,
A.
Valentini
, and
V.
Veryazov
, “
Modern quantum chemistry with [Open]Molcas
,”
J. Chem. Phys.
152
,
214117
(
2020
).
84.
G.
Li Manni
,
I.
Fdez Galván
,
A.
Alavi
,
F.
Aleotti
,
F.
Aquilante
,
J.
Autschbach
,
D.
Avagliano
,
A.
Baiardi
,
J. J.
Bao
,
S.
Battaglia
,
L.
Birnoschi
,
A.
Blanco-González
,
S. I.
Bokarev
,
R.
Broer
,
R.
Cacciari
,
P. B.
Calio
,
R. K.
Carlson
,
R.
Carvalho Couto
,
L.
Cerdán
,
L. F.
Chibotaru
,
N. F.
Chilton
,
J. R.
Church
,
I.
Conti
,
S.
Coriani
,
J.
Cuéllar-Zuquin
,
R. E.
Daoud
,
N.
Dattani
,
P.
Decleva
,
C.
de Graaf
,
M. G.
Delcey
,
L.
De Vico
,
W.
Dobrautz
,
S. S.
Dong
,
R.
Feng
,
N.
Ferré
,
M.
Filatov(Gulak)
,
L.
Gagliardi
,
M.
Garavelli
,
L.
González
,
Y.
Guan
,
M.
Guo
,
M. R.
Hennefarth
,
M. R.
Hermes
,
C. E.
Hoyer
,
M.
Huix-Rotllant
,
V. K.
Jaiswal
,
A.
Kaiser
,
D. S.
Kaliakin
,
M.
Khamesian
,
D. S.
King
,
V.
Kochetov
,
M.
Krośnicki
,
A. A.
Kumaar
,
E. D.
Larsson
,
S.
Lehtola
,
M.-B.
Lepetit
,
H.
Lischka
,
P.
López Ríos
,
M.
Lundberg
,
D.
Ma
,
S.
Mai
,
P.
Marquetand
,
I. C. D.
Merritt
,
F.
Montorsi
,
M.
Mörchen
,
A.
Nenov
,
V. H. A.
Nguyen
,
Y.
Nishimoto
,
M. S.
Oakley
,
M.
Olivucci
,
M.
Oppel
,
D.
Padula
,
R.
Pandharkar
,
Q. M.
Phung
,
F.
Plasser
,
G.
Raggi
,
E.
Rebolini
,
M.
Reiher
,
I.
Rivalta
,
D.
Roca-Sanjuán
,
T.
Romig
,
A. A.
Safari
,
A.
Sánchez-Mansilla
,
A. M.
Sand
,
I.
Schapiro
,
T. R.
Scott
,
J.
Segarra-Martí
,
F.
Segatta
,
D.-C.
Sergentu
,
P.
Sharma
,
R.
Shepard
,
Y.
Shu
,
J. K.
Staab
,
T. P.
Straatsma
,
L. K.
Sørensen
,
B. N. C.
Tenorio
,
D. G.
Truhlar
,
L.
Ungur
,
M.
Vacher
,
V.
Veryazov
,
T. A.
Voß
,
O.
Weser
,
D.
Wu
,
X.
Yang
,
D.
Yarkony
,
C.
Zhou
,
J. P.
Zobel
, and
R.
Lindh
, “
The OpenMolcas Web: A community-driven approach to advancing computational chemistry
,”
J. Chem. Theory Comput.
19
,
6933
6991
(
2023
).
85.
T.
Shiozaki
,
W.
Győrffy
,
P.
Celani
, and
H.-J.
Werner
, “
Communication: Extended multi-state complete active space second-order perturbation theory: Energy and nuclear gradients
,”
J. Chem. Phys.
135
,
081106
(
2011
).
86.
N.
Forsberg
and
P.-Å.
Malmqvist
, “
Multiconfiguration perturbation theory with imaginary level shift
,”
Chem. Phys. Lett.
274
,
196
204
(
1997
).
87.
J. B.
Schriber
and
F. A.
Evangelista
, “
Communication: An adaptive configuration interaction approach for strongly correlated electrons with tunable accuracy
,”
J. Chem. Phys.
144
,
161106
(
2016
).
88.
J. B.
Schriber
and
F. A.
Evangelista
, “
Adaptive configuration interaction for computing challenging electronic excited states with tunable accuracy
,”
J. Chem. Theory Comput.
13
,
5354
5366
(
2017
).
89.
L.
Tong
,
M.
Nolan
,
T.
Cheng
, and
J.
Greer
, “
A Monte Carlo configuration generation computer program for the calculation of electronic states of atoms, molecules, and quantum dots
,”
Comput. Phys. Commun.
131
,
142
163
(
2000
).
90.
J. P.
Coe
and
M. J.
Paterson
, “
State-averaged Monte Carlo configuration interaction applied to electronically excited states
,”
J. Chem. Phys.
139
,
154103
(
2013
).
91.
A.
Prentice
,
J.
Coe
, and
M.
Paterson
(
2023
). “
Generalsci 1.0
,” Zenodo. https://doi.org/10.5281/zenodo.10203394
92.
R.
Crespo-Otero
and
M.
Barbatti
, “
Spectrum simulation and decomposition with nuclear ensemble: Formal derivation and application to benzene, furan and 2-phenylfuran
,”
Theor. Chem. Acc.
131
,
1237
(
2012
).
93.
S.
Mai
,
P.
Marquetand
, and
L.
González
, “
Nonadiabatic dynamics: The SHARC approach
,”
WIREs Comput. Mol. Sci.
8
,
e1370
(
2018
).
94.
J. W.
Park
and
T.
Shiozaki
, “
On-the-fly CASPT2 surface-hopping dynamics
,”
J. Chem. Theory Comput.
13
,
3676
3683
(
2017
).
95.
Y.
Gan
,
L.
Yue
,
X.
Guo
,
C.
Zhu
, and
Z.
Cao
, “
Multi-state nonadiabatic deactivation mechanism of coumarin revealed by ab initio on-the-fly trajectory surface hopping dynamic simulation
,”
Phys. Chem. Chem. Phys.
19
,
12094
12106
(
2017
).
96.
T. V.
Papineau
,
D.
Jacquemin
, and
M.
Vacher
, “
Which electronic structure method to choose in trajectory surface hopping dynamics simulations? Azomethane as a case study
,”
J. Phys. Chem. Lett.
15
,
636
643
(
2024
).
97.
A.
Udvarhazi
and
M. A.
El-Sayed
, “
Vacuum-ultraviolet spectra of cyclic ketones
,”
J. Chem. Phys.
42
,
3335
3336
(
1965
).
98.
I.
Fdez Galván
,
M. G.
Delcey
,
T. B.
Pedersen
,
F.
Aquilante
, and
R.
Lindh
, “
Analytical state-average complete-active-space self-consistent field nonadiabatic coupling vectors: Implementation with density-fitted two-electron integrals and application to conical intersections
,”
J. Chem. Theory Comput.
12
,
3636
3653
(
2016
).
99.
D.
Bellshaw
,
R. S.
Minns
, and
A.
Kirrander
, “
Correspondence between electronic structure calculations and simulations: Nonadiabatic dynamics in CS2
,”
Phys. Chem. Chem. Phys.
21
,
14226
14237
(
2019
).
100.
A.
Kirrander
,
K.
Saita
, and
D. V.
Shalashilin
, “
Ultrafast X-ray scattering from molecules
,”
J. Chem. Theory Comput.
12
,
957
967
(
2016
).
101.
K.
Acheson
and
A.
Kirrander
, “
Robust inversion of time-resolved data via forward-optimization in a trajectory basis
,”
J. Chem. Theory Comput.
19
,
2721
2734
(
2023
).
102.
C. C.
Pemberton
,
Y.
Zhang
,
K.
Saita
,
A.
Kirrander
, and
P. M.
Weber
, “
From the (1B) spectroscopic state to the photochemical product of the ultrafast ring-opening of 1,3-cyclohexadiene: A spectral observation of the complete reaction path
,”
J. Phys. Chem. A
119
,
8832
(
2015
).
103.
M.
Tudorovskaya
,
R. S.
Minns
, and
A.
Kirrander
, “
Effects of probe energy and competing pathways on time-resolved photoelectron spectroscopy: The ring-opening of 1,3-cyclohexadiene
,”
Phys. Chem. Chem. Phys.
20
,
17714
17726
(
2018
).
104.
J.
Janoš
and
P.
Slavíček
, “
What controls the quality of photodynamical simulations? Electronic structure versus nonadiabatic algorithm
,”
J. Chem. Theory Comput.
19
,
8273
8284
(
2023
).
Published open access through an agreement with University of Oxford