Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2024 Feb 5;20(5):2167–2180. doi: 10.1021/acs.jctc.3c01259

Time-Dependent Resonant Inelastic X-ray Scattering of Pyrazine at the Nitrogen K-Edge: A Quantum Dynamics Approach

Antonia Freibert †,‡,*, David Mendive-Tapia , Nils Huse , Oriol Vendrell ‡,*
PMCID: PMC10938531  PMID: 38315564

Abstract

graphic file with name ct3c01259_0008.jpg

We calculate resonant inelastic X-ray scattering spectra of pyrazine at the nitrogen K-edge in the time domain including wavepacket dynamics in both the valence and core-excited state manifolds. Upon resonant excitation, we observe ultrafast non-adiabatic population transfer between core-excited states within the core-hole lifetime, leading to molecular symmetry distortions. Importantly, our time-domain approach inherently contains the ability to manipulate the dynamics of this process by detuning the excitation energy, which effectively shortens the scattering duration. We also explore the impact of pulsed incident X-ray radiation, which provides a foundation for state-of-the-art time-resolved experiments with coherent pulsed light sources.

Introduction

The development of high-brilliance synchroton X-ray radiation sources13 and X-ray free electron lasers47 has enabled techniques that require high photon flux such as resonant inelastic X-ray scattering (RIXS)8 and its extension into the ultrafast time domain.917 RIXS constitutes a Raman scattering process in which the system is resonantly excited into short-lived core-hole states and spontaneously decays back to the electronic ground and excited states.8,18 This technique combines the element specificity of core-level spectroscopy with the ability to reach valence-excited states across a wide spectral range (>20 eV) even in the condensed phase19 and at a spectral resolution that is not limited by the core-hole lifetime broadening, making it a versatile and promising tool to study the local electronic structure in complex molecular systems. RIXS to ground and valence-excited states has already been applied in various areas such as the nature of hydrogen-bond interactions,2022 the photochemistry of transition-metal complexes,2329 ultrafast dynamics of molecules on surfaces,30,31 proton transfer dynamics,13,32,33 or core-excited state dynamics.3436

Within the dipole approximation, RIXS follows electronic symmetry selection rules that offer valuable insights into the symmetry of occupied and unoccupied orbitals in molecules.3740 For instance, by strictly adhering to the dipole selection rules during both the absorption and emission steps, the parity of the system must remain unchanged throughout the RIXS process. However, molecules with equivalent atoms always possess multicenter core orbitals that are delocalized over these atoms, resulting in nearly degenerate core-excited states, commonly referred to as the core-hole localization problem.4146 These states couple vibronically through a non-totally symmetric normal mode, thus corresponding to a symmetry-allowed conical intersection47 and resulting in a final localization of the core holes.18,37,48,49 This dynamical distortion of symmetry can enable transitions that are otherwise electronically symmetry-forbidden38,39,50 and significantly impact the intensity of their RIXS signal.49

One effective method to control or even prevent symmetry breaking in RIXS is to adjust the excitation energy to below the X-ray absorption resonance.51,52 More generally, this detuning of the excitation energy allows for the manipulation of the effective lifetime in the core-excited states and, consequently, the scattering duration.51,53,54 Hence, by varying the detuning, it is possible to control and manipulate dynamical processes within the femtosecond scattering duration, such as dissociation,53,55 vibrational collapse,56,57 or symmetry distortion.36,5860

There are different theoretical approaches to simulating RIXS spectra, which can be broadly classified into time-independent and time-dependent methods. In general, the quantum mechanical description of X-ray Raman scattering is based on second-order time-dependent perturbation theory which gives rise to the well-known Kramers–Heisenberg–Dirac (KHD) expression for the polarizabilty tensor in the frequency domain.6163 Starting from this sum-over-states form, time-independent methods rely on an eigenstate representation of the system, ideally considering all molecular vibronic states that contribute to the resonance effect.64,65 Although various efficient techniques have been developed,6669 as the size and complexity of the molecular system increase, summing over all the eigenstates of the vibrational modes in the molecular excited state becomes rapidly computationally demanding. Therefore, time-dependent strategies based on wave packet propagation on the excited state manifold have been proposed as alternatives, where the scattering amplitude is determined by the half-Fourier transform of the time-dependent overlap between the final and initial vibrational states.63,70 Various levels of theory have been employed, including real-time propagation,71,72 Green’s function techniques,73,74 and methods based on time-dependent density functional theory,75,76 demonstrating the extensive potential and applicability of time-domain formalisms.

In the present work, we examine the RIXS process of pyrazine at the nitrogen K-edge by using a comprehensive time-domain approach. We employ a full quantum mechanical treatment of both, the nuclear and electronic degrees of freedom, under symmetry-selective excitations in a diabatic representation of the Hamiltonian within the multiconfiguration time-dependent Hartree (MCTDH)77 framework. To accurately depict dynamic processes occurring within the ultrashort core-hole lifetime and their manipulation through changes in the excitation frequency, we explicitly incorporate nuclear motion on the core-excited state manifold. Additionally, we explicitly introduce an arbitrary coherent spectral distribution, e.g., an incoming ultrashort X-ray pulse, and investigate how it manifests in the resulting spectra, thereby enabling an optimal interplay between theory and experiment.

Methodology

Model Hamiltonian

The full Hamiltonian H employed in the nuclear quantum dynamics simulations consists of a molecular Hamiltonian Hmol supplemented by the interaction with an external electromagnetic field Hint acting as a time-dependent perturbation

graphic file with name ct3c01259_m001.jpg 1

Following the semi-classical ansatz of ref (78), the coupling to the external field is assumed to be accurately captured by the dipole approximation

graphic file with name ct3c01259_m002.jpg 2

where Inline graphic is the electric field of the photon beam and μαβ is the transition dipole moment between the electronic states |α⟩ and |β⟩, and where the Condon approximation and a single polarization direction are assumed. Within the rotating wave approximation, the electric field is represented by

graphic file with name ct3c01259_m004.jpg 3

with the temporal envelope function Inline graphic of the laser pulse centered at t0 and with carrier frequency ω0. In particular, we refer to a continuous wave (CW) experiment if Inline graphic.

As in ref (78), we decouple the valence and core-excited states,79,80 yielding the following matrix representation of the molecular Hamiltonian

graphic file with name ct3c01259_m007.jpg 4

where Hv and Hc are sub-Hamiltonians acting on the manifolds of the valence and core-excited electronic states, respectively. In both subspaces, a vibronic coupling model up to second order47,48,81 is employed, leading to coupled potential energy surfaces (PESs) in a diabatic representation.82,83 In this framework, each submatrix in eq 4 is expressed as

graphic file with name ct3c01259_m008.jpg 5

where the zeroth-order Hamiltonian is constituted by the ground-state Hamiltonian in the harmonic approximation

graphic file with name ct3c01259_m009.jpg 6

with ωi representing the frequency of mode Qi. The diabatic potential matrix Wx covers all changes in the excited state surfaces compared to the ground state, whose matrix elements are in this work given by

graphic file with name ct3c01259_m010.jpg 7
graphic file with name ct3c01259_m011.jpg 8

with the vertical energy displacement E(α) of the α-th electronic state, the linear intrastate coupling constant κ(α)i related to the gradients of the adiabatic potentials at the Franck–Condon point, the quadratic intrastate coupling constant γ(α)i enclosing changes in the frequencies, and the linear interstate coupling constants λ(αβ)i belonging to the non-adiabatic interaction between the α-th and β-th electronic states under displacements of the Qi mode.

For strongly anharmonic modes, the harmonic expression of the diabatic potentials is replaced by Morse potentials

graphic file with name ct3c01259_m012.jpg 9

where D(α)0 denotes the state-specific dissociation energy, α(α)i defines the curvature of the potential, and Q0 is the equilibrium position.

The parameters required to describe the diabatic molecular Hamiltonian were obtained from electronic structure calculations performed at the coupled cluster singles and doubles (CCSD) level of theory84 and its extensions for excited states using Dunning’s correlation consistent basis set aug-cc-pVDZ.85 Ground-state geometry optimization and normal mode evaluation were computed using the quantum chemistry software package Gaussian.86 Excited-state calculations were performed using the quantum chemistry software package Q-Chem87 where adiabatic energies, energy gradients, and non-adiabatic coupling terms for the valence-excited states were obtained employing the equation-of-motion (EOM-) CCSD approach88 while core-excited state properties were computed using the frozen-core/core–valence-separated (fc-CVS-) EOM-CCSD method.89 The latter two quantities were projected from Cartesian coordinates to the normal mode coordinate representation; for this purpose, the VCHam tools within the Heidelberg MCTDH package were used90

graphic file with name ct3c01259_m013.jpg 10
graphic file with name ct3c01259_m014.jpg 11

where Hel denotes the electronic Hamiltonian. In the case of the most relevant electronic states, the corresponding excited state parameters were refitted through a least-squares procedure applied to a series of ab initio single-point energy calculations along each normal mode.

Derivation of Spectral Properties

The underlying mechanism of RIXS is a two-photon process that first involves interaction with an incident X-ray beam, creating an evolving wavepacket on the core-excited state manifold, followed by spontaneous emission of a photon. The time-dependent picture of RIXS can hence be retrieved by second-order time-dependent perturbation theory for the light–matter interaction.70,91

Assuming a monochromatic and perturbative CW excitation, the RIXS cross-section IRIXS can be obtained from the time evolution of the second-order correction to the system wave function92

graphic file with name ct3c01259_m015.jpg 12

Following Lee and Heller,70,92 the time-dependent second-order perturbation wave function is recursively defined by

graphic file with name ct3c01259_m016.jpg 13

with the first-order wave function

graphic file with name ct3c01259_m017.jpg 14

where |ϕi⟩ ≡ |ϕi(−∞)⟩ is the initial unperturbed wave function with eigenenergy Ei, μI and μS are the transition dipole moment operators along the direction of the incident and the scattering field, εI and εS are monochromatic with frequency Inline graphic and Inline graphic, respectively, and where Inline graphic is defined by Inline graphic. Furthermore, Γc denotes the intrinsic lifetime broadening of the core-excited states. Changing the variables τ = t′ – t″, the first-order wave function can be rewritten as

graphic file with name ct3c01259_m022.jpg 15
graphic file with name ct3c01259_m023.jpg 16

with the Raman wave function Inline graphic given by

graphic file with name ct3c01259_m025.jpg 17

Although the Raman wave function is not an eigenfunction of either Hamiltonian involved, it forms a pseudo-time-independent intermediate state, which contains all dynamical information prior to the scattering event. Finally, inserting eqs 16 and 13 into eq 12 and including the inverse final state lifetime Γf leads to a time-dependent expression for the total RIXS spectrum

graphic file with name ct3c01259_m026.jpg 18

where Inline graphic and Inline graphic is an evolving wavepacket on the ground and valence-excited state manifold. Furthermore, Inline graphic defines the energy loss of the system.

Performing the time-integral in eq 18 yields the frequency-dependent form of the RIXS cross section

graphic file with name ct3c01259_m030.jpg 19

where the scattering amplitude αfi is governed by the well-known KHD formula

graphic file with name ct3c01259_m031.jpg 20

with |i⟩, |n⟩, and |f⟩ being the initial, intermediate, and final vibronic eigenstates involved in the RIXS process with energies Ei, En, and Ef, respectively. Furthermore, the Lorentzian line shape function

graphic file with name ct3c01259_m032.jpg 21

with full width at half-maximum (fwhm) Γ causes a phenomenological broadening. The KHD expression thus provides a static perspective of the RIXS event, where the scattering amplitudes are dependent on the Franck–Condon overlaps of all relevant states as well as the amount of detuning Inline graphic from each intermediate state.

Both, the time-dependent (eq 18) and time-independent (eq 19) expressions for the RIXS cross sections rely on CW conditions and thus neglect any contributions stemming from the spectral content of the incident radiation. In order to incorporate the spectral content required for a finite duration of the incident field, we go beyond the CW picture and assume an Inline graphic-shaped X-ray pulse as given in eq 3, describing a more realistic situation for time-resolved experiments. Without loss of generality, the pulse is centered at t0 = 0 and has the carrier frequency Inline graphic. The first-order wave function now reads

graphic file with name ct3c01259_m036.jpg 22

while the second-order wave function is still recursively defined as given in eq 13. Using the Fourier representation of the envelope function

graphic file with name ct3c01259_m037.jpg 23

and again changing the variables τ = t′ – t″, the first-order wave function can be rewritten as

graphic file with name ct3c01259_m038.jpg

where * denotes the convolution and Inline graphic is given by

graphic file with name ct3c01259_m040.jpg 28

Analogously to eqs 13 and 16, we recursively define Inline graphic by

graphic file with name ct3c01259_m042.jpg 29

yielding the following expression for the second-order wave function

graphic file with name ct3c01259_m043.jpg 30

since the convolution is linear. Within the spontaneous emission process, a transition from every eigenstate component of a wavepacket occurs independently, populating a different final state for large enough t (see Appendix A) and thus

graphic file with name ct3c01259_m044.jpg

and use of eq 12 leads to

graphic file with name ct3c01259_m045.jpg 33

This dependence demonstrates that the impact of a limited pulse duration has two distinct aspects. First, when longitudinally coherent X-ray pulses with a wide spectral range are used, more resonances are encompassed within the coherent pulse excitation. This not only affects the finer vibronic substructure and peak intensity in the final spectra but also decreases the spectral sensitivity of the detuning effect. Additionally, the finite duration of the incident X-ray beam causes a general broadening of the RIXS spectra, given by the convolution with the squared absolute value of the Fourier transform of the incident field envelope function. Note that an incoherent spectral distribution as encountered at synchrotron radiation facilities is not described by a spectral convolution but rather a spectrally weighted average.

Quantum Dynamics

The nuclear time-dependent Schrödinger equation was solved employing the Heidelberg implementation of the multilayer (ML-) MCTDH method.9396 In this approach, MCTDH is recursively applied through different layers of effective modes, within which the equations of motion are determined from the Dirac–Frenkel variational principle.97,98 The layer structure as well as the discrete variable representation, number of grid points, and single-particle function (SPF) basis size were adapted from the ML-2 model in ref (93) without modifications except for the number of electronic states. In particular, this ML ansatz for a full dimensional wave function including all vibrational normal modes was proven to offer a reasonable balance of quality and computational effort. The details are reprinted in Figure 1 for completeness. Core-excited state propagations were run until the pseudo-time-independent intermediate state is reached while dynamics happening in the valence states were run for 100 fs. The data output was written every 0.1 fs, guaranteeing a sufficient time resolution and frequency span for subsequent Fourier transforms.

Figure 1.

Figure 1

Tree structure of the ML-MCTDH simulations, including all 24 normal modes. The maximum depth of the tree is five layers where the first one separates the 24 vibrational coordinates and the discrete electronic degree of freedom in the single-set formulation. The circles represent the node in the layer structure, and the number n of SPFs (blue) is given next to the link lines. The last layer comprises the vibrational normal modes, where the number N of primitive functions (red) is used to represent the grid. All values are taken from ref (93).

Results and Discussion

Parametrization of the Model Hamiltonian

Pyrazine (C4H4N2) belongs to point group D2h at neutral ground-state equilibrium geometry and possesses 24 vibrational normal modes. The vibrational frequencies computed at the CCSD/aug-cc-pVDZ level of theory are listed in Table 1 along with their symmetry and a comparison to the experimental data.

Table 1. Harmonic Ground-State Vibrational Frequencies (in cm–1) Obtained at the CCSD/aug-cc-pVDZ Level along with the Experimental Data99a.

mode Wilson symmetry cm–1 Exp.
ν1 ν16a au 350 341
ν2 ν16b b3u 425 420
ν3 ν6a ag 604 596
ν4 ν6b b3g 709 704
ν5 ν4 b2g 737 756
ν6 ν11 b3u 805 785
ν7 ν10a b1g 943 919
ν8 ν5 b2g 950 983
ν9 ν17a au 980 960
ν10 ν12 b1u 1033 1021
ν11 ν1 ag 1038 1015
ν12 ν18b b2u 1091 1063
ν13 ν14 b2u 1146 1149
ν14 ν18a b1u 1162 1136
ν15 ν9a ag 1253 1230
ν16 ν3 b3g 1367 1346
ν17 ν19b b2u 1441 1416
ν18 ν19a b1u 1518 1484
ν19 ν8b b3g 1595 1525
ν20 ν8a ag 1650 1582
ν21 ν7b b3u 3196 3040
ν22 ν13 b1u 3197 3012
ν23 ν20b b2u 3213 3063
ν24 ν2 ag 3218 3055
a

The modes are labeled by ascending frequency and compared to Wilson’s notation.100

The diabatic Hamiltonian Hmol comprises two sub-Hamiltonian Hv and Hc, accounting for the valence- and core-excited states dynamics, respectively. Hc involves two core-excited states that are nearly degenerate at the Franck–Condon point. In the vicinity of that point, the next energetically higher electronic states are well separated and can hence be disregarded. The four vibrational modes with symmetry B1u are potential candidates to vibronically couple the two nearly degenerate core-excited states where the conical intersection formed along these modes takes place exactly at the Franck–Condon point. Cuts of the PESs along the two strongest coupling modes, ν10(B1u) and ν18(B1u), are shown in Figure 2. The lifetime of the core-excited states is assumed to be 8 fs101 captured by an imaginary energy term in the core Hamiltonian Hc.

Figure 2.

Figure 2

Cuts through the diabatic PESs along the ν10(b1u) and ν18(b1u) vibrational normal modes that mainly drive the symmetry breaking in the core-excited states. Label and symmetry for each state can be found in Table 2. The adiabatic energies (black points) are obtained from ab initio calculations using the (fc-CVS-) EOM-CCSD/aug-cc-pVDZ method.

Since RIXS has the ability to reach higher-lying valence-excited states, we include 20 electronic states in Hv. These states consist of the ground state and the 19 energetically lowest valence-excited singlet states, covering an approximate spectral range of 10 eV. The computed vertical excitation energies and symmetries for all electronic states are listed in Table 2. It is worth noting that the ordering of very closely lying states can differ between different levels of theory.102,103 This concerns not only S2(B2u) and S3(Au) but also the region with a high density of states. To maintain a consistent model within this study, we treat every valence-excited state at the same level of theory without any adjustment of the vertical excitation energy.

Table 2. State Symmetries and Vertical Excitation Energies E(α) (in eV) at the Franck–Condon Point of All Electronic States Considered in This Worka.

state symmetry E(α)
S0 Ag 0.00
S1 B3u 4.32
S2 B2u 5.07
S3 Au 5.13
S4 B2g 6.01
S5 Ag 6.68
S6 B1u 6.91
S7 B1g 7.02
S8 B1g 7.11
S9 B2u 7.28
S10 B1u 7.47
S11 B3u 7.66
S12 B2u 7.94
S13 B3g 8.00
S14 Ag 8.04
S15 Au 8.12
S16 B1u 8.14
S17 B1g 8.35
S18 Au 8.36
S19 B2g 8.53
X1 B2g 402.30
X2 B3u 402.30
a

Valence-excited state energies were evaluated using EOM-CCSD while core-excited state energies were obtained using (fc-CVS-) EOM-CCSD. In both cases, the aug-cc-pVDZ basis set was used.

Both sub-Hamiltonians, Hv and Hc, were then approximated by a vibronic coupling model including all vibrational degrees of freedom. The highest frequency mode ν24 was fitted by Morse potentials to account for anharmonicity. Since the density of valence-excited states above 6 eV increases rapidly, we used the linear intra- and interstate coupling constants as obtained from EOM-CCSD/aug-cc-pvDZ calculations at the Franck–Condon point to parametrize these electronic states (S5 – S19). The linear coupling parameters, which mainly drive the core-excited state dynamics at short times, are listed in Table 3. A complete list of all parameters is further provided in the Supporting Information.

Table 3. Linear Intra- and Interstate Coupling Constants κ(n)i and λ(nm)i, Respectively, for the Core-Excited States, Xn, Obtained in This Work.

  κ3 κ11 κ15 κ20
X1 0.02738 –0.04034 0.05568 0.10433
X2 0.02627 –0.04050 0.05619 0.10457
  λ10 λ14 λ18 κ22
(X1, X2) 0.08680 0.01326 0.09910 0.03014

The transition dipole moments and oscillator strengths for each pair of valence- and core-excited states were computed at the Franck–Condon point using the same level of theory. We neglect transitions where the transition dipole moment was below 0.01 and thus less than 10% of the strongest core–valence transition. The corresponding transition dipole moments are listed in Table 4.

Table 4. Transition Dipole Moments μfi between States i and f Obtained from fc-CVS-EOM-CCSD/aug-cc-pVDZ Calculations.

state transition μfi
X2 ← S0 0.10
X1 ← S1 0.06
X2 ← S4 0.06
X1 ← S6 0.02
X1 ← S16 0.04
X1 ← S18 0.04

Symmetry Distortion Induced by Core-Excited State Dynamics

We consider the scattering event initiated by the resonant absorption of an X-ray photon, which promotes an electron from a nitrogen (N-)1s orbital to the lowest unoccupied molecular orbital (LUMO). Pyrazine contains two equivalent N atoms, so a linear combination of both N-1s orbitals builds symmetric and antisymmetric core-orbitals that are nearly degenerate. Excitation from these core-orbitals leads in turn to two nearly degenerate core-excited states, X1(B2g) and X2(B3u). Although only transitions from the antisymmetric core orbitals are allowed by symmetry, both core-excited states must be considered to account for nonadiabatic transitions induced by vibronic coupling. Figure 3 shows the static X-ray absorption spectrum, resulting from an electronic transition between the antisymmetric core orbital and the LUMO. The absorption spectrum was obtained from the Fourier transform of the autocorrelation function of the dipole-operated ground state, thus equivalent to a δ-pulse excitation.92,104,105

Figure 3.

Figure 3

Computed static X-ray absorption spectrum of pyrazine at the nitrogen K-edge. The vertical lines indicate the excitation energies used to simulate the RIXS spectra. Here, the energy related to the dashed black line corresponds to the vertical excitation energy of X2 used to calculate the RIXS spectrum shown in this section. The energies indicated by the colored lines relate to the detuning effect discussed in the next section.

The related diabatic population transfer is illustrated in Figure 4. After an instantaneous, vertical excitation at time 0, the bright X2(B3u) state rapidly depopulates into the dark X1(B2g) state, leading to an almost equal distribution of population between the two states within the short core-hole lifetime. This ultrafast, nonadiabatic population transfer can be traced back to the symmetry-allowed conical intersection between these states, which is formed in the immediate vicinity of the Franck–Condon point (see Figure 2) and primarily driven by the two asymmetric normal modes ν10 and ν18. These dynamics result in a final localization of the core-hole.

Figure 4.

Figure 4

Core-excited state populations of the diabatic X1(B2g) (green) and X2(B3u) (magenta) states starting from X2(B3u) at the Franck–Condon point. The solid lines present the total diabatic state population including the core-hole decay while the dashed lines correspond to the normalized diabatic state population.

The impact of symmetry breaking induced by core-excited state dynamics is clearly seen in the RIXS spectrum of pyrazine at the nitrogen K-edge presented in Figure 5. The spectrum was obtained using eq 18, assuming a monochromatic X-ray photon beam with carrier frequency equal to the vertical excitation energy from the ground state to the excited state X2. It displays five prominent bands arising from six electronic transitions. The asymmetric shape of the elastic peak already suggests that the system undergoes nuclear displacements on the core-excited state manifold. As expected from the transition dipole moments (see Table 4), the elastic peak is the dominant feature of the spectrum but without obscuring any contributions from inelastic transition due to the significant energy gap between the ground and valence-excited states. Moreover, the spectrum reveals four bands of inelastic electronic transitions located at approximately 4.5, 6.0, 7.0, 8.1, and 8.6 eV where the latter two exhibit a large spectral overlap and are thus barely distinguishable in the overall spectrum. While the emission band located at 6.0 eV is attributable to the electronic X2 → S4 transition, the other four bands originate from population of the optically dark X1(B2g) state which can only be accessed through nonadiabatic population transfer from the bright X2(B3u) state. Consequently, the symmetry breaking caused by ultrafast core-excited state dynamics allows for four additional electronic transitions that would otherwise be forbidden for two consecutive dipole transitions (i.e., by quadrupole selection rules) within this spectral range.

Figure 5.

Figure 5

Simulated RIXS spectrum of pyrazine at the nitrogen K-edge. The contributions of the transitions to S0, S1, S4, S6, S16, and S18 to the total spectrum are highlighted in purple, blue, green, cyan, red, and orange, respectively. The elastic peak (purple) is downscaled to 25% for better visualization. Before performing the Fourier transformation in eq 18, the window function Inline graphic was applied to reduce the Gibbs phenomenon. Here, we assume a damping time of T = 30 fs in order to include broadening caused by dephasing mechanisms.

Dynamical Control by Detuning

Dynamical processes in the core-excited states can be controlled by detuning the frequency of the incident radiation that is used to prepare the intermediate state of the system. The effective scattering duration in RIXS is determined by the time the wavepacket spends in the intermediate core-excited states, i.e., the time interval [0, τ] that contributes to the Raman wave function (eq 17). Due to the uncertainty relation for the energy, the propagation time shortens the further the excitation energy is from resonance and is given by106

graphic file with name ct3c01259_m046.jpg 34

where Ω is the amount of detuning and Γc inversely determines the core-hole lifetime.

The dependence of the RIXS signal for pyrazine on the incoming photon frequency, ωI, is shown in Figure 6. The four inelastic emission channels stemming from the dark X1(B2g) state exhibit a strongly detuning-dependent intensity. In particular, when far from resonance, i.e., for incoming excitation energies below ∼402.0 eV, these bands almost disappear. Moreover, in this region, the elastic peak exhibits a symmetric Lorentzian line shape indicative of a single transition. Both observations show that the effective scattering duration becomes so short for large detuning that dynamical effects are negligible. In contrast, for excitation energies greater than 402.0 eV, the signature of vibrational progressions is apparent on the positive energy side of the elastic peak, and inelastic Raman transitions stemming from both core-excited states, X1 and X2, are prominent. Their shapes differ for different excitation energies, reflecting the nuclear motion in the core-excited states. In particular, for excitation energies greater than the vertical excitation energy, a rather broad and asymmetrical shape can be observed for each loss peak.

Figure 6.

Figure 6

Dependence of the RIXS profile on the incident photon energy. All spectra are independently normalized to their respective elastic peak height before rescaling the signals in the gray spectral region to 25%. The energies of the incident radiation used for these calculation are also highlighted in the photoabsorption band, as shown in Figure 3.

Spectral Distribution due to Finite Pulse Duration

While the previous sections are based on light–matter interaction with monochromatic plane X-ray waves, in the following section, we investigate the influence of the incident X-ray spectrum on the resolution of the RIXS spectra. The RIXS cross sections are calculated using eq 33 where the external electric field was given by a normalized Gaussian-shaped X-ray pulse, i.e., the pulse envelope function is

graphic file with name ct3c01259_m047.jpg 35

where t0 denotes the pulse center and the standard deviation σ is linked to the temporal fwhm duration Ft by Inline graphic. According to eq 33, a spectral broadening of the RIXS spectrum is caused by the convolution with the square of the absolute value of the Fourier transform of the envelope function. For an envelope function of the form in eq 35, this is again a Gaussian function

graphic file with name ct3c01259_m049.jpg 36

with spectral fwhm Inline graphic. In particular, the fwhm ratio is here Ft·Fω ≈ 3.921.

Due to the reciprocal connection between temporal and spectral width, shorter pulses create spectrally broader wavepackets and vice versa, leading to different signatures of the RIXS signal. This aspect is described by the term

graphic file with name ct3c01259_m051.jpg 37

in eq 33 and can be considered independently from the general broadening of the signal described by the convolution with Inline graphic although both aspects have the same origin. Figure 7 illustrates both effects on the RIXS spectra of pyrazine for different incoming X-ray pulses where the temporal duration was varied, but the carrier frequency of the pulse was kept fixed at 402.3 eV. The spectra shown in blue are evaluated in accordance with eq 37, while the orange shadowed spectra represent the total cross section as obtained from eq 33.

Figure 7.

Figure 7

Dependence of RIXS signals on the duration of the incident radiation field. Each spectrum was derived using eq 33 (orange shadow) and eq 37 (blue line), where a Gaussian X-ray pulse was used to trigger the core-excited state dynamics. The carrier frequency was kept fixed while the temporal standard deviation varies from 1 to 8 fs (equivalent to Γc/8 to Γc). The same window function as that in Figure 5 was applied before performing the Fourier transformation.

In general, the different RIXS spectra are in good agreement. However, the relative intensities vary depending on the length of the X-ray pulse. Specifically, emission channels starting from the dark X1 at around 6.0, 7.0, and 8.4 eV appear to be more likely for shorter pulse duration. Furthermore, the shape of the spectral lines changes slightly depending on the distribution of the field. This is evident in the X1 → S1 emission band, where distinct progressions become more prominent with longer pulse duration.

While for pulses with FWHM of about a third of the dephasing time of the valence excitations or longer, the broadening resulting from the spectral distribution of the field is less significant compared to other dephasing effects, shorter pulses cause a more pronounced broadening of the signal, as shown by the orange shadow. This effect should hence be considered for optimal spectral and temporal resolution as well as the detuning capability. Especially for very short pulses, the finer details of the vibronic structure in the emission bands are almost completely suppressed, which can complicate the analysis of core-excited state dynamics when the effectiveness of detuning as a control mechanism is also reduced.

Conclusions

In this study, we benchmarked a fully time-dependent approach within the MCTDH framework to simulate RIXS spectra. We used pyrazine as a model system to examine this resonant scattering process at the nitrogen K-edge, considering wavepacket dynamics in all of the electronic states of the model. Our findings reveal ultrafast symmetry breaking of the molecule in the core-excited states, leading to a significant alteration of the symmetry selection rules. This behavior underscores the crucial role of non-adiabatic nuclear core-excited state dynamics for the correct description of RIXS signals. Furthermore, we explored the detuning dependence of the RIXS spectra in our time-dependent approach. Our results highlight the importance of detuning as a control mechanism for ultrafast dynamical processes, which allows for predicting not only the intensity of both inelastic and elastic emission channels but also the vibrational substructure of individual electronic transitions. Additionally, we investigated how the duration of a coherent X-ray pulse affects the RIXS spectrum.

In summary, we have described RIXS as a fully time-dependent process including electronic and nuclear dynamics in all core- and valence-excited states of the model and with a temporally arbitrary incident X-ray field. Our work demonstrates how the spectral distribution of the X-ray pulse impacts the individual RIXS pathways, and how it leads to a general broadening effect of the loss spectra. This research provides insights into the intricacies of inelastic X-ray scattering in the presence of core-hole symmetry breaking, under detuning, and for excitations with coherent X-ray pulses on the order of or shorter than the core-hole lifetime. Furthermore, our time-dependent simulation approach can be used for modeling ultrafast RIXS experiments at pulsed X-ray sources such as free-electron lasers.

Acknowledgments

A.F. acknowledges financial support from the International Max Planck Graduate School for Ultrafast imaging & Structural Dynamics (IMPRS-UFAST) and from the Christiane-Nüsslein-Vollhard-Foundation. O.V. acknowledges financial support from the German Science Foundation (DFG) through the project number 493826649. This work is supported by the Cluster of Excellence ‘CUI: Advanced Imaging of Matter’ of the Deutsche Forschungsgemeinschaft (DFG)—EXC 2056—project ID 390715994 (A.F. and N.H.).

Appendix A

Emission from a Wavepacket

In order to derive an expression for the probability that a photon with energy ℏω is emitted from a wavepacket, we consider a system containing two electronic states. The total Hamiltonian reads

graphic file with name ct3c01259_m053.jpg 38
graphic file with name ct3c01259_m054.jpg 39

where the Hamiltonian Inline graphic describes the unperturbed system and the interaction Hamiltonian Inline graphic is treated as perturbation. Here, the molecular Hamiltonian Inline graphic of the two-state system is given by

graphic file with name ct3c01259_m058.jpg 40

where En and |n⟩, n ∈ {i, f}, are the eigenenergies and eigenstates, respectively, of the initial and final electronic state. The light Hamiltonian Inline graphic is further defined through

graphic file with name ct3c01259_m060.jpg 41

with the vacuum state |vac⟩ where we ignore directions and polarizations for simplicity. For an interaction operator Inline graphic between the molecule and photon, the light–matter interaction term can be written as

graphic file with name ct3c01259_m062.jpg 42
graphic file with name ct3c01259_m063.jpg 43

Assuming a wavepacket has been initially created as a superposition of the initial electronic state, i.e.,

graphic file with name ct3c01259_m064.jpg 44

the transition rate Γω of emitting photon with energy ℏω is defined by

graphic file with name ct3c01259_m065.jpg 45

where the transition probability Pω(t)≔|⟨ω|ψ(t)⟩|2 can be evaluated using first-order perturbation theory

graphic file with name ct3c01259_m066.jpg 46

where |ψ(0)(0)⟩ = |ψ(0)⟩ and |ψ(1)(0)⟩ = 0. Considering only induced transitions between two different states, the probability then reads

graphic file with name ct3c01259_m067.jpg 47

where the first-order correction ψ(1) to the wavefunction of the system ψ is defined by

graphic file with name ct3c01259_m068.jpg 48

with

graphic file with name ct3c01259_m069.jpg 49
graphic file with name ct3c01259_m070.jpg 50
graphic file with name ct3c01259_m071.jpg 51
graphic file with name ct3c01259_m072.jpg 52

As ∫dω⟨ω′|ω⟩ = ∫dωδ(ω′ – ω) = 1, it follows

graphic file with name ct3c01259_m073.jpg 53
graphic file with name ct3c01259_m074.jpg 54
graphic file with name ct3c01259_m075.jpg 55
graphic file with name ct3c01259_m076.jpg 56

where Ωif≔ωi – ωf. Hence, the emission probability per unit time in first-order correction is given by

graphic file with name ct3c01259_m077.jpg 57
graphic file with name ct3c01259_m078.jpg 58
graphic file with name ct3c01259_m079.jpg 59

For Ωif ≠ ω and Ωjf ≠ ω, it holds

graphic file with name ct3c01259_m080.jpg 60

and in the same way

graphic file with name ct3c01259_m081.jpg 61

Therefore, it follows for the product of a zero and bounded sequence

graphic file with name ct3c01259_m082.jpg 62

and the transition rate Γω thus vanishes for Ωif, Ωjf ≠ ω.

Let now Ωif ≠ ω but Ωjf = ω. Applying the rule of l’Hopital

graphic file with name ct3c01259_m083.jpg 63

we can show that the fraction above is purely imaginary. Hence, it follows

graphic file with name ct3c01259_m084.jpg 64
graphic file with name ct3c01259_m085.jpg 65
graphic file with name ct3c01259_m086.jpg 66
graphic file with name ct3c01259_m087.jpg 67
graphic file with name ct3c01259_m088.jpg 68

and thus again Γω → 0 for t → ∞. Analogously, Γω is zero for Ωif = ω and Ωjf ≠ ω. Hence, only the case Ωif = ω and Ωjf = ω, so in particular i = j, contributes to the transition rate. Starting from eq 59, we get

graphic file with name ct3c01259_m089.jpg 69
graphic file with name ct3c01259_m090.jpg 70
graphic file with name ct3c01259_m091.jpg 71
graphic file with name ct3c01259_m092.jpg 72
graphic file with name ct3c01259_m093.jpg 73
graphic file with name ct3c01259_m094.jpg 74

where |iγ(t)⟩ denotes the propagation of the projected eigenstate |i⟩ onto the final excited state manifold. So, every eigenstate from the wavepacket emits independently, where the contribution of a certain eigenstate is determined by the evolution of the projected wavepacket |iγ(t)⟩ weighted by the initial population of state |i⟩. In particular, the spontaneous emission from a wavepacket is incoherent, in the sense that every transition from this wavepacket populates a different and hence orthogonal final state of the material system and the electromagnetic field.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.jctc.3c01259.

  • On-diagonal linear intrastate coupling constants κ(n)i, off-diagonal linear interstate coupling constants λ(nm)i, and on-diagonal bilinear (quadratic) coupling constants γ(n)i and the parameters concerning the Morse potential (PDF)

The authors declare no competing financial interest.

Supplementary Material

ct3c01259_si_001.pdf (142.5KB, pdf)

References

  1. Tavares P. F.; Leemann S. C.; Sjöström M.; Andersson Å. The MAXIV storage ring project. J. Synchrotron Radiat. 2014, 21, 862–877. 10.1107/S1600577514011503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Willmott P.An Introduction to Synchrotron Radiation: Techniques and Applications; Wiley: Hoboken, 2019. [Google Scholar]
  3. Willmott P. R.. In Magnetism and Accelerator-Based Light Sources; Bulou H., Joly L., Mariot J.-M., Scheurer F., Eds.; Springer International Publishing: Cham, 2021, pp 1–37. [Google Scholar]
  4. McNeil B. W. J.; Thompson N. R. X-ray free-electron lasers. Nat. Photonics 2010, 4, 814–821. 10.1038/nphoton.2010.239. [DOI] [Google Scholar]
  5. Pellegrini C.; Marinelli A.; Reiche S. The physics of x-ray free-electron lasers. Rev. Mod. Phys. 2016, 88, 015006. 10.1103/RevModPhys.88.015006. [DOI] [Google Scholar]
  6. Zhao Z.; Wang D.; Gu Q.; Yin L.; Gu M.; Leng Y.; Liu B. Status of the SXFEL Facility. Appl. Sci. 2017, 7, 607. 10.3390/app7060607. [DOI] [Google Scholar]
  7. Seddon E. A.; Clarke J. A.; Dunning D. J.; Masciovecchio C.; Milne C. J.; Parmigiani F.; Rugg D.; Spence J. C. H.; Thompson N. R.; Ueda K.; Vinko S. M.; Wark J. S.; Wurth W. Short-wavelength free-electron laser sources and science: a review. Rep. Prog. Phys. 2017, 80, 115901. 10.1088/1361-6633/aa7cca. [DOI] [PubMed] [Google Scholar]
  8. Ament L. J. P.; van Veenendaal M.; Devereaux T. P.; Hill J. P.; van den Brink J. Resonant inelastic X-ray scattering studies of elementary excitations. Rev. Mod. Phys. 2011, 83, 705–767. 10.1103/RevModPhys.83.705. [DOI] [Google Scholar]
  9. Zhang Y.; Biggs J. D.; Healion D.; Govind N.; Mukamel S. Core and valence excitations in resonant X-ray spectroscopy using restricted excitation window time-dependent density functional theory. J. Chem. Phys. 2012, 137, 194306. 10.1063/1.4766356. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Beye M.; Wernet P.; Schüßler-Langeheine C.; Föhlisch A. Time resolved resonant inelastic X-ray scattering: A supreme tool to understand dynamics in solids and molecules. J. Electron Spectrosc. Relat. Phenom. 2013, 188, 172–182. 10.1016/j.elspec.2013.04.013. [DOI] [Google Scholar]; Progress in Resonant Inelastic X-Ray Scattering
  11. Mukamel S.; Healion D.; Zhang Y.; Biggs J. D. Multidimensional Attosecond Resonant X-Ray Spectroscopy of Molecules: Lessons from the Optical Regime. Annu. Rev. Phys. Chem. 2013, 64, 101–127. 10.1146/annurev-physchem-040412-110021. [DOI] [PMC free article] [PubMed] [Google Scholar]; PMID: 23245522
  12. Wernet P.; Kunnus K.; Josefsson I.; Rajkovic I.; Quevedo W.; Beye M.; Schreck S.; Grübel S.; Scholz M.; Nordlund D.; Zhang W.; Hartsock R. W.; Schlotter W. F.; Turner J. J.; Kennedy B.; Hennies F.; de Groot F. M. F.; Gaffney K. J.; Techert S.; Odelius M.; Föhlisch A. Orbital-specific mapping of the ligand exchange dynamics of Fe(CO)5 in solution. Nature 2015, 520, 78–81. 10.1038/nature14296. [DOI] [PubMed] [Google Scholar]
  13. Eckert S.; Norell J.; Miedema P. S.; Beye M.; Fondell M.; Quevedo W.; Kennedy B.; Hantschmann M.; Pietzsch A.; Van Kuiken B. E.; Ross M.; Minitti M. P.; Moeller S. P.; Schlotter W. F.; Khalil M.; Odelius M.; Föhlisch A. Ultrafast Independent N-H and N-C Bond Deformation Investigated with Resonant Inelastic X-Ray Scattering. Angew. Chem., Int. Ed. 2017, 56, 6088–6092. 10.1002/anie.201700239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Jay R. M.; Norell J.; Eckert S.; Hantschmann M.; Beye M.; Kennedy B.; Quevedo W.; Schlotter W. F.; Dakovski G. L.; Minitti M. P.; Hoffmann M. C.; Mitra A.; Moeller S. P.; Nordlund D.; Zhang W.; Liang H. W.; Kunnus K.; Kubiček K.; Techert S. A.; Lundberg M.; Wernet P.; Gaffney K.; Odelius M.; Föhlisch A. Disentangling Transient Charge Density and Metal–Ligand Covalency in Photoexcited Ferricyanide with Femtosecond Resonant Inelastic Soft X-ray Scattering. J. Phys. Chem. Lett. 2018, 9, 3538–3543. 10.1021/acs.jpclett.8b01429. [DOI] [PubMed] [Google Scholar]; PMID: 29888918
  15. Lu H.; Gauthier A.; Hepting M.; Tremsin A. S.; Reid A. H.; Kirchmann P. S.; Shen Z. X.; Devereaux T. P.; Shao Y. C.; Feng X.; Coslovich G.; Hussain Z.; Dakovski G. L.; Chuang Y. D.; Lee W. S. Time-resolved RIXS experiment with pulse-by-pulse parallel readout data collection using X-ray free electron laser. Sci. Rep. 2020, 10, 22226. 10.1038/s41598-020-79210-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Chergui M.; Beye M.; Mukamel S.; Svetina C.; Masciovecchio C. Progress and prospects in nonlinear extreme-ultraviolet and X-ray optics and spectroscopy. Nat. Rev. Phys. 2023, 5, 578–596. 10.1038/s42254-023-00643-7. [DOI] [Google Scholar]
  17. Monney C.; Patthey L.; Razzoli E.; Schmitt T. Static and time-resolved resonant inelastic X-ray scattering: Recent results and future prospects. X-Ray Spectrom. 2023, 52, 216–225. 10.1002/xrs.3311. [DOI] [Google Scholar]
  18. Gel’mukhanov F.; Ågren H. Resonant X-ray Raman scattering. Phys. Rep. 1999, 312, 87–330. 10.1016/S0370-1573(99)00003-4. [DOI] [Google Scholar]
  19. Ochmann M.; Vaz da Cruz V.; Eckert S.; Huse N.; Föhlisch A. R-Group stabilization in methylated formamides observed by resonant inelastic X-ray scattering. Chem. Commun. 2022, 58, 8834–8837. 10.1039/D2CC00053A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Harada Y.; Takeuchi T.; Kino H.; Fukushima A.; Takakura K.; Hieda K.; Nakao A.; Shin S.; Fukuyama H. Electronic Structure of DNA Nucleobases and Their Dinucleotides Explored by Soft X-ray Spectroscopy. J. Phys. Chem. A 2006, 110, 13227–13231. 10.1021/jp062720j. [DOI] [PubMed] [Google Scholar]; PMID: 17149838
  21. Weinhardt L.; Ertan E.; Iannuzzi M.; Weigand M.; Fuchs O.; Bär M.; Blum M.; Denlinger J. D.; Yang W.; Umbach E.; Odelius M.; Heske C. Probing hydrogen bonding orbitals: resonant inelastic soft X-ray scattering of aqueous NH3. Phys. Chem. Chem. Phys. 2015, 17, 27145–27153. 10.1039/C5CP04898B. [DOI] [PubMed] [Google Scholar]
  22. Hussain A.; Huse N.; Vendrell O. Sensitivity of core-level spectroscopy to electrostatic environments of nitrile groups: An ab initio study. Struct. Dyn. 2017, 4, 054102. 10.1063/1.5003404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. van Schooneveld M. M.; Gosselink R. W.; Eggenhuisen T. M.; Al Samarai M.; Monney C.; Zhou K. J.; Schmitt T.; de Groot F. M. F. A Multispectroscopic Study of 3 d Orbitals in Cobalt Carboxylates: The High Sensitivity of 2p3d Resonant X-ray Emission Spectroscopy to the Ligand Field. Angew. Chem., Int. Ed. 2013, 52, 1170–1174. 10.1002/anie.201204855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Wernet P.; Kunnus K.; Schreck S.; Quevedo W.; Kurian R.; Techert S.; de Groot F. M. F.; Odelius M.; Föhlisch A. Dissecting Local Atomic and Intermolecular Interactions of Transition-Metal Ions in Solution with Selective X-ray Spectroscopy. J. Phys. Chem. Lett. 2012, 3, 3448–3453. 10.1021/jz301486u. [DOI] [PubMed] [Google Scholar]; PMID: 26290971
  25. Kunnus K.; Zhang W.; Delcey M. G.; Pinjari R. V.; Miedema P. S.; Schreck S.; Quevedo W.; Schröder H.; Föhlisch A.; Gaffney K. J.; Lundberg M.; Odelius M.; Wernet P. Viewing the Valence Electronic Structure of Ferric and Ferrous Hexacyanide in Solution from the Fe and Cyanide Perspectives. J. Phys. Chem. B 2016, 120, 7182–7194. 10.1021/acs.jpcb.6b04751. [DOI] [PubMed] [Google Scholar]; PMID: 27380541
  26. Hahn A. W.; Van Kuiken B. E.; Chilkuri V. G.; Levin N.; Bill E.; Weyhermüller T.; Nicolaou A.; Miyawaki J.; Harada Y.; DeBeer S. Probing the Valence Electronic Structure of Low-Spin Ferrous and Ferric Complexes Using 2p3d Resonant Inelastic X-ray Scattering (RIXS). Inorg. Chem. 2018, 57, 9515–9530. 10.1021/acs.inorgchem.8b01550. [DOI] [PubMed] [Google Scholar]; PMID: 30044087
  27. Jay R. M.; Eckert S.; Fondell M.; Miedema P. S.; Norell J.; Pietzsch A.; Quevedo W.; Niskanen J.; Kunnus K.; Föhlisch A. The nature of frontier orbitals under systematic ligand exchange in (pseudo-)octahedral Fe(II) complexes. Phys. Chem. Chem. Phys. 2018, 20, 27745–27751. 10.1039/C8CP04341H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Jay R. M.; Eckert S.; Van Kuiken B. E.; Ochmann M.; Hantschmann M.; Cordones A. A.; Cho H.; Hong K.; Ma R.; Lee J. H.; Dakovski G. L.; Turner J. J.; Minitti M. P.; Quevedo W.; Pietzsch A.; Beye M.; Kim T. K.; Schoenlein R. W.; Wernet P.; Föhlisch A.; Huse N. Following Metal-to-Ligand Charge-Transfer Dynamics with Ligand and Spin Specificity Using Femtosecond Resonant Inelastic X-ray Scattering at the Nitrogen K-Edge. J. Phys. Chem. Lett. 2021, 12, 6676–6683. 10.1021/acs.jpclett.1c01401. [DOI] [PMC free article] [PubMed] [Google Scholar]; PMID: 34260255
  29. Biasin E.; Nascimento D. R.; Poulter B. I.; Abraham B.; Kunnus K.; Garcia-Esparza A. T.; Nowak S. H.; Kroll T.; Schoenlein R. W.; Alonso-Mori R.; Khalil M.; Govind N.; Sokaras D. Revealing the bonding of solvated Ru complexes with valence-to-core resonant inelastic X-ray scattering. Chem. Sci. 2021, 12, 3713–3725. 10.1039/D0SC06227H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Dell’Angela M.; Anniyev T.; Beye M.; Coffee R.; Föhlisch A.; Gladh J.; Katayama T.; Kaya S.; Krupin O.; LaRue J.; Møgelhøj A.; Nordlund D.; Nørskov J. K.; Öberg H.; Ogasawara H.; Öström H.; Pettersson L. G. M.; Schlotter W. F.; Sellberg J. A.; Sorgenfrei F.; Turner J. J.; Wolf M.; Wurth W.; Nilsson A. Real-Time Observation of Surface Bond Breaking with an X-ray Laser. Science 2013, 339, 1302–1305. 10.1126/science.1231711. [DOI] [PubMed] [Google Scholar]
  31. Öström H.; Öberg H.; Xin H.; LaRue J.; Beye M.; Dell’Angela M.; Gladh J.; Ng M. L.; Sellberg J. A.; Kaya S.; Mercurio G.; Nordlund D.; Hantschmann M.; Hieke F.; Kühn D.; Schlotter W. F.; Dakovski G. L.; Turner J. J.; Minitti M. P.; Mitra A.; Moeller S. P.; Föhlisch A.; Wolf M.; Wurth W.; Persson M.; Nørskov J. K.; Abild-Pedersen F.; Ogasawara H.; Pettersson L. G. M.; Nilsson A. Probing the transition state region in catalytic CO oxidation on Ru. Science 2015, 347, 978–982. 10.1126/science.1261747. [DOI] [PubMed] [Google Scholar]
  32. Jeyachandran Y. L.; Meyer F.; Nagarajan S.; Benkert A.; Bär M.; Blum M.; Yang W.; Reinert F.; Heske C.; Weinhardt L.; Zharnikov M. Ion-Solvation-Induced Molecular Reorganization in Liquid Water Probed by Resonant Inelastic Soft X-ray Scattering. J. Phys. Chem. Lett. 2014, 5, 4143–4148. 10.1021/jz502186a. [DOI] [PubMed] [Google Scholar]; PMID: 26278946
  33. Savchenko V.; Brumboiu I. E.; Kimberg V.; Odelius M.; Krasnov P.; Liu J.-C.; Rubensson J.-E.; Björneholm O.; Såthe C.; Gråsjö J.; Dong M.; Pietzsch A.; Föhlisch A.; Schmitt T.; McNally D.; Lu X.; Polyutov S. P.; Norman P.; Iannuzzi M.; Gel’mukhanov F.; Ekholm V. Vibrational resonant inelastic X-ray scattering in liquid acetic acid: a ruler for molecular chain lengths. Sci. Rep. 2021, 11, 4098. 10.1038/s41598-021-83248-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Marchenko T.; Carniato S.; Journel L.; Guillemin R.; Kawerk E.; Žitnik M.; Kavčič M.; Bučar K.; Bohinc R.; Petric M.; da Cruz V. V.; Gel’mukhanov F.; Simon M. Electron dynamics in the core-excited CS2 molecule revealed through resonant inelastic X-ray scattering spectroscopy. J. Phys.: Conf. Ser. 2015, 635, 112012. 10.1088/1742-6596/635/11/112012. [DOI] [Google Scholar]
  35. Žitnik M.; Kavčič M.; Bohinc R.; Bučar K.; Mihelič A.; Cao W.; Guillemin R.; Journel L.; Marchenko T.; Carniato S.; Kawerk E.; Piancastelli M.; Simon M. Resonant inelastic X-ray spectroscopy of atoms and simple molecules: Satellite features and dependence on energy detuning and photon polarization. J. Electron Spectrosc. Relat. Phenom. 2015, 204, 356–364. 10.1016/j.elspec.2015.05.018. [DOI] [Google Scholar]; Gas phase spectroscopic and dynamical studies at Free-Electron Lasers and other short wavelength sources
  36. Eckert S.; Vaz da Cruz V.; Ochmann M.; von Ahnen I.; Föhlisch A.; Huse N. Breaking the Symmetry of Pyrimidine: Solvent Effects and Core-Excited State Dynamics. J. Phys. Chem. Lett. 2021, 12, 8637–8643. 10.1021/acs.jpclett.1c01865. [DOI] [PMC free article] [PubMed] [Google Scholar]; PMID: 34472857
  37. Gel’mukhanov F.; Ågren H. Resonant inelastic X-ray scattering with symmetry-selective excitation. Phys. Rev. A 1994, 49, 4378–4389. 10.1103/PhysRevA.49.4378. [DOI] [PubMed] [Google Scholar]
  38. Skytt P.; Guo J.; Wassdahl N.; Nordgren J.; Luo Y.; Ågren H. Probing symmetry breaking upon core excitation with resonant X-ray fluorescence. Phys. Rev. A 1995, 52, 3572–3576. 10.1103/PhysRevA.52.3572. [DOI] [PubMed] [Google Scholar]
  39. Luo Y.; Ågren H.; Gel’mukhanov F.; Guo J.; Skytt P.; Wassdahl N.; Nordgren J. Symmetry-selective resonant inelastic X-ray scattering of C60. Phys. Rev. B 1995, 52, 14479–14496. 10.1103/PhysRevB.52.14479. [DOI] [PubMed] [Google Scholar]
  40. Glans P.; Gunnelin K.; Skytt P.; Guo J.-H.; Wassdahl N.; Nordgren J.; Ågren H.; Gel’mukhanov F. K.; Warwick T.; Rotenberg E. Resonant X-Ray Emission Spectroscopy of Molecular Oxygen. Phys. Rev. Lett. 1996, 76, 2448–2451. 10.1103/PhysRevLett.76.2448. [DOI] [PubMed] [Google Scholar]
  41. Hergenhahn U.; Kugeler O.; Rüdel A.; Rennie E. E.; Bradshaw A. M. Symmetry-Selective Observation of the N 1s Shape Resonance in N2. J. Phys. Chem. A 2001, 105, 5704–5708. 10.1021/jp0038456. [DOI] [Google Scholar]
  42. Kosugi N. Exchange interaction in core excitation of diatomic systems. Chem. Phys. 2003, 289, 117–134. 10.1016/S0301-0104(02)00791-7. [DOI] [Google Scholar]; Decay Processes in Core-excited Species
  43. Bagus P. S.; Schaefer H. F. Localized and Delocalized 1s Hole States of the O2+ Molecular Ion. J. Chem. Phys. 2003, 56, 224–226. 10.1063/1.1676850. [DOI] [Google Scholar]
  44. Ehara M.; Nakatsuji H.; Matsumoto M.; Hatamoto T.; Liu X.-J.; Lischke T.; Prümper G.; Tanaka T.; Makochekanwa C.; Hoshino M.; Tanaka H.; Harries J. R.; Tamenori Y.; Ueda K. Symmetry-dependent vibrational excitation in N 1s photoionization of N2: Experiment and theory. J. Chem. Phys. 2006, 124, 124311. 10.1063/1.2181144. [DOI] [PubMed] [Google Scholar]
  45. Cederbaum L. S.; Domcke W. Localized and delocalized core holes and their interrelation. J. Chem. Phys. 1977, 66, 5084–5086. 10.1063/1.433763. [DOI] [Google Scholar]
  46. Xu Y.-C.; Wang S.-X.; Du X.-J.; Wang L.-H.; Zhu L.-F. Probing the delocalized core-hole via inner-shell excitation in N2. New J. Phys. 2022, 24, 053036. 10.1088/1367-2630/ac6bb7. [DOI] [Google Scholar]
  47. Köppel H.; Domcke W.; Cederbaum L. S.. Advances in Chemical Physics; John Wiley & Sons, Ltd, 1984, pp 59–246. [Google Scholar]
  48. Domcke W.; Cederbaum L. Vibronic coupling and symmetry breaking in core electron ionization. Chem. Phys. 1977, 25, 189–196. 10.1016/0301-0104(77)87075-4. [DOI] [Google Scholar]
  49. Cederbaum L. S. Symmetry breaking and localization in resonant photon emission. J. Chem. Phys. 1995, 103, 562–567. 10.1063/1.470090. [DOI] [Google Scholar]
  50. Ma Y.; Wassdahl N.; Skytt P.; Guo J.; Nordgren J.; Johnson P. D.; Rubensson J.-E.; Boske T.; Eberhardt W.; Kevan S. D. Soft-X-ray resonant inelastic scattering at the C K edge of diamond. Phys. Rev. Lett. 1992, 69, 2598–2601. 10.1103/PhysRevLett.69.2598. [DOI] [PubMed] [Google Scholar]
  51. Skytt P.; Glans P.; Guo J.-H.; Gunnelin K.; Såthe C.; Nordgren J.; Gel’mukhanov F. K.; Cesar A.; Ågren H. Quenching of Symmetry Breaking in Resonant Inelastic X-Ray Scattering by Detuned Excitation. Phys. Rev. Lett. 1996, 77, 5035–5038. 10.1103/PhysRevLett.77.5035. [DOI] [PubMed] [Google Scholar]
  52. Cesar A.; Gel’mukhanov F.; Luo Y.; Ågren H.; Skytt P.; Glans P.; Guo J.; Gunnelin K.; Nordgren J. Resonant x-ray scattering beyond the Born–Oppenheimer approximation: Symmetry breaking in the oxygen resonant x-ray emission spectrum of carbon dioxide. J. Chem. Phys. 1997, 106, 3439–3456. 10.1063/1.474111. [DOI] [Google Scholar]
  53. Gelmukhanov F.; Ågren H. X-ray resonant scattering involving dissociative states. Phys. Rev. A 1996, 54, 379–393. 10.1103/PhysRevA.54.379. [DOI] [PubMed] [Google Scholar]
  54. Gel’mukhanov F.; Sałek P.; Privalov T.; Ågren H. Duration of x-ray Raman scattering. Phys. Rev. A 1999, 59, 380–389. 10.1103/PhysRevA.59.380. [DOI] [Google Scholar]
  55. Björneholm O.; Sundin S.; Svensson S.; Marinho R. R. T.; Naves de Brito A.; Gel’mukhanov F.; Ågren H. Femtosecond Dissociation of Core-Excited HCl Monitored by Frequency Detuning. Phys. Rev. Lett. 1997, 79, 3150–3153. 10.1103/PhysRevLett.79.3150. [DOI] [Google Scholar]
  56. Sundin S.; Kh. Gel’mukhanov F.; Ågren H.; Osborne S. J.; Kikas A.; Björneholm O.; Ausmees A.; Svensson S. Collapse of Vibrational Structure in the Auger Resonant Raman Spectrum of CO by Frequency Detuning. Phys. Rev. Lett. 1997, 79, 1451–1454. 10.1103/physrevlett.79.1451. [DOI] [Google Scholar]
  57. Gel’mukhanov F.; Privalov T.; Ågren H. Collapse of vibrational structure in spectra of resonant X-ray Raman scattering. Phys. Rev. A 1997, 56, 256–264. 10.1103/PhysRevA.56.256. [DOI] [Google Scholar]
  58. Harada Y.; Tokushima T.; Takata Y.; Takeuchi T.; Kitajima Y.; Tanaka S.; Kayanuma Y.; Shin S. Dynamical Symmetry Breaking under Core Excitation in Graphite: Polarization Correlation in Soft X-Ray Recombination Emission. Phys. Rev. Lett. 2004, 93, 017401. 10.1103/PhysRevLett.93.017401. [DOI] [Google Scholar]
  59. Hennies F.; Polyutov S.; Minkov I.; Pietzsch A.; Nagasono M.; Gel’mukhanov F.; Triguero L.; Piancastelli M.-N.; Wurth W.; Ågren H.; Föhlisch A. Nonadiabatic Effects in Resonant Inelastic X-Ray Scattering. Phys. Rev. Lett. 2005, 95, 163002. 10.1103/PhysRevLett.95.163002. [DOI] [PubMed] [Google Scholar]
  60. Maganas D.; Kristiansen P.; Duda L.-C.; Knop-Gericke A.; DeBeer S.; Schlögl R.; Neese F. Combined Experimental and Ab Initio Multireference Configuration Interaction Study of the Resonant Inelastic X-ray Scattering Spectrum of CO2. J. Phys. Chem. C 2014, 118, 20163–20175. 10.1021/jp505628y. [DOI] [Google Scholar]
  61. Kramers H. A.; Heisenberg W. Über die Streuung von Strahlung durch Atome. Z. Phys. 1925, 31, 681–708. 10.1007/BF02980624. [DOI] [Google Scholar]
  62. Dirac P. A. M. The quantum theory of dispersion. Proc. R. Soc. London, Ser. A 1927, 114, 710–728. [Google Scholar]
  63. Tannor D. J.; Heller E. J. Polyatomic Raman scattering for general harmonic potentials. J. Chem. Phys. 1982, 77, 202–218. 10.1063/1.443643. [DOI] [Google Scholar]
  64. Herzberg G.; Teller E. Schwingungsstruktur der Elektronenübergänge bei mehratomigen Molekülen. Z. Phys. Chem. 1933, 21B, 410–446. 10.1515/zpch-1933-2136. [DOI] [Google Scholar]
  65. Albrecht A. C. On the Theory of Raman Intensities. J. Chem. Phys. 1961, 34, 1476–1484. 10.1063/1.1701032. [DOI] [Google Scholar]
  66. Neugebauer J.; Hess B. A. Resonance Raman spectra of uracil based on Kramers–Kronig relations using time-dependent density functional calculations and multireference perturbation theory. J. Chem. Phys. 2004, 120, 11564–11577. 10.1063/1.1697371. [DOI] [PubMed] [Google Scholar]
  67. Jensen L.; Autschbach J.; Schatz G. C. Finite lifetime effects on the polarizability within time-dependent density-functional theory. J. Chem. Phys. 2005, 122, 224115. 10.1063/1.1929740. [DOI] [PubMed] [Google Scholar]
  68. Jensen L.; Zhao L. L.; Autschbach J.; Schatz G. C. Theory and method for calculating resonance Raman scattering from resonance polarizability derivatives. J. Chem. Phys. 2005, 123, 174110. 10.1063/1.2046670. [DOI] [PubMed] [Google Scholar]
  69. Rappoport D.; Shim S.; Aspuru-Guzik A. Simplified Sum-Over-States Approach for Predicting Resonance Raman Spectra. Application to Nucleic Acid Bases. J. Phys. Chem. Lett. 2011, 2, 1254–1260. 10.1021/jz200413g. [DOI] [PubMed] [Google Scholar]; PMID: 26295418
  70. Lee S.; Heller E. J. Time-dependent theory of Raman scattering. J. Chem. Phys. 1979, 71, 4777–4788. 10.1063/1.438316. [DOI] [Google Scholar]
  71. Thomas M.; Latorre F.; Marquetand P. Resonance Raman spectra of ortho-nitrophenol calculated by real-time time-dependent density functional theory. J. Chem. Phys. 2013, 138, 044101. 10.1063/1.4776218. [DOI] [PubMed] [Google Scholar]
  72. Mattiat J.; Luber S. Efficient calculation of (resonance) Raman spectra and excitation profiles with real-time propagation. J. Chem. Phys. 2018, 149, 174108. 10.1063/1.5051250. [DOI] [PubMed] [Google Scholar]
  73. Ovander L. N.; Shadura V. A. Green function method and its application to the resonance Raman effect. J. Raman Spectrosc. 2001, 32, 587–590. 10.1002/jrs.735. [DOI] [Google Scholar]
  74. Kas J. J.; Rehr J. J.; Soininen J. A.; Glatzel P. Real-space Green’s function approach to resonant inelastic x-ray scattering. Phys. Rev. B 2011, 83, 235114. 10.1103/physrevb.83.235114. [DOI] [Google Scholar]
  75. Guthmuller J.; Champagne B. Resonance Raman Scattering of Rhodamine 6G as Calculated by Time-Dependent Density Functional Theory: Vibronic and Solvent Effects. J. Phys. Chem. A 2008, 112, 3215–3223. 10.1021/jp7112279. [DOI] [PubMed] [Google Scholar]; PMID: 18327928
  76. Ma H.; Liu J.; Liang W. Time-Dependent Approach to Resonance Raman Spectra Including Duschinsky Rotation and Herzberg–Teller Effects: Formalism and Its Realistic Applications. J. Chem. Theory Comput. 2012, 8, 4474–4482. 10.1021/ct300640c. [DOI] [PubMed] [Google Scholar]; PMID: 26605607
  77. Meyer H.-D.; Manthe U.; Cederbaum L. The multi-configurational time-dependent Hartree approach. Chem. Phys. Lett. 1990, 165, 73–78. 10.1016/0009-2614(90)87014-I. [DOI] [Google Scholar]
  78. Freibert A.; Mendive-Tapia D.; Huse N.; Vendrell O. Femtosecond x-ray absorption spectroscopy of pyrazine at the nitrogen K-edge: on the validity of the Lorentzian limit. J. Phys. B: At., Mol. Opt. Phys. 2021, 54, 244003. 10.1088/1361-6455/ac3846. [DOI] [Google Scholar]
  79. Cederbaum L. S.; Domcke W.; Schirmer J. Many-body theory of core holes. Phys. Rev. A 1980, 22, 206–222. 10.1103/PhysRevA.22.206. [DOI] [Google Scholar]
  80. Worth G. A.; Cederbaum L. S. Beyond Born-Oppenheimer: Molecular Dynamics Through a Conical Intersection. Annu. Rev. Phys. Chem. 2004, 55, 127–158. 10.1146/annurev.physchem.55.091602.094335. [DOI] [PubMed] [Google Scholar]; PMID: 15117250
  81. Cederbaum L.; Domcke W.; Köppel H.; Von Niessen W. Strong vibronic coupling effects in ionization spectra: The “mystery band” of butatriene. Chem. Phys. 1977, 26, 169–177. 10.1016/0301-0104(77)87041-9. [DOI] [Google Scholar]
  82. Baer M. Adiabatic and diabatic representations for atom-molecule collisions: Treatment of the collinear arrangement. Chem. Phys. Lett. 1975, 35, 112–118. 10.1016/0009-2614(75)85599-0. [DOI] [Google Scholar]
  83. Pacher T.; Cederbaum L. S.; Köppel H.. Advances in Chemical Physics; John Wiley & Sons, Ltd, 1993, pp 293–391. [Google Scholar]
  84. Purvis G. D.; Bartlett R. J. A full coupled-cluster singles and doubles model: The inclusion of disconnected triples. J. Chem. Phys. 1982, 76, 1910–1918. 10.1063/1.443164. [DOI] [Google Scholar]
  85. Dunning T. H. Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. 10.1063/1.456153. [DOI] [Google Scholar]
  86. Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams-Young D.; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian 16. Revision A.03; Gaussian Inc: Wallingford CT, 2016.
  87. Epifanovsky E.; Gilbert A. T. B.; Feng X.; Lee J.; Mao Y.; Mardirossian N.; Pokhilko P.; White A. F.; Coons M. P.; Dempwolff A. L.; Gan Z.; Hait D.; Horn P. R.; Jacobson L. D.; Kaliman I.; Kussmann J.; Lange A. W.; Lao K. U.; Levine D. S.; Liu J.; McKenzie S. C.; Morrison A. F.; Nanda K. D.; Plasser F.; Rehn D. R.; Vidal M. L.; You Z.-Q.; Zhu Y.; Alam B.; Albrecht B. J.; Aldossary A.; Alguire E.; Andersen J. H.; Athavale V.; Barton D.; Begam K.; Behn A.; Bellonzi N.; Bernard Y. A.; Berquist E. J.; Burton H. G. A.; Carreras A.; Carter-Fenk K.; Chakraborty R.; Chien A. D.; Closser K. D.; Cofer-Shabica V.; Dasgupta S.; de Wergifosse M.; Deng J.; Diedenhofen M.; Do H.; Ehlert S.; Fang P.-T.; Fatehi S.; Feng Q.; Friedhoff T.; Gayvert J.; Ge Q.; Gidofalvi G.; Goldey M.; Gomes J.; González-Espinoza C. E.; Gulania S.; Gunina A. O.; Hanson-Heine M. W. D.; Harbach P. H. P.; Hauser A.; Herbst M. F.; Hernández Vera M.; Hodecker M.; Holden Z. C.; Houck S.; Huang X.; Hui K.; Huynh B. C.; Ivanov M.; Jász Á.; Ji H.; Jiang H.; Kaduk B.; Kähler S.; Khistyaev K.; Kim J.; Kis G.; Klunzinger P.; Koczor-Benda Z.; Koh J. H.; Kosenkov D.; Koulias L.; Kowalczyk T.; Krauter C. M.; Kue K.; Kunitsa A.; Kus T.; Ladjánszki I.; Landau A.; Lawler K. V.; Lefrancois D.; Lehtola S.; Li R. R.; Li Y.-P.; Liang J.; Liebenthal M.; Lin H.-H.; Lin Y.-S.; Liu F.; Liu K.-Y.; Loipersberger M.; Luenser A.; Manjanath A.; Manohar P.; Mansoor E.; Manzer S. F.; Mao S.-P.; Marenich A. V.; Markovich T.; Mason S.; Maurer S. A.; McLaughlin P. F.; Menger M. F. S. J.; Mewes J.-M.; Mewes S. A.; Morgante P.; Mullinax J. W.; Oosterbaan K. J.; Paran G.; Paul A. C.; Paul S. K.; Pavošević F.; Pei Z.; Prager S.; Proynov E. I.; Rák Á.; Ramos-Cordoba E.; Rana B.; Rask A. E.; Rettig A.; Richard R. M.; Rob F.; Rossomme E.; Scheele T.; Scheurer M.; Schneider M.; Sergueev N.; Sharada S. M.; Skomorowski W.; Small D. W.; Stein C. J.; Su Y.-C.; Sundstrom E. J.; Tao Z.; Thirman J.; Tornai G. J.; Tsuchimochi T.; Tubman N. M.; Veccham S. P.; Vydrov O.; Wenzel J.; Witte J.; Yamada A.; Yao K.; Yeganeh S.; Yost S. R.; Zech A.; Zhang I. Y.; Zhang X.; Zhang Y.; Zuev D.; Aspuru-Guzik A.; Bell A. T.; Besley N. A.; Bravaya K. B.; Brooks B. R.; Casanova D.; Chai J.-D.; Coriani S.; Cramer C. J.; Cserey G.; DePrince A. E.; DiStasio R. A.; Dreuw A.; Dunietz B. D.; Furlani T. R.; Goddard W. A.; Hammes-Schiffer S.; Head-Gordon T.; Hehre W. J.; Hsu C. P.; Jagau T. C.; Jung Y.; Klamt A.; Kong J.; Lambrecht D. S.; Liang W.; Mayhall N. J.; McCurdy C. W.; Neaton J. B.; Ochsenfeld C.; Parkhill J. A.; Peverati R.; Rassolov V. A.; Shao Y.; Slipchenko L. V.; Stauch T.; Steele R. P.; Subotnik J. E.; Thom A. J. W.; Tkatchenko A.; Truhlar D. G.; Van Voorhis T.; Wesolowski T. A.; Whaley K. B.; Woodcock H. L.; Zimmerman P. M.; Faraji S.; Gill P. M. W.; Head-Gordon M.; Herbert J. M.; Krylov A. I. Software for the frontiers of quantum chemistry: An overview of developments in the Q-Chem 5 package. J. Chem. Phys. 2021, 155, 084801. 10.1063/5.0055522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Stanton J. F.; Bartlett R. J. The equation of motion coupled-cluster method. A systematic biorthogonal approach to molecular excitation energies, transition probabilities, and excited state properties. J. Chem. Phys. 1993, 98, 7029–7039. 10.1063/1.464746. [DOI] [Google Scholar]
  89. Vidal M.; Feng X.; Epifanovsky E.; Krylov A. I.; Coriani S. New and Efficient Equation-of-Motion Coupled-Cluster Framework for Core-Excited and Core-Ionized States. J. Chem. Theory Comput. 2019, 15, 3117–3133. 10.1021/acs.jctc.9b00039. [DOI] [PubMed] [Google Scholar]
  90. Worth G. A.; Beck M. H.; Jäckle A.; Meyer H.; Otto F.; Brill M.; Vendrell O.. The MCTDH Package. version 8.6, 2020. http://mctdh.uni-hd.de. (accessed date 21.01.2024).
  91. Williams S. O.; Imre D. G. Raman spectroscopy: time-dependent pictures. J. Phys. Chem. 1988, 92, 3363–3374. 10.1021/j100323a012. [DOI] [Google Scholar]
  92. Tannor D. J.Introduction to Quantum Mechanics: A Time-Dependent Perspective; University Science Books; 1st ed., 2007. [Google Scholar]
  93. Vendrell O.; Meyer H.-D. Multilayer multiconfiguration time-dependent Hartree method: Implementation and applications to a Henon–Heiles Hamiltonian and to pyrazine. J. Chem. Phys. 2011, 134, 044135. 10.1063/1.3535541. [DOI] [PubMed] [Google Scholar]
  94. Wang H.; Thoss M. Multilayer formulation of the multiconfiguration time-dependent Hartree theory. J. Chem. Phys. 2003, 119, 1289–1299. 10.1063/1.1580111. [DOI] [Google Scholar]
  95. Manthe U. A multilayer multiconfigurational time-dependent Hartree approach for quantum dynamics on general potential energy surfaces. J. Chem. Phys. 2008, 128, 164116. 10.1063/1.2902982. [DOI] [PubMed] [Google Scholar]
  96. Meyer H.-D. Studying molecular quantum dynamics with the multiconfiguration time-dependent Hartree method. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2012, 2, 351–374. 10.1002/wcms.87. [DOI] [Google Scholar]
  97. Dirac P. A. M. Note on Exchange Phenomena in the Thomas Atom. Math. Proc. Cambridge Philos. Soc. 1930, 26, 376–385. 10.1017/S0305004100016108. [DOI] [Google Scholar]
  98. Frenkel J.. In Wave mechanics; advanced general theory; The international series of monographs on physics; The Clarendon press: Oxford, 1934. [Google Scholar]
  99. Innes K.; Ross I.; Moomaw W. R. Electronic states of azabenzenes and azanaphthalenes: A revised and extended critical review. J. Mol. Spectrosc. 1988, 132, 492–544. 10.1016/0022-2852(88)90343-8. [DOI] [Google Scholar]
  100. Wilson E. B. The Normal Modes and Frequencies of Vibration of the Regular Plane Hexagon Model of the Benzene Molecule. Phys. Rev. 1934, 45, 706–714. 10.1103/PhysRev.45.706. [DOI] [Google Scholar]
  101. Prince K. C.; Vondráček M.; Karvonen J.; Coreno M.; Camilloni R.; Avaldi L.; de Simone M. A critical comparison of selected 1s and 2p core hole widths. J. Electron Spectrosc. Relat. Phenom. 1999, 101–103, 141–147. 10.1016/S0368-2048(98)00436-8. [DOI] [Google Scholar]
  102. Sala M.; Lasorne B.; Gatti F.; Guérin S. The role of the low-lying dark nπ* states in the photophysics of pyrazine: a quantum dynamics study. Phys. Chem. Chem. Phys. 2014, 16, 15957–15967. 10.1039/C4CP02165G. [DOI] [PubMed] [Google Scholar]
  103. Tsuru S.; Vidal M. L.; Pápai M.; Krylov A. I.; Møller K. B.; Coriani S. Time-resolved near-edge X-ray absorption fine structure of pyrazine from electronic structure and nuclear wave packet dynamics simulations. J. Chem. Phys. 2019, 151, 124114. 10.1063/1.5115154. [DOI] [PubMed] [Google Scholar]
  104. Heller E. J. Quantum corrections to classical photodissociation models. J. Chem. Phys. 1978, 68, 2066–2075. 10.1063/1.436029. [DOI] [Google Scholar]
  105. Schinke R.Photodissociation Dynamics: Spectroscopy and Fragmentation of Small Polyatomic Molecules. In Cambridge Monographs on Atomic, Molecular and Chemical Physics; Cambridge University Press, 1993. [Google Scholar]
  106. Heller E. J.; Sundberg R.; Tannor D. Simple aspects of Raman scattering. J. Phys. Chem. 1982, 86, 1822–1833. 10.1021/j100207a018. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ct3c01259_si_001.pdf (142.5KB, pdf)

Articles from Journal of Chemical Theory and Computation are provided here courtesy of American Chemical Society

RESOURCES