Quantum dynamics simulations are becoming a standard tool for simulating photo-excited molecular systems involving a manifold of coupled states, known as non-adiabatic dynamics. While these simulations have had many successes in explaining experiments and giving details of non-adiabatic transitions, the question remains as to their predictive power. In this work, we present a set of quantum dynamics simulations on cyclobutanone using both grid-based multi-configuration time-dependent Hartree and direct dynamics variational multi-configuration Gaussian methods. The former used a parameterized vibronic coupling model Hamiltonian, and the latter generated the potential energy surfaces on the fly. The results give a picture of the non-adiabatic behavior of this molecule and were used to calculate the signal from a gas-phase ultrafast electron diffraction (GUED) experiment. Corresponding experimental results will be obtained and presented at a later stage for comparison to test the predictive power of the methods. The results show that over the first 500 fs after photo-excitation to the S2 state, cyclobutanone relaxes quickly to the S1 state, but only a small population relaxes further to the S0 state. No significant transfer of population to the triplet manifold is found. It is predicted that the GUED experiments over this time scale will see signals related mostly to the C–O stretch motion and elongation of the molecular ring along the C–C–O axis.

Molecular quantum dynamics (QD) has become a powerful computational tool for understanding fundamental reactivity. By solving the time-dependent Schrödinger equation, these simulations can follow the time evolution of nuclei after a molecular system is prepared in a particular way, for example, in a molecular beam or pump–probe experiment. The key feature is that the quantum nature of the nuclei is taken into account, which is required for a complete description of systems involving either tunneling or non-adiabatic effects.

A present question, which is to be addressed in this paper, is whether QD simulations have predictive power, which is necessary for them to become a mature and trusted tool. This work addresses a challenge issued last year and aims to predict outcomes of experiments on cyclobutanone conducted at the SLAC megaelectronvolt ultrafast electron diffraction facility. These experiments commenced in late January 2024. The molecules will be photo-excited at 200 nm with a short (∼100 fs wide) pulse, and diffraction images will be recorded over many picoseconds. Up to 200 fs, images will be recorded every 30 fs, and then, the delay time will be increased. The challenge to simulation is to produce signals that can be compared to the experimental outputs and aid in their interpretation. Ultrafast electron diffraction is proving its value by providing data on structural dynamics in photochemistry.1,2 Examples are the dissociation of water,3 the ring-opening of cyclohexadiene,4 and imaging the passage of CH3I as it passes through a conical intersection.5 

Cyclobutanones are an interesting target as they exhibit distinct chemical reactivity compared to cyclic ketones with larger rings, primarily owing to their inherent ring strain of around 105–120 kJ mol−1.6,7 Alongside the considerable ring strain, there is a heightened electrophilicity of the carbonyl carbon atom, a ring puckering induced by steric interaction of substituents at C2 and C4, and significant photochemical reactivity. Despite this, cyclobutanones were long considered as academic curiosities.8 Since the mid-1970s, the role of cyclobutanones as synthetic reactants and intermediates has expanded, enabling the production of a wide variety of compounds with diverse applications; a small selection can be seen in the following articles.9–12 Reviews of relatively standard unsubstituted cyclobutanone reactions can be found in Refs. 6 and 8; for example, in a more recent publication, it has been reported that an isonitrile-based, four-component Ugi reaction,13 utilizing cyclobutanone, effectively produced an aspartame analog.14 

The earliest synthetic protocol for the production of cyclobutanone, from cyclobutanecarboxylic acid, authored by N. Kirscher,15 suffered from inefficiency, yielded low quantities, and featured several sequential reaction steps. Since then, several high-yield synthetic methods have been devised. One synthetic scheme entails a dialkylation process (utilizing 1-bromo-3-chloropropane) of 1,3-dithiane. This is followed by deprotection of the ketone through treatment with a mercury salt and cadmium carbonate.16 

There has been extensive research into the photo-induced behaviors of cyclobutanone, and its derivatives, both theoretically and experimentally. An early study of cyclobutanone showed two channels to photolytic decay: the C2 channel, whereby ketene and ethylene are produced, and the C3 channel, whereby cyclopropane/propene and CO are produced, with a ratio of 2:3.17 

Further theoretical and experimental studies have shown that the ratio of photoproducts is dependent on the wavelength of excitation. Upon excitation between 340 and 240 nm, the (nπ*) S1 state is accessed, with the absorption maximum at around 280 nm.18,19 Upon excitation to this state, three vibrational modes are activated corresponding to CO stretching, CO out-of-plane wagging, and ring puckering.20 Studies have indicated that following photoexcitation at shorter wavelengths (<315 nm range) an IC process via α-cleavage occurs resulting in the production of photofragments with a ratio (C2:C3) of ∼2:1.21,22 In the 345-315 nm range, an ISC process occurs via a triplet state, followed by the production of the photofragments, with a ratio (C2:C3) of ∼1:2 at 326.3 nm22,23 and ∼1:7 at 343.7 nm.22 This indicates an activation barrier to the breaking of the CC bond, a ring-opening process, which has been found to be between 9.6 kJ mol−120 and 29 kJ mol−1.24,25 By comparison, the barrier to ring opening of cyclopentanone and cyclohexanone is in excess of 63 kJ mol−1.20 Experimentally, the formation of hot ketene fragments from cyclobutanone in a cyclohexane solution occurred with a time delay of 0.25 ps, indicating an ultrafast ring-opening pathway. The subsequent relaxation of the vibrationally hot photoproducts and the growth of the ketene fundamental band were fitted to biexponential functions with shared time constants of 7 and 550 ps.24 A theoretical study using the ab initio multiple spawning method estimated the S1 lifetime to be around 484 fs with an average time taken for the α-cleavage occurring in 176.6 fs.26 It should be noted that theoretical findings are in the gas phase.

Experimental and theoretical studies have also explored the second excited (S2) state of cyclobutanone and other cyclic ketones. In the Franck–Condon region, the S2 state exhibits Rydberg 3s character and undergoes an internal conversion (IC) process, facilitated by the low frequency ring puckering mode, to the S1 (nπ*) state.27 The rate at which this 3s state decays was found, experimentally, to be around 0.74 ps. A complementary theoretical study, running quantum dynamics on potentials generated from a linear vibronic coupling Hamiltonian using five degrees of freedom, found the decay of the 3s state to be around 0.95 ps.28,29 It was also shown that in addition to the ring puckering mode, the CO out-of-plane deformation was significant. By comparison, the experimentally obtained ratio of the relative rates of the 3s → nπ* process in cyclobutanone, cyclopentanone, and cyclohexanone was determined to be 13:2:1.27 

The calculations presented here will use a standard work-flow for the study of a photo-excited molecule, in this case cyclobutanone, with different techniques building up a picture of the molecular dynamics. The work-flow has four stages. The first is to choose an appropriate level of quantum chemistry to describe the excited-state potential energy surfaces and couplings. The second is to build a model Hamiltonian using the vibronic-coupling scheme. This will provide information on which vibrational modes are excited on photo-excitation and which modes provide non-adiabatic coupling between the electronic states. The third step is to run grid-based QD simulations using the Multi-Configuration Time-Dependent Hartree (MCTDH) method. This will provide an accurate description of the short-time dynamics, along with absorption spectra. Finally, direct QD will be performed using the variational Multi-Configuration Gaussian (vMCG) method. In these calculations, the potential energy surfaces are calculated on the fly, allowing the molecule to undergo long-range motion, such as fragmentation, which is not possible in the MCTDH calculations due to the nature of the model Hamiltonian. The vMCG description also has an underlying trajectory nature, which will enable a straightforward simulation of the experimental signal and relate it to the evolving molecular geometry.

To provide a reference for the challenge, the original manuscript has been deposited on the arXiv at DOI 10.48550/arXiv.2402.09933.

All quantum chemistry calculations were performed using the Molpro 2022 package.30–32 The MCTDH and DD-vMCG simulations used a development version of the Quantics Suite,33,34 with the vibronic coupling Hamiltonians built using the VCHam package.35 

The starting point is to choose the level of electronic structure theory able to describe the system of interest, here cyclobutanone, balancing accuracy with cost. We start by finding and characterizing the ground-state equilibrium structure. For this, the coupled-cluster singles and doubles (CCSD) method was used with a 6-311++G** basis set.

The next step is to choose the level of theory needed to describe the excited-states. To describe potential long range motion, a multi-configurational method is best and as cyclobutanone is quite small with only five heavy atoms, a complete active space self-consistent field (CASSCF) wavefunction is possible. To select the complete active spaces (CASs), a number of spaces were tried: (6, 6), (6, 5), (8, 8), and (8, 10). State averaging over either five or six states was also investigated. The character of the states and their energies were compared to experimental results and previous calculations.20,21,37 After this, it was decided that a CAS(8, 10) with eight electrons in ten orbitals was required.

The CAS orbitals include the π orbitals and low lying Rydberg orbitals and are shown in Fig. 1. After calculating excitation energies at the C2v structure, it was decided that state-averaging over six states and the large, 6-311++G** basis set was required for stable results. Additionally, the inclusion of a second-order perturbation theory correction (CASPT2), using the RS2C method implemented in Molpro, was used to provide improved energies.

FIG. 1.

The 6-311++G** CAS(8,10) orbitals of cyclobutanone where orbitals 19 and 20 are the HOMO and LUMO, respectively.

FIG. 1.

The 6-311++G** CAS(8,10) orbitals of cyclobutanone where orbitals 19 and 20 are the HOMO and LUMO, respectively.

Close modal
Assuming a diabatic electronic basis, the Hamiltonian can be written in matrix form as
(1)
where T̂ is the kinetic energy operator, 1 is a unit matrix, and W is the diabatic potential matrix. Thus, in this representation, the kinetic energy operator is diagonal in the electronic states and the nuclear-electronic, non-adiabatic, coupling is represented by potential-like functions as the off-diagonal elements of the potential matrix. The key feature is that, unlike the usual adiabatic representation, the non-adiabatic couplings do not contain singularities.36,37
The vibronic coupling model represents the diabatic potentials using Taylor expansions around the Franck–Condon (FC) point and truncated to a low order,
(2)
In the standard Linear Vibronic Coupling (LVC) scheme employed here, the zero-order Hamiltonian is taken to be the ground-state Hamiltonian in the Harmonic approximation,
(3)
Here, the coordinates used are mass-frequency scaled normal modes with ωα representing the ground-state frequencies. The zero-order diabatic matrix
(4)
is a diagonal matrix with the energies of the states at the Franck–Condon point. The first-order expansion matrices
(5)
(6)
contain the linear on-diagonal coupling parameters, κα(i) along with the off-diagonal parameters, λα(ij). The former are the gradients of the potential surfaces at the Franck–Condon point, while the latter are the non-adiabatic couplings between states i and j. Higher-order matrices contain only on-diagonal terms as required to model the diabatic surfaces, hence the name linear vibronic coupling. Terms up to fourth-order were required for some modes due to the anharmonicity of the surfaces.
The parameters are only non-zero if they obey the symmetry rule that the product of normal modes and states involved in the expansion term must contain the totally symmetric representation. For example, the linear on-diagonal terms in terms of the related Hamiltonian matrix element are
(7)
where Hel is the usual electronic Hamiltonian with ψi, the electronic wavefunction at the Franck–Condon geometry, and the derivative is also evaluated at this point. This matrix element is only non-zero if Γi ×Γi ×ΓαA1, where Γi, Γα are the irreducible representations of the electronic states and normal mode, respectively, and A1 is the totally symmetric irrep. Thus, only totally symmetric vibrations have non-zero κ parameters. In a similar way, λ parameters are only non-zero if Γi ×Γj ×ΓαA1, and thus, only vibrations with the correct symmetry can couple states. These rules are easily extended to higher-order terms.
The MCTDH method provides a variational solution to the time dependent Schrödinger equation (TDSE) using the wavefunction ansatz,
(8)
i.e., a full direct product expansion of the wavefunction in terms of p sets of low-dimensional basis functions, φjκ(κ)(qκ,t), known as single-particle functions (SPFs). These functions are time-dependent and depend on a set of physical coordinates qκ = (Qα, Qβ, …). They are represented in a time-independent primitive basis, χr,
(9)
which provides an underlying grid. The functions |s⟩ are time-independent vectors that define the electronic state populated. This is usually referred to as the single-set formulation of MCTDH as a single set of SPFs are used to describe all states.

The method is thus a contraction scheme from the full basis set to a variational one. The resulting equations of motion for the time evolution of the SPFs and the expansion coefficients Aj1jps are well described in the literature38,39 and will be not be given here.

For large systems, the Multi-Layer Multi-Configuration Time-Dependent Hartree (ML-MCTDH) variant must be used.40–42 In this, the SPFs are expanded in the form of the MCTDH ansatz, Eq. (8). These new basis functions (SPFs of the first layer) can in turn be expanded in this form. This procedure is continued to provide layers of functions with a set of primitive grid functions forming the lowest layer. In this way, a full tensor contraction scheme is set up and used to variationally solve the TDSE.

The ML-MCTDH method was used in this work to allow for simulations, including all degrees of freedom. The Hamiltonian is provided by the vibronic model described above. The primitive basis functions used for all coordinates were harmonic oscillator discrete variable representations (DVRs).38 Calculations need to be converged with respect to the basis sets. In order to achieve this, the size of the primitive basis is checked to ensure that the wavepacket does not significantly populate the end grid points (population less than 10−6). A measure of the quality of the time-dependent basis set is given by the natural populations. These are the eigenvalues of the reduced density matrix associated with each set of basis functions.38,42 These eigenvalues are related to the importance of the functions, and if the lowest eigenvalue is small, adding further functions to the basis set will not significantly change the wavefunction. To keep a reasonable quality of wavefunction, the lowest natural populations are kept below 10−3. The latter was ensured throughout the calculation by dynamically growing the basis in any layer whenever the population of the least important SPF approaches this limit.43 

The MCTDH calculations using the vibronic coupling Hamiltonian are able to simulate the absorption spectrum from the Fourier transform of the autocorrelation function,
(10)
The function g(t) is used to remove artifacts from the Fourier transform. The chosen form is
(11)
where the first term is to ensure that the autocorrelation function goes to zero at the end of the simulation, time T, and the second term is a damping function. The damping time used was 150 fs.

In addition, diabatic state populations are directly obtained from the wavefunction, which give a rate of electronic relaxation as the molecule changes electronic configuration. The short-time dynamics may also be analyzed for the evolution of the molecular geometry in terms of excitation of the vibrational coordinates.

To create a more flexible wavefunction with a direct connection to molecular structures, the wavefunction ansatz can be written as a superposition of Gaussian functions,
(12)
where each Gaussian function is a separable product of one-dimensional Gaussian functions,
(13)
and s is the state index. These 1D Gaussian functions have the form
(14)
where ζ, ξ, and ηj are quadratic, linear, and scalar parameters, respectively.
Solving the TDSE using this ansatz, and the Dirac–Frenkel variational principle, leads to the variational Multiconfigurational Gaussian (vMCG) method.44,45 The expansion coefficients evolve with equations similar to those for MCTDH. However, the non-orthogonality of the GWP basis must be taken into account,
(15)
with the overlap matrix, S, Hamiltonian matrix, H, and overlap time-derivative, τ,
(16)
(17)
(18)
In the standard approach, the widths of the Gaussians, ζ, are kept fixed (frozen Gaussians) and the scalar parameter, η, is fixed by the requirement for the Gaussians to be normalized and for the phase to be kept zero. The time-dependence is then carried by the linear parameters, ξ. Collecting the set of parameters for a multi-dimensional Gaussian into a vector, Λ = ξ, the equations of motion (EOMs) for the Gaussian functions can be written as
(19)
Using the relationship between the general form of Eq. (14) and the Gaussian wavepackets of Heller,46 the linear parameters can be written in terms of the coordinate and momentum at the center of the Gaussian,
(20)
and the vector X in Eq. (19) is related to the classical equations of motion,
(21)
Thus, the Gaussians move along trajectories that have a classical component with an additional variational coupling. This is due to the C matrix, which is composed of Gaussian overlaps and moments, and the YR vector, which is composed of residual Hamiltonian matrix elements in the Gaussian basis after the classical terms have been removed. For full details of the vMCG equations of motion, see Ref. 45. This variational coupling leads to faster convergence of the wavefunction.
Due to the localized nature of a Gaussian function, it is reasonable to calculate integrals of the potential energy using a Local Harmonic Approximation (LHA) in which the potential is expanded to second order around the center of a Gaussian,
(22)
Solutions to the TDSE no longer converge on the exact result, but the integrals can all be performed analytically and the algorithm can be used for direct dynamics simulations.

Direct dynamics simulations use this formalism to calculate the potential from quantum chemistry calculations on the fly. The LHA requires the energies, gradients, and Hessians at the Gaussian central coordinate, and these can all be provided by a quantum chemistry calculation. In the standard DD-vMCG protocol, these quantities are calculated and stored in a database. New points are added to the database only when a GWP has coordinates that are significantly different from any point stored in the database. The surfaces experienced by the evolving GWPs are provided by Shepard interpolation between the points in the database. Calculations are run in the diabatic picture, and the potentials are diabatized using the propagation diabatization scheme. For full details, see Ref. 47.

The major effort in the vMCG method is the inversion of the C matrix in Eq. (19), which has the size of (Nf×n)2, where Nf is the number of degrees of freedom and n is the number of basis functions. This effort can be reduced by partitioning the system and making the wavefunction a multi-configurational product from two (or more) sets of functions,
(23)
This is the G-MCTDH ansatz.44 The EOMs are the same as Eqs. (15) and (19), but now, the function Gi is a configuration, i.e., a product of basis functions, Gi=Gi1(1)Gi2(2), and the propagation of the Gaussian parameters includes mean-field operators connecting the different partitions.
GUED uses high energy pulses of electrons to provide structural information in the form of a diffraction pattern. Electron scattering by a potential field (a molecule) can be described by measuring the momentum transfer from an electron to the target on collision. For a set of atoms under the independent atom model, this can be formulated to give an intensity as a function of scattered momentum transfer s that has two parts: an “atomic” due to scattering directly off an atom and a “molecular” due to interference from scattering off neighboring atoms.48 The atomic part is
(24)
and the molecular part is written as
(25)
where N is the number of atoms, fi are the “atomic form factors” accounting for the structure of an atom in terms of nuclei and electrons, and rij denotes the inter-atomic distance between the ith and jth atom. The total intensity is simply the sum of these two terms. Actual measurements record the scattering signal as a modified scattering intensity,
(26)
where the time dependence has been added to indicate that the signal is obtained by scattering off the sample at time t. Note that the atomic component is time-independent due to the independent atom model. It is more useful to record the signal as the difference between the signal at t and an initial (reference) signal,
(27)
as the change in the signal gives the structural changes in the molecule without needing to extract Imol. This signal can be related to the change in the “pair distribution function” (PDF) by integrating over s to give the signal as a function of inter-atomic distances,
(28)
where α is a smoothing factor. Changes in the molecular geometry can thus be seen in the PDF as losses and rises due to atoms moving relative to each other. It is, however, summed over all the pairs in a molecule, so each inter-nuclear distance may be related to a number of different atom pairs, making interpretation with a molecular simulation difficult. From a simulation, the PDF can be simulated by using the atomic coordinates along a trajectory to get first the molecular intensity, Eq. (25), and then obtain the modified scattering intensity before calculating the difference PDF of Eq. (28).
Finally, it needs to be taken into account that a molecule is not a classical object moving along a single trajectory. The scattering takes place from the molecular density given by the evolving wavepacket. This can be taken into account by integrating over the molecular configurations,
(29)
where |Ψ(q, t)|2 is the probability of the molecule being in configuration q at time t.
In vMCG, a useful approximation for calculating the expectation values of operators that depend only on the atomic coordinates is
(30)
(31)
where rj is the molecular configuration defined by the center coordinate of the GWP and O(qj) is the value of the operator at that point. The expectation value is thus approximated by a weighted sum along the GWP trajectories, with the weighting given by the gross Gaussian populations,
(32)
where Sij is the GWP overlap matrix. The final difference PDF from the signal calculated over the molecular density can thus be approximated in this way as a weighted sum,
(33)
where Imol,i is the molecular scattering intensity from vMCG trajectory i. In fact, this can be carried through to the procedure for calculating the difference PDF and one can simply provide a weighted sum of difference PDFs.

The modified scattering intensity for each trajectory at each time step was calculated using a code to simulate electron diffraction signals provided by Wolf et al.49 with subsequent weighting according to Eq. (33) to get the final time-dependent signal. The time-dependent difference pair distribution function ΔP(r, t) was then obtained using smin=0Å1, smax=10Å1, and α=0.036Å2 to match the experimental spatial resolution of 0.6Å.

The nuclear coordinates used in the simulations are mass-frequency scaled normal modes. As a result, displacements are dimensionless and have the scale that 1 unit of displacement stretches the normal mode to have a potential energy equal to the zero-point energy (12ω). In the following, distances are simply recorded as x units of displacement or when it is clear from the context that we are referring to a displacement, x units.

Using CCSD/6-311++G**, two structures of cyclobutanone were found at critical points on the ground-state potential surface. The first is planar with C2v symmetry. The second is bent and has Cs symmetry. These are shown in Fig. 2, and the coordinates, along with bond lengths and angles, are given in the supplementary material. The atom numbering was chosen to emphasize the proximities of the carbon atoms to the oxygen. Consequently, atom 1 is the closest carbon to the oxygen, carbons 2 and 3 are the symmetrically equivalent central atoms, and carbon 4 is the furthest away.

FIG. 2.

The optimized Cs and C2v ground-state structures of cyclobutanone with the atom numbering used throughout this paper. The angle shown on the Cs structure is the ring-puckering angle calculated at the CCSD/6-311++G** level of theory.

FIG. 2.

The optimized Cs and C2v ground-state structures of cyclobutanone with the atom numbering used throughout this paper. The angle shown on the Cs structure is the ring-puckering angle calculated at the CCSD/6-311++G** level of theory.

Close modal

The ring puckering angle between the C1C2C3 and C4C2C3 planes was found to be 171.7°, and the angle of the C–O bond to the C1C2C3 plane was 12.3°. Frequency calculations show that the Cs structure is a minimum and the C2v structure is a transition state. The frequencies of the vibrational modes of both structures are listed in the supplementary material. With one exception, the transition mode, changes in the normal modes, and frequencies are minimal between the two structures.

The transition mode with an imaginary frequency of 80 cm−1 is the ring puckering motion, ν1, and at Cs, it has a real frequency of 100 cm−1. The difference in energy between the structures ΔE=EC2vECs is the barrier height to the out-of-plane bend and has a value of 0.0087 eV. The frequency of the transition mode, ν1, means it has a zero point energy level at 0.0062 eV. Consequently, while the barrier is very small, the low frequency of the transition mode means that it is a stable structure, and at 0 K, the equilibrium structure will be Cs. However, at room temperature, a significant population of molecules will be in excited vibrational levels and the most probable structure will be closer to C2v. It was also found that at the CASSCF and CASPT2 levels of theory used in the dynamics calculations, the C2v structure is the minimum energy on the ground-state. We therefore will need to examine the dynamics starting from both structures.

The excitation energies and characters from both structures to the lowest two singlet and three triplet states are listed in Table I. Calculations used a CAS with eight electrons in ten orbitals and a 6-311++G** basis. The orbitals are shown in Fig. 1.

TABLE I.

Excitation energies (eV) of cyclobutanone calculation at the CASSCF and CASPT2 levels of theory with a CAS(8,10)/6-311++G** wavefunction. The symmetry labels refer to the structure that has been optimized using CCSD/6-311++G**.

C2vCs
StateCharacterCASSCFCASPT2CASSCFCASPT2Expt.
S0(A1 0.0 −0.0087a  
S1(A2πy*πx* 4.93 4.16 5.19 4.21 3.61b 
S2(B2πy* sRy 5.81 6.32 5.86 6.23 6.11c 
T1(A2πy*πx* 4.59 3.77 4.58 3.84  
T2(B2πy* sRy 5.73 6.27 5.85 6.18  
T3(A1πxπx* 6.44 6.69 6.48 6.24  
C2vCs
StateCharacterCASSCFCASPT2CASSCFCASPT2Expt.
S0(A1 0.0 −0.0087a  
S1(A2πy*πx* 4.93 4.16 5.19 4.21 3.61b 
S2(B2πy* sRy 5.81 6.32 5.86 6.23 6.11c 
T1(A2πy*πx* 4.59 3.77 4.58 3.84  
T2(B2πy* sRy 5.73 6.27 5.85 6.18  
T3(A1πxπx* 6.44 6.69 6.48 6.24  
a

Energy calculated CCSD/6-311++G**.

b

Adiabatic 0–0 transition, Ref. 22.

c

Adiabatic 0–0 transition, Ref. 50.

Little difference is found for the excitation from the two structures. In agreement with previous work,20,25,29 the lowest singlet state is an excitation from the in-plane πy* to the π* orbital at around 4.2 eV. The second state is an excitation out of the same orbital into a diffuse Rydberg orbital lying around 6.3 eV. Interestingly, the CASPT2 correction shifts this state up in energy. The T1 and T2 states are the same configurations as S1 and S2, but lie slightly lower in energy. The T3 state lies above S2 at 6.7 eV and is a π to π* excitation.

The vibronic coupling model Hamiltonian of Sec. II B was obtained using the mass-frequency scaled normal modes of the ground-state C2v structure, with the imaginary frequency of ν1 taken as real. The ground-state and excitation energies were calculated at the CASPT2 corrected SA6-CAS(8,10)/6-311++G** level of theory at 21 points along each vibrational mode, treating the singlet and triplet manifolds separately. The parameters of the model were then chosen to optimize the fit between the adiabatic potential surfaces of the model and the calculated adiabatic energies.51 Due to the anharmonicities of the potentials, quartic potentials (expansion to fourth order) were used for all modes. The diabatic coupling, however, was truncated to first order (linear vibronic coupling).

The singlet and triplet manifolds were then merged into one operator, and the spin–orbit couplings calculated at the Franck–Condon point were added as constant values between the singlet and triplet states using the magnitude of the complex components to give an effective coupling treating each triplet manifold as a single state. The excitation energies are given in Table I. The spin–orbit couplings, given as a normalized average over the three components of the triplet state, are given in Table II.

TABLE II.

Averaged spin–orbit coupling (cm−1) for cyclobutanone at the CASSCF level of theory with a SA6-CAS(8,10)/6-311++G** wavefunction.

StateT1(A2)T2(B2)T3(A1)
S0(A133.94 1.90 0.00 
S1(A20.00 0.76 15.42 
S2(B20.64 0.00 1.45 
StateT1(A2)T2(B2)T3(A1)
S0(A133.94 1.90 0.00 
S1(A20.00 0.76 15.42 
S2(B20.64 0.00 1.45 

The model provides information on the key vibrational modes: those that are directly vibrationally excited on electronic excitation and those that provide coupling between the states. Tables III and IV list the linear κ and λ parameters. Coupling strengths are defined by the ratios κα(i)/ωα and λα(ij)/ωα, where ωα is the frequency of the mode. Using these strengths as a criteria, a subset of 12 key modes can be identified. The ν5, ν16, ν19, ν20, and ν21 modes are the totally symmetric vibrations with the highest κ coupling strengths, i.e., they have significant gradients on the potential surfaces at the Franck–Condon point. Modes ν4 and ν9, with A2 symmetry, carry the coupling between S0 and S1. Modes ν3 and ν17, with B2 symmetry, couple between S0 and S2, while modes ν2 and ν25, with B1 symmetry, carry the coupling between S1 and S2. Finally, ν1 has symmetry B1 and dominates the coupling between the T1 and T2 states. The coupling between T1 and T3 is dominated by ν4 and between the T2 and T3 states by the ν11 mode. This latter coupling is, however, quite weak. Consequently, ν11 is not included in the key modes as the high energy T3 state is unlikely to be important for the short term dynamics. The key vibrations are plotted in Fig. 3.

TABLE III.

On-diagonal linear parameters from the vibronic coupling model of cyclobutatone calculated at the CASPT2 level with a SA6-CAS(10,8)/6-311++G** wavefunction.

ModeSymmetryκ(2) (eV)κ(3) (eV)κ(4) (eV)κ(5) (eV)κ(6) (eV)
ν5 A1 0.0727 0.0272 0.0435 0.0423 0.0571 
ν7 A1 0.0509 −0.0347 0.0186 −0.0118 −0.0118 
ν10 A1 0.0135 0.0331 0.0445 0.0116 0.0116 
ν16 A1 −0.0577 −0.1895 −0.0197 −0.0728 0.0073 
ν19 A1 −0.0419 −0.1246 −0.0370 −0.0970 −0.0155 
ν20 A1 0.0508 0.1094 0.0377 0.1110 0.0475 
ν21 A1 −0.3530 0.1541 −0.3070 −0.7209 0.0881 
ν23 A1 −0.0008 −0.0047 −0.0098 −0.0884 −0.0884 
ν24 A1 −0.0174 −0.0251 −0.0005 0.0007 0.0007 
ModeSymmetryκ(2) (eV)κ(3) (eV)κ(4) (eV)κ(5) (eV)κ(6) (eV)
ν5 A1 0.0727 0.0272 0.0435 0.0423 0.0571 
ν7 A1 0.0509 −0.0347 0.0186 −0.0118 −0.0118 
ν10 A1 0.0135 0.0331 0.0445 0.0116 0.0116 
ν16 A1 −0.0577 −0.1895 −0.0197 −0.0728 0.0073 
ν19 A1 −0.0419 −0.1246 −0.0370 −0.0970 −0.0155 
ν20 A1 0.0508 0.1094 0.0377 0.1110 0.0475 
ν21 A1 −0.3530 0.1541 −0.3070 −0.7209 0.0881 
ν23 A1 −0.0008 −0.0047 −0.0098 −0.0884 −0.0884 
ν24 A1 −0.0174 −0.0251 −0.0005 0.0007 0.0007 
TABLE IV.

Off-diagonal linear parameters from the vibronic coupling model of cyclobutatone calculated at the CASPT2 level with a SA6-CAS(10,8)/6-311++G** wavefunction. Modes with all parameters less than 10−2 are omitted.

ModeSymmetryλ(1,2)/ω (eV)λ(1,3) (eV)λ(2,3) (eV)λ(4,5) (eV)λ(4,6) (eV)λ(5,6) (eV)
ν1 B1 ⋯ ⋯ 0.0006 0.0569 ⋯ ⋯ 
ν2 B1 ⋯ ⋯ −0.0277 −0.0035 ⋯ ⋯ 
ν3 B2 ⋯ −0.2513 ⋯ ⋯ ⋯ ⋯ 
ν4 A2 −0.1699 ⋯ ⋯ ⋯ 0.0489 ⋯ 
ν8 B2 ⋯ 0.0149 ⋯ ⋯ ⋯ ⋯ 
ν9 A2 0.1913 ⋯ ⋯ ⋯ −0.0476 ⋯ 
ν11 B2 ⋯ −0.0307 ⋯ ⋯ ⋯ 0.0364 
ν12 B1 ⋯ ⋯ −0.0019 −0.0006 ⋯ ⋯ 
ν14 B1 ⋯ ⋯ 0.0006 0.0527 ⋯ ⋯ 
ν17 B2 ⋯ 0.1293 ⋯ ⋯ ⋯ −0.0172 
ν18 B2 ⋯ 0.0739 ⋯ ⋯ ⋯ 0.0319 
ν22 B2 ⋯ −0.0701 ⋯ ⋯ ⋯ ⋯ 
ν25 B1 ⋯ ⋯ 0.1351 −0.0761 ⋯ ⋯ 
ν26 A2 0.0069 ⋯ ⋯ ⋯ −0.0285 ⋯ 
ν27 B1 ⋯ ⋯ −0.0002 −0.0543 ⋯ ⋯ 
ModeSymmetryλ(1,2)/ω (eV)λ(1,3) (eV)λ(2,3) (eV)λ(4,5) (eV)λ(4,6) (eV)λ(5,6) (eV)
ν1 B1 ⋯ ⋯ 0.0006 0.0569 ⋯ ⋯ 
ν2 B1 ⋯ ⋯ −0.0277 −0.0035 ⋯ ⋯ 
ν3 B2 ⋯ −0.2513 ⋯ ⋯ ⋯ ⋯ 
ν4 A2 −0.1699 ⋯ ⋯ ⋯ 0.0489 ⋯ 
ν8 B2 ⋯ 0.0149 ⋯ ⋯ ⋯ ⋯ 
ν9 A2 0.1913 ⋯ ⋯ ⋯ −0.0476 ⋯ 
ν11 B2 ⋯ −0.0307 ⋯ ⋯ ⋯ 0.0364 
ν12 B1 ⋯ ⋯ −0.0019 −0.0006 ⋯ ⋯ 
ν14 B1 ⋯ ⋯ 0.0006 0.0527 ⋯ ⋯ 
ν17 B2 ⋯ 0.1293 ⋯ ⋯ ⋯ −0.0172 
ν18 B2 ⋯ 0.0739 ⋯ ⋯ ⋯ 0.0319 
ν22 B2 ⋯ −0.0701 ⋯ ⋯ ⋯ ⋯ 
ν25 B1 ⋯ ⋯ 0.1351 −0.0761 ⋯ ⋯ 
ν26 A2 0.0069 ⋯ ⋯ ⋯ −0.0285 ⋯ 
ν27 B1 ⋯ ⋯ −0.0002 −0.0543 ⋯ ⋯ 
FIG. 3.

The 12 key normal modes of cyclobutanone calculated at the CCSD/6-311++G** level on the ground-state C2v structure. Arrows indicate the motion of the atoms during each vibration. The symmetry labels are given in the supplementary material.

FIG. 3.

The 12 key normal modes of cyclobutanone calculated at the CCSD/6-311++G** level on the ground-state C2v structure. Arrows indicate the motion of the atoms during each vibration. The symmetry labels are given in the supplementary material.

Close modal

The parameters for the full final Hamiltonian are given as the quantics operator file in the supplementary material. Cuts through the potential surfaces for the full singlet and triple manifold are also given in the supplementary material. In the triplet manifold, the major features are due to the near degenerate T2 and T3 states, which are also near degenerate with the S2 state. The spin–orbit coupling, however, is weak. In this manifold, ν2, ν3, and ν16 show double well structures in T2, while in ν21, the T2 and T3 states cross at the Franck–Condon point and T2 appears to be dissociative. Mode ν21, the C–O stretch vibration, is the most interesting mode in the singlet manifold. The S2/S1 crossing is seen to be at a negative value along this mode and in a downhill direction from the FC point. The S1 state is also long and fairly flat out to positive values, but not dissociative. The ring stretching mode along the C–C–O axis, ν5, is also potentially interesting as it has a strong gradient at the FC point in all excited states.

ML-MCTDH simulations were run for 200 fs on the full singlet and triplet vibronic coupling operator of Sec. III C. The tree defining the layering along with the number of functions is given in the supplementary material. Two simulations were started with the initial wavepacket centered either at the C2v transition state geometry or the Cs minimum energy structure. In both cases, the wavepacket was a separable product of Gaussian functions with a width of 1/2 along each normal mode, i.e., the harmonic ground-state wavefunction in the mass-frequency scaled normal modes of the model. The wavepackets were then placed in the S2 state in a vertical excitation. At the end of this, less than 0.1% population flows into the T1 state, and the populations of the T2 and T3 are even smaller. This is not surprising, given the small size of the coupling. In both the S1 and S2 states, there is zero spin–orbit coupling with the closest lying triplet state (T1 and T2, respectively).

Of course, the vibronic coupling model treats the spin–orbit coupling with only a constant value taken at the Frank–Condon point. In order to discern the effect of the inclusion of nuclear motion on the spin–orbit coupling and the amount of ISC, DD-vMCG simulations were run, including all six states, calculating the spin–orbit coupling at each point added to the DB. All properties were calculated at the CAS(8,10) level. The wavepacket was partitioned into two parts with the 12 key modes in one partition and treated with 16 GWPs. The remaining modes in the second partition were treated with 8 GWPs. While this is a small basis set, 128 trajectories allows for a reasonable searching of the main configuration space to find any regions where ISC occurs.

At the end of the simulation, all three triplet states had accumulated ∼0.1% of the population. Consequently, while the inclusion of the nuclear motion has indeed increased the ISC, it is still not likely to be significant for the short-term dynamics (over the first 0.5 ps). As a result, the triplet states will be ignored in the final analysis.

ML-MCTDH simulations were run for 200 fs on the singlet manifold vibronic coupling model of cyclobutanone. Two sets of simulations were carried out as in the simulations of Sec. III D, including the triplet states, i.e., the initial wavepacket was centered at either the C2v or Cs structures. In each case, two simulations were performed, including either all modes or only the key 12 modes: ν1, ν2, ν3, ν4, ν5, ν9, ν16, ν17, ν19, ν20, ν21, and ν25. The tree structures and basis set sizes for converged calculations are given in the supplementary material.

Diabatic state populations for the first 100 fs after a vertical excitation to the S2 state are shown in Fig. 4 as there is little change during the subsequent period up to 200 fs. Starting from the planar C2v structure and including all modes [Fig. 4(c)], a fast step-wise decay of population is seen to S1 and 90% of the population is lost from S2. Only a small population rise is seen in S0, with ∼5% having reached the ground-state after 100 fs. A similar picture is seen starting from the Cs minimum energy structure [Fig. 4(d)]. However, there is more population transfer to the ground-state, with ∼15% transfer to S0 by 100 fs. Simulations including only the key 12 modes [Figs. 4(a) and 4(b)] show very similar population dynamics, indicating that these are the key modes representing the short-time dynamics. Interestingly, the decay to the ground-state is slightly less in the full, 27D calculations. These populations are similar to previous MCTDH dynamics on a simpler vibronic coupling model.29 The system retaining a significant population in S1 on this time-scale is supported by the fact that cyclobutanone is known to fluoresce.52 

FIG. 4.

State populations from the MCTDH calculation on the singlet manifold, vibronic coupling model of cyclobutanone. Calculations (a) and (c) start from the C2v ground-state structure, while (b) and (d) start from Cs. (a) and (b) Only the key 12 modes, while (c) and (d) are the full 27D calculations.

FIG. 4.

State populations from the MCTDH calculation on the singlet manifold, vibronic coupling model of cyclobutanone. Calculations (a) and (c) start from the C2v ground-state structure, while (b) and (d) start from Cs. (a) and (b) Only the key 12 modes, while (c) and (d) are the full 27D calculations.

Close modal

The minimum energy conical intersections (MECIs) between the S2/S1 and S1/S0 states from the vibronic coupling model were obtained by first minimizing the energy gap between the states of interest and then minimizing the energy along the intersection seam using a Lagrange constraint. The geometries in terms of the normal modes are given in the supplementary material. The S2/S1 intersection was found to be at 5.4 units of displacement away from the FC point. Large displacements are made along the ν5, ν16, ν19, ν20, and ν21 modes. The energy of the intersection is at 5.97 eV and is below the FC point, at 6.32 eV. In a recently introduced classification of conical intersection types,53 this is a direct sloped intersection, which provides the step-wise fast transfer observed. The S1/S0 intersection is further away from the FC point, at 11.5 units of displacement at an energy of 4.58 eV, and is thus energetically accessible. However, the displacements required to reach it involve large-scale motions along ν1 and all the B2 vibrations; hence, this intersection is not directly accessed and the transfer to the ground-state is limited.

Absorption spectra calculated from the Fourier transform of the autocorrelation function are shown in Fig. 5. Similar results are obtained with the full and 12D simulations. In order to remove artifacts due to the finite propagation time, T, the autocorrelation function was multiplied by an exponential damping of 150 fs. The spectra are shifted by the ground-state zero-point energy. Spectra for absorption to S1 were also calculated from simulations starting with a vertical excitation to S1.

FIG. 5.

Spectra from ML-MCTDH calculation on the singlet manifold, vibronic coupling model of cyclobutanone. (a) and (c) The S1 spectrum starting from either the (a) C2v or (c) Cs structure. (b) and (d) The S2 spectrum starting from either the (b) C2v or (d) Cs structure.

FIG. 5.

Spectra from ML-MCTDH calculation on the singlet manifold, vibronic coupling model of cyclobutanone. (a) and (c) The S1 spectrum starting from either the (a) C2v or (c) Cs structure. (b) and (d) The S2 spectrum starting from either the (b) C2v or (d) Cs structure.

Close modal

The experimental spectrum shows a broad, featureless band from 3.8 eV to 5.0 eV, with the maximum at around 4.2 eV,20 representing the S1 band. This is a dark state with low intensity. The calculated spectrum from the C2v structure is found to be in the correct energy region. However, the maximum of this band is lower in energy; by 0.4 eV, it is too narrow and too structured. Starting from the Cs structure, the fine structure of the band is much reduced, although the band is too broad and high in energy.

The experimental S2 band is also broad, but with a weak progression of peaks between at 6.0 eV and 6.8 eV, with a maximum at 6.5 eV.20 The calculated spectrum from the C2v structure lies in the correct energy region (around 0.2 eV lower) and has too much structure. The calculated spectrum from the Cs structure is, again, too broad and high in energy.

These results indicate that the spectra are better reproduced by excitation from the C2v structure despite it being a transition state. This is unsurprising due to the low transition barrier, meaning that at room temperature, the ground state wavefunction will not be purely in a ground-state Cs minimum. It will also access the C2v structure.

Finally, the expectation values of the normal modes can be examined to see which modes are excited. These are shown in Sec. S4 of the supplementary material. Starting from the C2v structure, the modes ν5, ν16, ν19, ν20, and ν21 all show large oscillations, particularly ν21 and ν5. All are totally symmetric vibrations, and there is no significant loss of symmetry, or out-of-plane motion, over the 200 fs, although mode ν2 is starting to be activated after 150 fs. Mode ν21, corresponding to the C–O stretch, is seen to extend and oscillates around a positive displacement. Starting from the non-planar Cs structure, the expectation values for the non-symmetric modes ν1 and ν2 now also undergo significant dynamics. Mode ν1 starts with a value around 2 units, and this decreases, showing that the ring becomes more planar over time. ν2, moves to a negative value, showing the oxygen moves further out of plane.

The final step to give a more complete molecular picture of the dynamics, and obtain the GUED signal, was to perform direct dynamics simulations using the DD-vMCG algorithm. This calculates the potential surfaces on the fly and provides a better description of any long range motion than can be provided by the vibronic coupling model used in the MCTDH calculations of Sec. III E. The surfaces were calculated only at the CASSCF level in order to save time. While this will provide less accurate energetics than the CASPT2 calculations used in the model Hamiltonian, the topography of the surfaces and couplings should be good enough to capture the major atomic motions.

The coordinates used for the simulation are the mass-frequency scaled normal modes of the C2v transition state. Points were added to the database keeping the C2v symmetry of the surface by generating appropriate symmetry replicas of each point. An initial calculation of 100 fs was also run starting with the C2v geometry, and 8 GWPs, so as to ensure the symmetry of the initial points in the database.

The sample in the experiments will be cooled to below room temperature, and hence, the main simulations were started from the Cs minimum. The wavepacket was divided into two partitions with the key 12 modes in one partition and the remaining modes in the second. The database was grown by running an initial simulation for 500 fs, with 16 GWPs in the first partition and 8 in the second. A second simulation collecting further points was then run for 500 fs with 32/16 GWPs in the two partitions. Points were added whenever a structure had an atom over 0.2 bohr away from any structure already in the database. The final database had 22 391 structures. A production run was then carried out for 500 fs, again with a 32/16 GWP basis set, utilizing only the points existing in the database, i.e., not collecting new points. The population dynamics of this final simulation is similar to the model Hamiltonian results shown in Fig. 4. However, less crossing out of the S2 state occurs, resulting in 46% of the population in S1, 42% in S2, and 12% in S0.

The MECIs between the S2/S1 and S1/S0 surfaces, provided by the direct dynamics quantum chemistry database, were obtained using an optimization procedure in the same way as those in the vibronic coupling model. The MECI geometries, in both normal mode and Cartesian coordinates, are given in the supplementary material. In both cases, the MECI is slightly closer to the FC point, with the S2/S1 MECI now 3.2 units of displacement away and the S1/S0 MECI 8.3 units away. The energies have also changed, with the S2 MECI moving down to 5.2 eV and the S1/S0 MECI up to 4.90 eV.

The simulation provides 512 trajectories (the centers of the GWP configurations). Using the procedure in Sec. II E, these can then be used to simulate a PDF, as would be produced by a GUED experiment. The modified scattering intensity, the simulation of the original signal, is shown in the supplementary material. The difference PDF as a function of time is shown in Fig. 6(b), with the reference PDF in Fig. 6(a). The reference is the pair distribution function of the initial, Cs minimum energy structure. The peak at 1.5 Å is due to all of the nearest neighbors. The peak at 2.3 Å is due to the distances C1–C4, C2–C3 (both across the ring) as well as C2–O and C3–O. The peak at 3.2 Å is due to C4–O, and the weak peak at 4.3 Å is due to the distance from the protons attached to C4 to O. Atom numbering in this section follows the numbering in Fig. 2.

FIG. 6.

Pair Distribution Function (PDF) from 512 DD-vMCG trajectories calculated on CASSCF potential surfaces. (a) Reference PDF for the initial structure. (b) Difference PDF as a function of time. The blue peaks denote loss with respect to the reference spectrum of (a), while red peaks denote gain.

FIG. 6.

Pair Distribution Function (PDF) from 512 DD-vMCG trajectories calculated on CASSCF potential surfaces. (a) Reference PDF for the initial structure. (b) Difference PDF as a function of time. The blue peaks denote loss with respect to the reference spectrum of (a), while red peaks denote gain.

Close modal

In Fig. 6, loss of intensity is observed in the neighboring atom peak at 1.5 Å with a signal gain around 1.8 Å This is due to the bonds vibrating. The major feature, however, is the strong loss of the 2.3 Å peak and gains at 3 Å and 3.6 Å occurring with a periodicity of 65 fs. The question arises as to what the two new distances are. Plots of the four distances that make up the 2.3 Å peak as a function of time, averaged over the trajectories using the GGP weighting, are shown in Fig. 7. The across ring distance C2–C3 is seen to undergo a small amplitude vibration. The across ring distance C1–C4, however, undergoes a large amplitude vibration, stretching to 2.5 Å, which, along with the C2–O and C3–O distances that extend out to 3 Å, is responsible for the gain just below 3 Å. The gain at 3.5 Å is due to the C4–O vibration and the loss of the original signal masked by the gain at 2.5 Å.

FIG. 7.

Distances between atoms in cyclobutanone from 512 DD-vMCG trajectories calculated on CASSCF potential surfaces, weighted by the gross Gaussian population of each trajectory. Some distances are not included as they are equivalent by symmetry.

FIG. 7.

Distances between atoms in cyclobutanone from 512 DD-vMCG trajectories calculated on CASSCF potential surfaces, weighted by the gross Gaussian population of each trajectory. Some distances are not included as they are equivalent by symmetry.

Close modal

The overall picture, hence, is that the molecule stretches along the C–C–O axis. Examination of the bond lengths shows clearly that the C2–C4 distance changes minimally, and the stretch is dominated by the C–O group moving away from the other carbon atoms. The dynamics divides into two regimes. In the first 300 fs, the vibrations are coherent and stay in a tight grouping, whereas at later times, these distances spread, and in most, the vibrational motion dies away. An exception is the C–O bond, which appears to be increasing in amplitude. A final note on the dynamics is that there is little change in the ring puckering mode, with an average angle ranging from 166° to 173°. The oxygen out-of-plane motion is a little stronger, with the out-of-plane angle ranging from 12° to −7°. In contrast to earlier AIMS simulations starting from S126 and static calculations of the S1 surface,25 there are no signs of ring opening over the 500 fs.

We have performed quantum dynamics simulations to predict the dynamics of cyclobutanone as it will be observed by a gas-phase ultrafast electron diffraction experiment over 500 fs following the initiation of the dynamics by excitation to the S2 electronic state. Cyclobutanone is a non-trivial molecule for simulations. It has recently been shown that the population dynamics for the related cyclopropanone molecule is highly sensitive to the electronic structure method used.54 The ground-state equilibrium structure is a shallow well, and the molecule easily converts between puckered structures through a planar transition state. Consequently, the choice of the initial conditions is also not straightforward as the geometry will be temperature dependent. The potential involvement of triplet states brings a further challenge.

The final results from the simulations indicate that the triplet states are not significantly populated over the first 500 fs. However, it is also found that while the initial relaxation from S2 is fast, taking place in around 50 fs, the subsequent relaxation from S1 is slow. Thus, at later time-scales, there will be crossing to the triplet manifold, and this can result in photo-fragmentation.55 Our simulations also show that in the singlet manifold, over 500 fs, the molecule does little more than vibrate along the central C–C–O axis. The C–O moiety is increasing in energy toward the end of this time, and it is to be expected that the molecule will fragment to C3H6 + C–O at later times. No ring puckering dynamics is observed. The pair distribution function was obtained, which would be measured by the electron scattering. It is seen that this plot is not easy to interpret without the simulations. The simulations give a molecular picture to the pair distances as each peak is due to multiple pairs, as well as overlap. This function, however, clearly reflects the dynamics seen in the simulation.

The simulations are of course not definitive. While a nuclear wavepacket, including quantum effects, was used, the vMCG basis set was not very large, and so this lack of convergence may mean some features are not seen. The potential surfaces used in the final simulations were also only at the CASSCF level, which may not be sufficient for a good description away from equilibrium geometries. The approximations used in the dynamics, e.g., use of the LHA for integrals, and the propagation diabatization procedure may also bring errors. While it is not presently feasible, future work could remove these uncertainties by making it possible to calculate the signal using the full accuracy of the MCTDH method run on the flexible surfaces provided by the DD-vMCG simulations. The integrals required for the GUED signal are not trivial in a grid-based method but are, in principle, possible. In addition, finally, it should be noted that the independent atom model used in calculating the GUED signal may not fully account for changes in the electron density brought about by the non-adiabatic dynamics that couple the electronic motion to the nuclei. A comparison with the experimental result will be a great test of the present capabilities and will show what is still in need of improvement.

The supplementary material includes the coordinates of the ground state structures, vibrational frequencies, cuts through the vibronic coupling model potentials, basis sets used in the ML-MCTDH calculations, information on conical intersections in the singlet manifold, and a plot of the difference scattering spectrum that would be obtained from a GUED experiment. Details of available datasets containing the files from the simulations presented are also given.

This work was funded by the EPSRC under the program Grant COSMOS (No. EP/X026973/1). O.B. also acknowledges UCL for funding. A.F. acknowledges financial support from the Cluster of Excellence “CUI: Advanced Imaging of Matter” of the Deutsche Forschungsgemeinschaft (DFG)—EXC 2056—Project No. 390715994, from the International Max Planck Graduate School for Ultrafast imaging & Structural Dynamics (IMPRS-UFAST) and from the Christiane Nüsslein-Vollhard Foundation.

The authors have no conflicts to disclose.

Olivia Bennett: Conceptualization (equal); Investigation (equal); Methodology (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Antonia Freibert: Investigation (equal); Methodology (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). K. Eryn Spinlove: Conceptualization (equal); Investigation (equal); Methodology (equal); Writing – original draft (equal); Writing – review & editing (equal). Graham A. Worth: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (lead); Investigation (equal); Methodology (equal); Supervision (lead); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available within the article and its supplementary material.

1.
Y.
Liu
,
S. L.
Horton
,
J.
Yang
,
J. P. F.
Nunes
,
X.
Shen
,
T. J. A.
Wolf
,
R.
Forbes
,
C.
Cheng
,
B.
Moore
,
M.
Centurion
et al,
Phys. Rev. X
10
,
021016
(
2020
).
2.
E. G.
Champenois
,
D. M.
Sanchez
,
J.
Yang
,
J. P.
Figueira Nunes
,
A.
Attar
,
M.
Centurion
,
R.
Forbes
,
M.
Guhr
,
K.
Hegazy
,
F.
Ji
et al,
Science
374
,
178
(
2021
).
3.
J.
Cao
,
X.
Wang
, and
D.
Zhong
,
Science
374
,
34
(
2021
).
4.
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
et al,
Nat. Chem.
11
,
504
(
2019
).
5.
J.
Yang
,
X.
Zhu
,
T. J. A.
Wolf
,
Z.
Li
,
J. P. F.
Nunes
,
R.
Coffee
,
J. P.
Cryan
,
M.
Gühr
,
K.
Hegazy
,
T. F.
Heinz
et al,
Science
361
,
64
(
2018
).
6.
D.
Belluš
and
B.
Ernst
,
Angew. Chem., Int. Ed. Engl.
27
,
797
(
1988
).
7.
R. D.
Bach
and
O.
Dmitrenko
,
J. Am. Chem. Soc.
128
,
4598
(
2006
).
8.
B. M.
Trost
, in
Small Ring Compounds in Organic Synthesis I
, edited by
A.
de Meijere
(
Springer
,
Heidelberg, Germany
,
1986
), pp.
3
82
, ISBN: 978-3-540-39769-4.
9.
A.
Lumbroso
,
S.
Catak
,
S.
Sulzer-Mossé
, and
A.
De Mesmaeker
,
Tetrahedron Lett.
56
,
2397
(
2015
).
10.
S.
Izquierdo
,
F.
Rúa
,
A.
Sbai
,
T.
Parella
,
A.
Álvarez-Larena
,
V.
Branchadell
, and
R.
Ortuño
,
J. Org. Chem.
70
,
7963
(
2005
).
11.
E.
Lee-Ruff
and
D.
Wells
,
Nucleosides, Nucleotides Nucleic Acids
27
,
484
(
2008
).
12.
M.
Mihovilovic
,
P.
Kapitan
,
J.
Rydz
,
F.
Rudroff
,
F.
Ogink
, and
M.
Fraaije
,
J. Mol. Catal. B: Enzym.
32
,
135
(
2005
).
14.
M. C.
Pirrung
and
J.
Wang
,
J. Org. Chem.
74
,
2958
(
2009
).
15.
N.
Kishner
,
Zh. Russ. Fiz.-Khim. O-va.
37
,
106
(
1905
).
16.
D.
Seebach
and
A.
Beck
,
Org. Synthetic.
51
,
76
(
1971
).
17.
S. W.
Benson
and
G. B.
Kistiakowsky
,
J. Am. Chem. Soc.
64
,
80
(
1942
).
18.
D. C.
Moule
,
J. Chem. Phys.
64
,
3161
(
1976
).
19.
J. C.
Hemminger
,
H. A. J.
Carless
, and
E. K. C.
Lee
,
J. Am. Chem. Soc.
95
,
682
(
1973
).
20.
E. W.-G.
Diau
,
C.
Kötting
, and
A. H.
Zewail
,
ChemPhysChem
2
,
294
(
2001
).
21.
R. J.
Campbell
,
E. W.
Schlag
, and
B. W.
Ristow
,
J. Am. Chem. Soc.
89
,
5098
(
1967
).
22.
K. Y.
Tang
and
E. K. C.
Lee
,
J. Phys. Chem.
80
,
1833
(
1976
).
23.
J. C.
Hemminger
and
E. K. C.
Lee
,
J. Chem. Phys.
56
,
5284
(
1972
).
24.
M.-H.
Kao
,
R. K.
Venkatraman
,
M. N. R.
Ashfold
, and
A. J.
Orr-Ewing
,
Chem. Sci.
11
,
1991
(
2020
).
25.
S.-H.
Xia
,
X.-Y.
Liu
,
Q.
Fang
, and
G.
Cui
,
J. Phys. Chem. A
119
,
3569
(
2015
).
26.
L.
Liu
and
W.-H.
Fang
,
J. Chem. Phys.
144
,
144317
(
2016
).
27.
T. S.
Kuhlman
,
T. I.
Sølling
, and
K. B.
Møller
,
ChemPhysChem
13
,
820
(
2012
).
28.
T. S.
Kuhlman
,
S. P. A.
Sauer
,
T. I.
Sølling
, and
K. B.
Møller
,
J. Chem. Phys.
137
,
22A522
(
2012
).
29.
T. S.
Kuhlman
,
S. P. A.
Sauer
,
T. I.
Sølling
, and
K. B.
Møller
,
EPJ Web Conf.
41
,
02033
(
2013
).
30.
H.-J.
Werner
,
P. J.
Knowles
,
P.
Celani
,
W.
Györffy
,
A.
Hesselmann
,
D.
Kats
,
G.
Knizia
,
A.
Köhn
,
T.
Korona
,
D.
Kreplin
et al,
MOLPRO, version 2022.1, a package of ab initio programs
,
2022
, see https://www.molpro.net.
31.
H.-J.
Werner
,
P. J.
Knowles
,
G.
Knizia
,
F. R.
Manby
, and
M.
Schütz
,
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
2
,
242
(
2012
).
32.
H.-J.
Werner
,
P. J.
Knowles
,
F. R.
Manby
,
J. A.
Black
,
K.
Doll
,
A.
Heßelmann
,
D.
Kats
,
A.
Köhn
,
T.
Korona
,
D. A.
Kreplin
et al,
J. Chem. Phys.
152
,
144107
(
2020
).
33.
G. A.
Worth
,
K.
Giri
,
G. W.
Richings
,
I.
Burghardt
,
M. H.
Beck
,
A.
Jäckle
, and
H.-D.
Meyer
,
Quantics package, version 2.0 development
,
2023
, see http://www.chem.ucl.ac.uk/quantics.
34.
G. A.
Worth
,
Comput. Phys. Commun.
248
,
107040
(
2020
).
35.
C.
Cattarius
,
A.
Markmann
, and
G. A.
Worth
,
The VCHAM program
,
2007
, see http://www.pci.uni-heidelberg.de/tc/usr/mctdh/.
36.
H.
Köppel
,
W.
Domcke
, and
L. S.
Cederbaum
,
Adv. Chem. Phys.
57
,
59
(
1984
).
37.
G. A.
Worth
and
L. S.
Cederbaum
,
Annu. Rev. Phys. Chem.
55
,
127
(
2004
).
38.
M. H.
Beck
,
A.
Jäckle
,
G. A.
Worth
, and
H.-D.
Meyer
,
Phys. Rep.
324
,
1
(
2000
).
39.
H.-D.
Meyer
,
F.
Gatti
, and
G. A.
Worth
,
Multidimensional Quantum Dynamics: MCTDH Theory and Applications
(
Wiley VCH
,
Weinheim, Germany
,
2009
), ISBN: 9783527320189.
40.
H.
Wang
and
M.
Thoss
,
J. Chem. Phys.
119
,
1289
(
2003
).
41.
U.
Manthe
,
J. Chem. Phys.
128
,
164116
(
2008
).
42.
O.
Vendrell
and
H.-D.
Meyer
,
J. Chem. Phys.
134
,
44135
(
2011
).
43.
D.
Mendive-Tapia
,
T.
Firmino
,
H.-D.
Meyer
, and
F.
Gatti
,
Chem. Phys.
482
,
113
(
2017
).
44.
I.
Burghardt
,
H.-D.
Meyer
, and
L. S.
Cederbaum
,
J. Chem. Phys.
111
,
2927
(
1999
).
45.
G. W.
Richings
,
I.
Polyak
,
K. E.
Spinlove
,
G. A.
Worth
,
I.
Burghardt
, and
B.
Lasorne
,
Int. Rev. Phys. Chem.
34
,
269
(
2015
).
46.
E. J.
Heller
,
J. Chem. Phys.
62
,
1544
(
1975
).
47.
G.
Christopoulou
,
A.
Freibert
, and
G. A.
Worth
,
J. Chem. Phys.
154
,
124127
(
2021
).
48.
M.
Centurion
,
T. J.
Wolf
, and
J.
Yang
,
Annu. Rev. Phys. Chem.
73
,
21
(
2022
).
49.
T. J.
Wolf
and
T. J.
Martinez
,
Diffraction_simulation
,
2021
, see https://github.com/thomasjawolf/diffraction_simulation.
50.
L.
O’Toole
,
P.
Brint
,
C.
Kosmidis
,
G.
Boulakis
, and
P.
Tsekeris
,
J. Chem. Soc., Faraday Trans.
87
,
3343
(
1991
).
51.
C.
Cattarius
,
G. A.
Worth
,
H.-D.
Meyer
, and
L. S.
Cederbaum
,
J. Chem. Phys.
115
,
2088
(
2001
).
52.
E. K. C.
Lee
,
R. G.
Shortridge
, and
C. F.
Rusbult
,
J. Am. Chem. Soc.
93
,
1863
(
1971
).
53.
S.
Gómez
,
E.
Spinlove
, and
G. A.
Worth
,
Phys. Chem. Chem. Phys.
26
,
1829
(
2024
).
54.
J.
Janoš
and
P.
Slavíček
,
J. Chem. Theory Comput.
19
,
8273
(
2023
).
55.
Y.
Chen
and
S.
Ye
,
Int. J. Quantum Chem.
97
,
725
(
2003
).
Published open access through an agreement with JISC Collections

Supplementary Material