Perspective: Ultrafast imaging of molecular dynamics using ultrafast low-frequency lasers, x-ray free electron laser and electron pulses
←
→
Page content transcription
If your browser does not render page correctly, please read the page content below
Perspective: Ultrafast imaging of molecular dynamics using ultrafast low-frequency lasers, x-ray free electron laser and electron pulses Ming Zhang,1 Zhengning Guo,1 Xiaoyu Mi,1 Zheng Li,1, 2, 3, ∗ and Yunquan Liu1, 2, 4, † 1 arXiv:2201.03197v1 [physics.atom-ph] 10 Jan 2022 State Key Laboratory for Mesoscopic Physics and Collaborative Innovation Center of Quantum Matter, School of Physics, Peking University, Beijing 100871, China 2 Collaborative Innovation Center of Extreme Optics, Shanxi University, Taiyuan, Shanxi 030006, China 3 Peking University Yangtze Delta Institute of Optoelectronics, Nantong, China 4 Center for Applied Physics and Technology, HEDPS, Peking University, Beijing 100871, China (Dated: January 11, 2022) Abstract The requirement of high space-time resolution and brightness is a great challenge for imaging atomic motion and making molecular movies. Important breakthroughs in ultrabright tabletop laser, x-ray and electron sources have enabled the direct imaging of evolving molecular structures in chemical processes. And recent experimental advances in preparing ultrafast laser and electron pulses equipped molecular imag- ing with femtosecond time resolution. This Perspectives present an overview of versatile imaging methods of molecular dynamics. High-order harmonic generation imaging maps the harmonic order of photons to the time delay for nuclear motion and can reach attosecond time resolution. Coulomb explosion imaging re- trieves molecular structural information by detecting the momentum vectors of fragmented ions. Diffraction imaging encodes molecular structural and electronic information in reciprocal space, achieving high spa- tial resolution by using x-ray or electrons with short wavelengths. The enhancement of x-ray and electron source brightness pushes diffraction into single molecular limit without the restriction of crystalline materi- als. Laser-induced electron diffraction is a self-imaging method based on coherent electron scattering. We also present various applications of these ultrafast imaging methods in resolving laser-induced nuclear and electronic dynamics, as well as dynamic symmetry evolution of molecules. ∗ zheng.li@pku.edu.cn † yunquan.liu@pku.edu.cn 1
FIG. 1. The state of the art methods and experimental tools for ultrafast imaging of molecular dynamics using tabletop laser, x-ray free electron laser (XFEL) light sources and ultrashort electron pulses. I. INTRODUCTION The imaging of complex molecular structures with electron or x-ray diffraction is challenging and it is especially difficult to extract dynamic structural information from their diffraction pat- terns with femtosecond time resolution, which is the characteristic time scale of atomic motion in molecules. Direct observation of atomic motions in real time during chemical reactions is thought as a Gedanken experiment for chemists [1, 2]. Given the extremely high spatial and temporal reso- lution required to image atomic motions along reaction coordinates, it was thought to be the purest form of a Gedanken experiment. It was even argued by Manfred Eigen in his Nobel lecture [3] that we could never achieve the necessary imaging technologies to see atoms rearrange in space during chemical reactions, as he argued that we would always have to use indirect methods for inferring the underlying mechanism of chemical reactions. Despite the challenge for high spatio-temporal resolution requirement, it has been a pedagogical tool for theoretical conceptualization of chem- ical reactions by transition state involving molecular structural dynamics at the barrier-crossing regions, and recent experimental advances in ultrabright tabletop laser, x-ray and electron sources 2
have brought the Gedanken experiment of directly imaging molecular motion into reality [4]. Considering the spatial resolution for experimentally imaging such structural dynamics, the length scale for chemical bonds is Ångstrom and the accurate value can be determined from static gas phase electron diffraction with 0.001 Åprecision [4, 5]. For example, in chemical reactions, intramolecular and intermolecular electron transfer play an important role in stabilizing the con- figuration under a proper energy [6, 7]. The magnitude order of molecular structural variation in electron transfer processes is 0.01 Å to 0.1 Å [8, 9], which is feasible to be imaged by high energy electrons. Apart from the spatial resolution requirements, achieving the femtosecond time resolu- tion is also an essential factor in catching the very moment in chemical reaction. In unimolecular reactions, time scale corresponding to the thermal sampling frequency of a barrier between reactant and product is 100 fs at room temperature. These frequencies in thermal sampling involve motions damped by anharmonic coupling in reaction coordinates. The time scale for intramolecular vibra- tional energy redistribution is also 100 fs, which is viewed as the shutter speed for imaging the molecular dynamics in reaction modes [10, 11]. For chemical process involving strongly repulsive excited state potentials like photodissociation, the time scale for well-defined reaction coordinates is 10 fs [12]. But for most reactions, the time scale for canonical time resolution is 100 fs to ps, for example, photoisomerization of retinal [13], electron transfer [14, 15], proton transfer [16] and C-H bond activation [17]. Thus we regard 10-100 fs as the time scale of chemistry. The challenge to image chemical process can be generalized as achieving the needed brightness of electron and light sources with sub-Å spatial resolution and subpicosecond temporal resolution. For more than a hundred years, sub-Å spatial resolution has been reached by using x-rays with sufficiently short wavelengths to determine atomic positions. Over 20 years ago, plasma gener- ation or similar electron pulse generation techniques via photoemission gave birth to the 100 fs pulses and the subpicosecond time resolution was reached. But these techniques are unable to make femtosecond movies directly in molecules because of their low brightness, and stroboscopic pump-probe method was applied to enhance the total brightness [18]. But the structural change and excitation dynamics make the sampling process irreversible, which limits the application scope for these techniques. With the advance of ultrafast high intensity laser technology, strong-field laser induced tunnel- ing, laser-induced electron diffraction, high harmonic generation imaging and Coulomb explosion imaging have been introduced as new tools into the field of ultrafast imaging of structural dy- namics. The imaging methods using high-order harmonic generation (HHG) and photoelectron 3
diffraction are used for reconstructing nuclear dynamics and tomographic imaging of molecu- lar Dyson orbitals. Using the relationship between HHG signal and the nuclear autocorrelation function [19], the nuclear motion of H+ + 2 and D2 can be reconstructed, and attosecond temporal resolution is reached through mapping between the harmonic order of photons to time delay of electron trajectory [20, 21]. Beyond nuclear degrees of freedom, the highest occupied molecular orbital (HOMO) of molecules, which is a good approximation of the Dyson orbital in outer va- lence photoionization, can be selectively and tomographically reconstructed by recording HHG spectra for different alignment angles under single active electron approximation [22, 23]. Apart from HHG imaging, the tunneling ionized photoelectron rescattering off molecular ion also pro- vides a self-imaging method, laser-induced electron diffraction (LIED) [24]. For example, the dissociation dynamics of di-ionized acetylene (C2 H2+ 2 ) can be visualized by extracting the tem- poral evolution of bond distances from LIED experimental data [25], and Dyson orbital of ion- ized molecules is incorporated in the photoelectron angular distribution (PAD) [26]. Coulomb explosion imaging (CEI) has become prevailing to determine the molecular structure by mea- suring the momentum vectors of fragmented ions following multiple ionization processes in the pump-probe experiments [27, 28], and one can map the measured fragment momentum data to the instantaneous geometry of the molecule at the time of the ionization. The CEI method is widely accepted because it allows convenient interpretation of the results, e.g. Coulomb explosion of methanol reveals its complex mechanisms for different production breakup channels including transient proton migration and long-range ”inverse harpooning” [29], the conformational isomers of 1,2-dibromoethane (C2 H4 Br2 ) can be distinguished by exploiting the three-body breakup chan- nel C2 H+ + + 4 + Br + Br containing both sequential and concerted breakup processes [30]. How- ever, the accurate reconstruction of molecular structure could hardly go beyond the measurement of bond length of diatomic molecules, because the complex atomic motion during the Coulomb explosion leads to the deviation of measured bond angle. And this in most cases hinders direct re- construction of molecular structure, unless a full time reversed propagation of molecular fragment trajectories is made possible. The required accuracy of momenta measurement is not within reach of current experimental techniques [31]. The development of ultrabright electron sources and the advent of x-ray free electron lasers (XFEL) broaden the application scope for imaging of molecules [32–34]. Compared to Coulomb explosion imaging, ultrafast coherent diffraction imaging provides an opportunity to directly im- age the molecular configuration and structural dynamics [4, 35]. High intensity and short pulse 4
length of XFEL can push diffraction to the single-molecule limit and reconstruct the image of non- crystalline objects [36] by detecting continuous diffraction intensity rather than Bragg spots. For example, the structure of lithium peroxide (Li2 O2 ) can be elucidated using a combination of x-ray and first-principles techniques [37]. Besides, x-ray diffraction of α-cyclodextrin shows rigid ring structures and anisotropic characteristics, which reveals the distance of neighboring glycosidic oxygen atoms and the orientation of individual molecules [38]. Ultrafast electron diffraction (UED) exhibits high spatial resolution, but suffered from space charge broadening which limits the temporal resolution [25], the issue is resolved by using rf pulse compression and relativistic effect of electrons of MeV energy, and the UED can currently have the time resolution of ∼50fs [35, 39]. Compared to x-ray scattering, electron scattering is more sensitive to the properties of electronic states such as electronic correlation [40], and goes beyond the independent atom model commonly used for traditional diffraction experiments, which neglects the redistribution of valence electrons during chemical-bond formation [41]. For example, the S1 (nπ ∗ ) state population evolution of pyridine is reflected by inelastic signal related to the two- electron correlation effect due to Coulomb repulsion between electrons [39]. Starting from the probability densities of molcular nuclei determined by ultrafast diffraction, the Wigner function and the density matrix of the quantum wave packet can be obtained by quantum tomography [35, 42]. In this Perspective, we emphasize the ultrafast imaging methods illustrated in Fig. 1, which make use of table-top laser, XFEL and electron sources. II. MOLECULAR IMAGING WITH TABLETOP INTENSE LASERS A. Ultrafast Imaging by Laser-induced Ionization In the prototypical strong-field interaction, the valence electron can be released through tunnel ionization from the molecule. The photoelectron momentum distributions (PMDs) encode struc- tural information of its origin. The measured far-field PMDs can be understood as a diffraction image of the source. Thus, in principle, it should be possible to retrieve structural information by analyzing the interference pattern. However, the source is not identical to the atomic or molecular orbital from which the electron is removed but rather the corresponding Dyson orbital. For the molecular species, the molecular frame photoelectron angular distributions can be measured. By 5
FIG. 2. Schematic figure of strong laser field tunneling ionization, recollision-based imaging, Coulomb ex- plosion imaging and diffraction imaging. (a) High harmonic generation imaging and photoelectron imaging are based on the recollision between the tunnel-ionized electron and parent ion. (b) Symmetry variation of methane cation is reflected by the changing of Coulomb explosion dissociation channels detected by different pump-probe time delay. (c) Gas-phase electron/x-ray diffraction imaging of molecules encodes structural and electronic information. 6
measuring the PMDs of the aligned molecules N2 , O2 [43–45], the molecular structures have been imaged. Recently, the strong-field direct ionization of the complex molecules has been demon- strated [46]. The photoelectron spectroscopy of CS2 reflects its nonadiabatic photodissociation processes involving vibrational quantum beat and different dissociation channels through singlet and triplet intermediate metastable states [47]. In the future, when the complex molecules are spa- tially confined, it would be possible to extract the structure for the complex molecules. Apart from detecting nuclear dynamics, intriguing approaches are recently developed to investigate ultrafast electron dynamics in molecules, such as the attosecond angular streaking (attoclock) with circular femtosecond laser fields, in which the photoelectrons can be streaked into different direction de- pending on the ionization instant in the molecular frame. The geometry of two-color co-rotating circularly polarized fields has been used as the novel attoclock scheme for atoms [48, 49]. Trans- ferring this technique from atoms to polyatomic molecules could provide a potential approach to study the ultrafast laser-molecular interaction and electron dynamics of molecules. B. Imaging Methods using Laser-induced Rescattering As shown in Fig. 2(a), besides direct electron tunneling ionization, the strong laser field can also accelerate the free electron in vacuum and eventually drives it back to the parent ion, predom- inantly resulting in rescattering or radiative recombination. The radiative recombination results in the emission of high-energy photons through high-harmonic generation. The three steps of HHG in semiclassical model [50] can be correlated with a pump-probe process: the tunneling ionization plays the role as the pump that launches nuclear dynamics, and recollision of electron and parent ion is the probe step for nuclear motion occurring during the time delay [21]. When the fraction of ionized electron wavepacket interferes with the orbital from which the electron was extracted, the electron density distribution oscillates as the wave packet propagates and leads to radiative emis- sion of harmonics [22]. By aligning the molecules, the spectrum of high harmonic generation can be measured, it is a powerful tool to investigate the electronic structure with attosecond temporal resolution. Alternatively, the rescattered portion of the electron wavepacket is exploited in laser- induced electron diffraction experiments as a coherent diffraction pattern of the molecular target. For example, LIED is used to retrieve multiple bond lengths from an aligned polyatomic molecule acetylene [51]. C. I. Blaga et al. use gas-phase LIED to map the structural responses of oxygen and nitrogen molecules to ionization, reaching the spatio-temporal resolution of 0.1-Å change in the 7
oxygen bond length occurring in the time scale of ∼5 fs [52]. Several challenges arise for experi- mental techniques including achieving high recollision energies and sufficient momentum transfer for the ionized electron [51]. In contrast with the conventional electron diffraction, the influence of strong laser field needs to be removed for extraction of field-free electron–ion cross sections based on the quantitative rescattering theory [53]. In the future, the application of laser-induced rescat- tering effect can potentially provide time-dependent images of the molecule at sub-femtosecond and few-picometer resolution for the complicated molecules using the mid-infrared or infrared lasers and the higher pondermotive potential. C. Coulomb Explosion Imaging The Coulomb explosion imaging method provides instantaneous 3D imaging of nuclear po- sitions [54], such as the structure of hydrogen dimers [55], laser-induced isomerization in ace- tonitrile (CH3 CN) [56] and photodissociation dynamics of chlorocarbonylsulfenyl chloride [27] and CH2 BrI [57]. In CEI experiments, the bonding electrons of molecules are stripped away by ultrashort lasers in a time scale shorter than nuclear motion, leaving positive ions in the positions of molecular nuclei. High intensity lasers is feasible to stripping many electrons, and typically, at least several binding electrons are scattered away. For visible laser, sufficiently high intensity (∼ 1014 W/cm2 ) is needed to successively remove the electrons from molecule and generate the in- teratomic Coulomb repulsion [58, 59]. In contrast, core ionization by x-ray will lead to cascading Auger decay and create highly charged molecules, which fall apart under Coulomb repulsion [60]. As the molecular potential suddenly changes to Coulomb repulsion dominated potential of the ion’s degree of freedom, the molecule dissociates and the CEI detector collects velocities by si- multaneous determination of molecular fragments from single molecule, and the 3D images of nuclear positions can be resolved by analyzing the momentum distribution of the fragments. The fragment trajectories for predicting detected velocity from initial conformations can be approx- imated using classical simulations with sufficient accuracy when at least one of the fragments charge is two or more. The density function of conformations including correlated variations within the molecular structure can be obtained by collecting structures of a molecular ensemble. The scheme of obtaining molecular structure from CEI is relatively direct and model independent, and classical trajectories approximation is usually sufficient especially for highly charged frag- ments. But the sudden approximation of electron stripping may not always be not fully satisfied, 8
leading to deviation from this simplified interpretation [54]. As shown in Fig. 2(b), by extracting the atomic motion with Coulomb explosion momentum mapping, the variation of molecular symmetry variation processes can be revealed. A prominent example is the spontaneous symmetry breaking, which is a fundamental phenomenon occurred in numerous fields of physics, such as the Higgs mechanism [61], and Jahn-Teller effect in molecular systems. The Jahn-Teller (JT) effect, arising from nonadiabatic coupling between electrons and nuclei, leads to distortion of nuclear configuration away from high symmetric points accompanied by removal of electronic state degeneracy [62]. For a typical linear vibronic coupling Hamiltonian of two state, the diabatic potential forming JT type conical intersection can be written as [63] V1 (x, y) λy V = , (1) λy V2 (x, y) where the tuning mode x and coupling mode y are two nuclear coordinates, and V1 and V2 are two diabatic state potential energy surface. Experimentally, the JT deformation of methane cation could be indirectly imaged by Coulomb explosion imaging in ∼10 fs time scale using tabletop laser, which is able to create ions up to triple charges by strong field ionization [64]. The strong pump pulse ionizes neutral methane and populates the CH+ 2 4 ground state F2 . After JT distortion from Td to C3v geometry, f2 and e vibrational modes lead to further distortion to C2v symmetry, where the electronic degeneracy is completely lifted [65]. A ∼20 fs time delay is observed between the two-body and three-body Coulomb explosion channels, which corresponds to the symmetry evolution from C3v to C2v , and is driven by the superposition of multiple vibrational modes [64]. However, the limitation of low cross sections in creating highly charged molecular ions through strong field valence ionization by tabletop lasers with wavelength in the IR to visible regime hin- ders the accuracy of structural analysis in CEI. Ideally, we expect to have an n-ion for an n-atomic molecule, in this manner, we can retrieve maximal information about the atomic positions in the molecule at the instant of Coulomb explosion. One of the solutions is to use ultrafast x-ray as probe to induce Coulomb explosion, this has been realized by free electron lasers, which can easily create highly charged molecular states via ionization cascades [31]. As an example, the proton migration process for acetylene dication can be revealed by x-ray CEI, which is related to the nature of chemical bonding forming and breakup. X-ray probe creates C2 D4+ 2+ 4 ion from C2 D2 dication for Coulomb explosion, and the CEI exhibits increase of average C-C-D angle with temporal evolution, which resolves the bending motion due to softening of the σ bonding of C-C-D atoms within 100 fs in 1 Σ+ g state [60]. However, even in 9
FIG. 3. Potential energy surface (PES) for CH+ 4 cation and Newton map of three-body Coulomb explosion momentum mapping at 9 fs and 29 fs pump-probe time delay. The CH+ 4 cation in a C3v geometry (at the peak of PES) undergoes JT distortion to a C2v geometry (at the bottom of PES). The spot-like structures correspond to different molecular symmetries of CH+ 4 . The momentum mapping of reference configuration C3v , C2v and D2d is based on the simulated result of Coulomb repulsion, which takes into account the atomic motion during CEI process. case with Coulomb explosion of 4+ ions for the four-atomic molecule, the disadvantage of CEI as an indirect imaging method prevents it to resolve the molecular dynamics by experiment alone and one has to incorporate input from theoretical simulations. In the case of CEI of C2 D2+ 2 dy- namics, the ab initio calculation predicts much longer timescale due to an isomerization barrier of ∼2eV [31]. Actually, the signal can be reproduced by the relative rotation of Coulomb exploded fragments, but large magnitude proton motion without leading to isomerization does exist on sub- 100 fs timescale because of ∼ 2.4 eV relaxation energy released from the weakening of C-C bond in high-lying 1πu−1 3σg−1 states [31]. The enhancement of momentum measurement accuracy of provides the possibility to resolve the rotational motion of Coulomb explosion fragment and to di- 10
rectly reconstruct molecular structure by time reversed propagation of the Newtonian equation of motion. However, the required resolution can be exceptionally high. For the dynamics of C2 D2+ 2 dication, the needed resolution can be estimated by considering a bending mode of vibrational period T ∼ 100 fs and of bond length r ∼ 1 Å. Due angular momentum conservation during the Coulomb explosion process, the tangential momentum p⊥ perpendicular to the bond explo- sion direction at the detector position R ∼ 0.1 m is given by 2πmr 2 /T = Rp⊥ , which gives a required accuracy p⊥ ∼ 10−8 a.u., far above the current experimental resolution of momentum measurements. III. ULTRAFAST X-RAY DIFFRACTION IMAGING Ultrafast coherent diffraction imaging provides a direct access to resolve the dynamical molec- ular structure, and thus offers an alternative route that is complementary to the Coulomb explosion imaging. As shown in Fig. 2(c), the electronic information such as evolution of electron den- sity [66], and the atomic positions in the molecule are encoded in the scattered wave of coherent incident x-ray [67], and can be determined by the measurement of diffraction intensity. The lost phase information of the electron form factor can be recovered by phase retrieval algorithms [68]. Various photochemical processes [17, 69, 70] and valence electron dynamics [71–73] have been investigated by the time-resolved diffraction. For crystalline samples, the diffraction signal at discrete Bragg peaks comes from interference enhancement of scattered waves from different lattice sites R ~ α . The total charge density in Fourier ~ space σ̂ = α σ̂α = α σ̂e−i~s·Rα is the sum of individual charge density, where ~s is the momen- P P tum transfer between incident and diffracted x-ray photon. The total diffraction intensity (apart from a constant coefficient) for N lattice sites is [74] (2) X X I = Tr(N ) (ρ̂(N ) σ̂ † σ̂) = Tr(1) (ρ̂(1) † α σ̂α σ̂α ) + Tr(2) (ρ̂αβ σ̂α† σ̂β ), (2) α=β α6=β where the atoms or molecules from different lattice sites are uncorrelated, so the second order (2) (1) (1) density matrix of two sites is direct product of the density matrix for each site ρ̂αβ = ρ̂α ⊗ ρ̂β . Suppose the excitation ratio from ground state |ψg i to excited state |ψe i is η, ρ̂(1) α = η |ψe i hψe | + (1 − η) |ψg i hψg | . (3) 11
The first term of diffraction intensity is the unimolecular contribution, X I1 = Tr(1) (ρ̂(1) † α σα σα ) α=β Xh i = ηTr(1) (|ψe i hψe | σ̂ † σ̂) + (1 − η)Tr(1) (|ψg i hψg | σ̂ † σ̂) α = N η ψe σ̂ † σ̂ ψe + (1 − η) ψg σ̂ † σ̂ ψg , (4) for the electron form factor fg/e = ψg/e |σ̂| ψg/e of the ground and excited state. Ignoring transition density terms hψg |σ̂| ψe i, we have the elastic unimolecular signal I1 = N (1 − η)|fg |2 + η|fe |2 . (5) The biomolecular contribution is (1) X I2 = Tr(2) (ρ̂(1) † α σ̂α ⊗ ρ̂β σ̂β ) α6=β ~ ~ X = ei~s·(Rα −Rβ ) Tr(1) (ρ̂(1) σ̂ † )Tr(1) (ρ̂(1) σ̂) α6=β ~ ~ X = ei~s·(Rα −Rβ ) |(1 − η)fg + ηfe |2 . (6) α6=β The diffraction signal I2 scales as N 2 , which is overwhelmingly predominant in the crystal diffrac- tion, because I1 is proportional to N. The interference of ground state and excited state form fac- tors appeared in I2 amplifies the change of diffraction signal, especially for weak excitation η ≪ 1. For example, the photoexcitation of polycrystalline ammonium sulfate [(NH4 )2 SO4 ] induces elec- tron transfer followed by a geometry stabilization process. By measuring the change of x-ray diffraction intensity of Bragg peaks ∆Ihkl , the electron flow pathways and proton translocation processes can be determined, which are reflected by the modification of structural factors [75]. For diffraction from gas phase molecules, the average of bimolecular diffraction I2 vanishes due to randomly distributed molecules and the only detected signal is I1 . X-ray free electron laser can satisfy the high intensity requirement and push diffraction to the single-molecule limit. Ultrafast laser pulses shorter than the time scale of the radiation damage enable the room temper- ature ”diffraction before destruction” technique for single molecule diffraction imaging [76] and provide high temporal resolution to image dynamical processes of molecules with femtosecond resolution. For strongly aligned 2,5-diiodo-benzonitrile molecule, the coherent x-ray diffraction pattern arising from the interference of two iodine atoms at both ends separating by 700 pm is ob- served using 2 keV (620 pm) XFEL [77]. Although the first interference maximum is not recorded 12
due to relative long wavelength, the distance of two iodine atoms can be derived from measured diffraction intensity by χ2 analysis with simulation results. A more precise result showing the first three maxima of iodine-iodine interference is obtained by increasing the photon energy to 9.5 keV [78]. However, the contemporary spatio-temporal resolution of ultrafast diffraction is, in certain cases, still insufficient to detect molecular dynamics of small amplitude motion, such as the Jahn- Teller structural distortion of photoexcited CF3 I molecules [79]. We simulated the diffraction pattern of CF3 I before and after the JT distortion on the CASSCF(6,5)/DZVP level using Ter- aChem program [80]. Because of the JT effect (see Fig. 4) [81], the bending of C-I bond lifts the degeneracy of 3 E states from C3v to Cs geometry, and the distortion angle for C-I bond bending θ = 4.5◦ corresponding to the minimum of A′ state potential energy curve. As the C3v symmetry breaks, the iodine atom fragments on a conical surface around the molecular axis, and generates an off-axis recoil to the CF3 group. The diffraction pattern and the percentage difference (PD) of signal between Franck-Condon (FC) geometry and JT geometry defined as [39] IJT (s) − IFC (s) PD(s) = × 100 . (7) IFC (s) are shown in Fig. 4. The percentage difference signal of diffraction pattern reflects the change of electronic density distribution and the molecular structure variation. However, the amplitude of signal is rather small (around 1%) and it is apparently difficult for ultrafast diffraction exper- iments to have sufficient resolution for revealing the structural symmetry breaking of such small amplitude. This pushes us to envisage new theory for ultrafast diffraction beyond classical ball-and-stick model of molecules, which can assist us to tackle open problems such as resolving small amplitude JT distortion. A savvy way to reveal Jahn-Teller effect in photoexcited CF3 I is to use the imprints of symmetry breaking left in the quantum state of molecules. During the C-I bond breakup pro- cess, the off-axis recoil momentum of the I atom acting on the CF3 fragment can obviously lead to its rotational excitation [82]. Reconstructing the quantum mechanical density matrix of the ex- cited rotational states of CF3 is equivalent to resolving the JT effect. Because the ultrafast coherent diffraction encodes the information of time-dependent probability density of the molecular quan- tum system, it naturally provides the possibility for tomographic reconstruction of the quantum states [83, 84]. Restricted by the dimension problem [85, 86], traditional quantum tomography (QT) method [87, 88] can only be applied to systems of a single degree of freedom, which cat- 13
FIG. 4. Rotationally averaged CF3 I diffraction pattern of 3 E states. Brown and black spheres are I and C atom, respectively, and the blue torus represents three F atoms. 1D and 2D x-ray diffraction pattern of Franck-Condon geometry is shown in the upper panel (arbitrary unit), and percentage difference (PD) diffraction pattern of JT distortion geometry (θ = 4.5◦ ) is shown in the lower panel. egorically excludes the rotation problem. However, the underlying nature of QT is to retrieve the phase information lost in measurement [35]. Analogous to crystallographic phase retrieval, the dimension problem can be solved by iterative projection algorithm using additional physical constraints of density matrix [42]. The generalized QT algorithm resolves the dimension problem and can be applied to make quantum version of ”molecular movie” for rotational and vibrational motions. IV. ULTRAFAST ELECTRON DIFFRACTION IMAGING Compared with x-ray diffraction, electron diffraction exhibits higher spatial resolution due to its shorter de Broglie wavelengths and larger scattering cross sections, which is especially important for the gas-phase diffraction experiments [25]. Gas phase ultrafast electron diffraction can provide 14
sub-Å spatial resolution and sub-picosecond temporal resolution using compact setups [89]. The spatial charge broadening due to Coulomb repulsion of electrons and group velocity mismatch be- tween electron and laser pulses decreases the temporal resolution [90], which can be remedied by relativistic electron acceleration or electron bunch compression [91, 92]. Because the longitudinal space-charge pulse elongation is proportional to 1/β 2 γ 5 , where β = v/c, γ = (1 − β 2 )−1/2 and v, c are the speed of electrons and vacuum light speed, respectively, the spatial charge effect can be greatly diminished for relativistic electrons [89, 93], electron sources with the MeV energy scale range enhances the time resolution to the 10 fs domain [94, 95]. Another advantage of electron diffraction is its sensitivity in detecting electronic correlation, which is incorporated in the inelastic part. This can be understood by the difference in the scatter- ing operators of x-ray diffraction σ̂X and electron diffraction σ̂e [67] X σ̂X = ei~s·~ri , (8) i ! 1 X ~ν i~s·R X σ̂e = 2 Nν e − ei~s·~ri , (9) s ν i where ~s is the momentum transfer vector and Nν is the charge carried by ν-th nuclear. The elastic and inelastic electron diffraction intensity for molecules in an electronic state |ψ0 i is [39] 2 IR X ~ X Ie(ela) (~s) = 4 Nν ei~s·Rν − ψ0 ei~s·~ri ψ0 , (10) s ν i 2 IR X X Ie(inela) (~s) = 4 ψ0 ei~s·~rij ψ0 − ψ0 ei~s·~ri ψ0 , (11) s i,j i and for the x-ray diffraction intensity [67] 2 (ela) X i~s·~ ri IX (~s) = IT ψ0 e ψ0 , (12) i 2 (inela) X X IX (~s) = IT ψ0 ei~s·~rij ψ0 − ψ0 ei~s·~ri ψ0 , (13) i,j i where IR and IT are Rutherford and Thomson scattering factors, respectively. The inelastic signal reflects electron correlation including exchange and Coulomb repulsion effects, which is related to two-electron reduced density ρ(2) (~r1 , ~r2 ) incorporating the probabil- ity of two-electron distance. Thus for the small angle scattering, which is essential for electron correlation [39], the additional denominator s4 significantly enlarges the signal of smaller s and 15
FIG. 5. Comparison of percentage difference (PD) signals of x-ray and electron diffraction signal of pyridine in the S0 and S1 (nπ ∗ ) states. Black, red and white spheres are C, N and H atoms, respectively. Isosurface representations of n and π ∗ molecular orbitals are shown of value 0.1Å−3/2 . The lone pair n orbital is localized at N and C atoms, where the electron correlation is much stronger than the anti-bonding π ∗ orbital. As shown in the right column, the PD diffraction signal of x-ray diffraction (above) is smaller than electron diffraction (below) due to the difference of elastic signal background. improves the accuracy of experimental measurements. Besides, the elastic background for elec- tron diffraction is much smaller, for the two terms of σ̂e contributing from electronic and nuclear parts cancel out at forward scattering direction, which eliminates the unchanged part of the signal and enhances the sensitivity of percentage difference signal. We simulated the percentage dif- ference signal of x-ray and electron diffraction for the pyridine molecule S0 and S1 (nπ ⋆ ) state on the CASSCF(10,9)/cc-pVTZ level. As shown in Fig. 5, the electron diffraction shows larger PD signals and could be more sensitive to the change of electronic correlation in electronic state transition. UED is also a direct imaging method for dynamic structural evolution, for example, the tempo- 16
rally evolving structure of photoexcited CF3 I molecule can be reconstructed by ultrafast gas-phase electron diffraction of laser aligned sample of isolated molecules [96]. The real space interatomic distances characterized by the pair distribution functions (PDF) can be extracted by Fourier trans- formation of diffraction pattern. Furthermore, the photodissociation and conical intersection dy- namics of CF3 I has been directly imaged [79]. Single-photon excitation to the dissociative 3 Q0 state leads to the loss of C-I and F-I atom pairs reflected by PDF. The umbrella and breathing vibra- tional motion are resolved by comparison of C-F bond length and F-C-F bond angle retrieved from PDF and simulation result. For the two-photon channel, real space trajectory encoded in PDF in- dicates the nonadiabatic transition from Rydberg 7s state to CF+ − 3 -I ion pair state and wavepacket bifurcation through conical intersection seam [79]. V. CONCLUDING REMARKS We have presented the current state of the art in imaging molecular dynamics with tabletop low-frequency ultrafast laser, x-ray free electron lasers and electron pulses. The imaging by laser induced photoionization and electron rescattering, and by ultrafast x-ray and electron diffraction are versatile and powerful tools for mapping dynamic molecular structure with numerous advan- tages. In comparison with spectroscopic methods, where the structural information is indirectly inferred based on exact knowledge of energy landscape, ultrafast imaging techniques are more effective to provide direct access to atom distances and intuitive view of chemical reactions [93]. However, to image complex molecule remains challenging for the indistinguishable overlap of the pair distribution functions from different atom pairs in the diffraction imaging. Instead of com- paring and optimizing theoretical model to best fit the data, it would be more powerful to direct retrieval of structural information without theoretical models [89]. Further development including brighter and shorter light or electron sources, higher spatio-temporal resolution and more robust structure retrieval methods will be needed for making molecular movies, especially for complex molecules. 17
VI. ACKNOWLEDGEMENT We thank the support of National Natural Science Foundation of China (Grant Nos. 92050201, 12174009, 11774013). [1] John C. Polanyi and Ahmed H. Zewail, “Direct observation of the transition state,” Acc. Chem. Res. 28, 119–132 (1995). [2] R.J. Dwayne Miller, “Mapping atomic motions with ultrabright electrons: The chemists’ gedanken experiment enters the lab frame,” Annu. Rev. Phys. Chem. 65, 583–604 (2014). [3] Manfred Eigen, “Immeasurably fast reactions,” Nobel Lecture (1967). [4] Anatoly A. Ischenko, Peter M. Weber, and R. J. Dwayne Miller, “Capturing chem- istry in action with electrons: Realization of atomically resolved reaction dynamics,” Chem. Rev. 117, 11066–11124 (2017). [5] A. A. Ischenko, I. V. Kochikov, and R. J. Dwayne Miller, “The effect of Coulomb repulsion on the space-time resolution limits for ultrafast electron diffraction,” J. Chem. Phys. 150, 054201 (2019). [6] Rudolph A. Marcus, “Electron transfer reactions in chemistry theory and experiment,” J. Electr. Chem. 438, 251–259 (1997). [7] E M Kosower and D Huppert, “Excited state electron and proton transfers,” Annu. Rev. Phys. Chem. 37, 127–156 (1986). [8] Tomáš Kubař and Marcus Elstner, “What governs the charge transfer in DNA? the role of DNA con- formation and environment,” J. Phys. Chem. B 112, 8788–8798 (2008). [9] Harald Oberhofer and Jochen Blumberger, “Charge constrained density functional molecular dynam- ics for simulation of condensed phase electron transfer reactions,” J. Chem. Phys. 131, 064101 (2009). [10] Jason R Dwyer, Christoph T Hebeisen, Ralph Ernstorfer, Maher Harb, Vatche B Deyirmenjian, Robert E Jordan, and R.J Dwayne Miller, “Femtosecond electron diffraction:’making the molecu- lar movie’,” Philos. Trans. R. Soc. A 364, 741–778 (2006). [11] G. Sciaini and R. D. Miller, “Femtosecond electron diffraction - heralding the era of atomically re- solved dynamics,” Rep. Prog. Phys. 74, 096101 (2011). [12] Brian Stankus, James M. Budarz, Adam Kirrander, David Rogers, Joseph Robinson, Thomas J. Lane, Daniel Ratner, Jerome Hastings, Michael P. Minitti, and Peter M. Weber, “Femtosecond photodissoci- 18
ation dynamics of 1,4-diiodobenzene by gas-phase x-ray scattering and photoelectron spectroscopy,” Faraday Discuss. 194, 525–536 (2016). [13] Philip J. M. Johnson, Alexei Halpin, Takefumi Morizumi, Leonid S. Brown, Valentyn I. Prokhorenko, Oliver P. Ernst, and R. J. Dwayne Miller, “The photocycle and ultrafast vibrational dynamics of bacteriorhodopsin in lipid nanodiscs,” Phys. Chem. Chem. Phys. 16, 21310–21320 (2014). [14] R.A. Marcus and Norman Sutin, “Electron transfers in chemistry and biology,” Biochim. Biophys. Acta 811, 265–322 (1985). [15] Mark Maroncelli, Jean MacInnis, and Graham R. Fleming, “Polar solvent dynamics and electron- transfer reactions,” Science 243, 1674–1681 (1989). [16] Paul F. Barbara, Gilbert C. Walker, and Terrance P. Smith, “Vibrational modes and the dynamic solvent effect in electron and proton transfer,” Science 256, 975–981 (1992). [17] Hubert Jean-Ruel, Meng Gao, Michal A. Kochman, Cheng Lu, Lai C. Liu, Ryan R. Cooney, Car- ole A. Morrison, and R. J. D. Miller, “Ring-closing reaction in diarylethene captured by femtosecond electron crystallography,” J. Phys. Chem. B 117, 15894–15902 (2013). [18] Vladimir A. Lobastov, Ramesh Srinivasan, and Ahmed H. Zewail, “Four-dimensional ultrafast elec- tron microscopy,” Proc. Natl. Acad. Sci. USA 102, 7069–7073 (2005). [19] Manfred Lein, “Attosecond probing of vibrational dynamics with high-harmonic generation,” Phys. Rev. Lett. 94, 053004 (2005). [20] Pengfei Lan, Marc Ruhmann, Lixin He, Chunyang Zhai, Feng Wang, Xiaosong Zhu, Qingbin Zhang, Yueming Zhou, Min Li, Manfred Lein, and Peixiang Lu, “Attosecond probing of nuclear dynamics with trajectory-resolved high-harmonic spectroscopy,” Phys. Rev. Lett. 119, 033201 (2017). [21] Sarah Baker, Joseph S Robinson, CA Haworth, H Teng, RA Smith, CC Chirilă, Manfred Lein, JWG Tisch, and Jonathan P Marangos, “Probing proton dynamics in molecules on an attosecond time scale,” Science 312, 424 (2006). [22] D. M. Villeneuve, J. Itatani, J. Levesque, D. Zeidler, Hiromichi Niikura, H. Pépin, J. C. Kieffer, and P. B. Corkum, “Tomographic imaging of molecular orbitals,” Nature 432, 867–871 (2004). [23] Serguel Patchkovskii, Zengxiu Zhao, Thomas Brabec, and D. M. Villeneuve, “High harmonic genera- tion and molecular orbital tomography in multielectron systems,” J. Chem. Phys. 126, 114306–114306 (2007). [24] Michael G Pullen, Benjamin Wolter, A-T Le, Matthias Baudisch, Michele Sclafani, Hugo Pires, Claus Dieter Schroeter, Joachim Ullrich, Robert Moshammer, Thomas Pfeifer, C. D. Lin, and 19
J. Biegert, “Influence of orbital symmetry on diffraction imaging with rescattering electron wave pack- ets,” Nat. Commun. 7, 11922 (2016). [25] Benjamin Wolter, Michael G Pullen, A-T Le, Matthias Baudisch, K Doblhoff-Dier, Arne Senftleben, Michael Hemmer, C Dieter Schröter, Joachim Ullrich, Thomas Pfeifer, et al., “Ultrafast electron diffraction imaging of bond breaking in di-ionized acetylene,” Science 354, 308 (2016). [26] C. M. Oana and Anna I. Krylov, “Dyson orbitals for ionization from the ground and electronically excited states within equation-of-motion coupled-cluster formalism: Theory, implementation, and ex- amples,” J. Chem. Phys. 127, 234106–234106 (2007). [27] Graham A. Cooper, S. T. Alavi, Wen Li, Suk K. Lee, and Arthur G. Suits, “Coulomb explosion dynamics of chlorocarbonylsulfenyl chloride,” J. Phys. Chem. A 125, 5481–5489 (2021). [28] J. Gagnon, Kevin F. Lee, D. M. Rayner, P. B. Corkum, and V. R. Bhardwaj, “Coincidence imaging of polyatomic molecules via laser-induced Coulomb explosion,” J. Phys. Chem. B 41, 215104 (2008). [29] Itamar Luzon, Ester Livshits, Krishnendu Gope, Roi Baer, and Daniel Strasser, “Making sense of Coulomb explosion imaging,” J. Phys. Chem. Lett. 10, 1361–1367 (2019). [30] Shashank Pathak, Razib Obaid, Surjendu Bhattacharyya, Johannes Burger, Xiang Li, Jan Tross, Travis Severt, Brandin Davis, Rene C. Bilodeau, Carlos A. Trallero-Herrero, Artem Rudenko, Nora Berrah, and Daniel Rolles, “Differentiating and quantifying gas-phase conformational isomers using Coulomb explosion imaging,” J. Phys. Chem. Lett. 11, 10205–10211 (2020). [31] Zheng Li, Ludger Inhester, Chelsea Liekhus-Schmaltz, Basile F. E. Curchod, James W. Snyder, Nikita Medvedev, James Cryan, Timur Osipov, Stefan Pabst, Oriol Vendrell, Phil Bucksbaum, and Todd J. Martinez, “Ultrafast isomerization in acetylene dication after carbon K-shell ionization,” Nat. Com- mun. 8, 453–7 (2017). [32] M. Bargheer, N. Zhavoronkov, Y. Gritsai, J. C. Woo, D. S. Kim, M. Woerner, and T. El- saesser, “Coherent atomic motions in a nanostructure studied by femtosecond x-ray diffraction,” Science 306, 1771–1773 (2004). [33] Shouhua Nie, Xuan Wang, Hyuk Park, Richard Clinite, and Jianming Cao, “Measurement of the elec- tronic grüneisen constant using femtosecond electron diffraction,” Phys. Rev. Lett. 96, 025901 (2006). [34] A. Cavalleri, C. W. Siders, F. L. H. Brown, D. M. Leitner, C. Tóth, J. A. Squier, C. P. J. Barty, K. R. Wilson, K. Sokolowski-Tinten, M. Horn von Hoegen, D. von der Linde, and M. Kammler, “Anharmonic lattice dynamics in germanium measured with ultrafast x-ray diffraction,” Phys. Rev. Lett. 85, 586–589 (2000). 20
[35] Zheng Li, Sandeep Gyawali, Anatoly A. Ischenko, Stuart Hayes, and R. J. D. Miller, “Mapping atomic motions with electrons: Toward the quantum limit to imaging chemistry,” ACS Photonics 7, 296–320 (2020). [36] Henry N. Chapman, “X-ray imaging beyond the limits,” Nat. Mater. 8, 299–301 (2009). [37] Maria K. Y. Chan, Eric L. Shirley, Naba K. Karan, Mahalingam Balasubramanian, Yang Ren, Jeffrey P. Greeley, and Tim T. Fister, “Structure of lithium peroxide,” J. Phys. Chem. Lett. 2, 2483–2486 (2011). [38] Kazuaki Kato, Kohzo Ito, and Taiki Hoshino, “Anisotropic amorphous x-ray diffraction attributed to the orientation of cyclodextrin,” J. Phys. Chem. Lett. 11, 6201–6205 (2020). [39] Jie Yang, Xiaolei Zhu, J. P. F. Nunes, Jimmy K. Yu, Robert M. Parrish, Thomas J. A. Wolf, Martin Centurion, Markus Gühr, Renkai Li, Yusong Liu, Bryan Moore, Mario Niebuhr, Suji Park, Xiaozhe Shen, Stephen Weathersby, Thomas Weinacht, Todd J. Martinez, and Xijie Wang, “Simultaneous observation of nuclear and electronic dynamics by ultrafast electron diffraction,” Science 368, 885– 889 (2020). [40] Markus Kowalewski, Kochise Bennett, and Shaul Mukamel, “Monitoring nonadiabatic avoided cross- ing dynamics in molecules by ultrafast x-ray diffraction,” Struct. Dyn. 4, 054101–054101 (2017). [41] Robin Santra, “Concepts in x-ray physics,” J. Phys. B: At. Mol. Opt. Phys. 42, 023001 (2009). [42] Ming Zhang, Shuqiao Zhang, Yanwei Xiong, Hankai Zhang, Anatoly A. Ischenko, Oriol Vendrell, Xiaolong Dong, Xiangxu Mu, Martin Centurion, Haitan Xu, R. J. D. Miller, and Zheng Li, “Quantum state tomography of molecules by ultrafast diffraction,” Nat. Commun. 12, 5441–5441 (2021). [43] M. Meckel, D. Comtois, D. Zeidler, A. Staudte, D. Pavicic, H. C. Bandulet, H. Pepin, J. C. Kieffer, R. Doerner, D. M. Villeneuve, and P. B. Corkum, “Laser-induced electron tunneling and diffraction,” Science 320, 1478–1482 (2008). [44] M. Meckel, A. Staudte, S. Patchkovskii, D. M. Villeneuve, P. B. Corkum, R. Dörner, and M. Spanner, “Signatures of the continuum electron phase in molecular strong-field photoelectron holography,” Nat. Phys. 10, 594–600 (2014). [45] Ming-Ming Liu, Min Li, Chengyin Wu, Qihuang Gong, André Staudte, and Yunquan Liu, “Phase structure of strong-field tunneling wave packets from molecules,” Phys. Rev. Lett. 116, 163004– 163004 (2016). [46] Joss Wiese, Jolijn Onvlee, Sebastian Trippel, and Jochen Küpper, “Strong-field ionization of complex molecules,” Phys. Rev. Research 3, 013089 (2021). 21
[47] Shutaro Karashima, Yoshi-Ichi Suzuki, and Toshinori Suzuki, “Ultrafast extreme ultraviolet photo- electron spectroscopy of nonadiabatic photodissociation of CS2 from 1 B2 (1 Σ+ u ) state: Product for- mation via an intermediate electronic state,” J. Phys. Chem. Lett. 12, 3755–3761 (2021). [48] Meng Han, Peipei Ge, Yun Shao, Qihuang Gong, and Yunquan Liu, “Attoclock photoelectron inter- ferometry with two-color corotating circular fields to probe the phase and the amplitude of emitting wave packets,” Phys. Rev. Lett. 120, 073202–073202 (2018). [49] Peipei Ge, Meng Han, Yongkai Deng, Qihuang Gong, and Yunquan Liu, “Universal description of the attoclock with two-color corotating circular fields,” Phys. Rev. Lett. 122, 013201–013201 (2019). [50] P. B. Corkum, “Plasma perspective on strong field multiphoton ionization,” Phys. Rev. Lett. 71, 1994– 1997 (1993). [51] Michael G Pullen, Benjamin Wolter, Anh-Thu Le, Matthias Baudisch, Michaël Hemmer, Arne Sen- ftleben, Claus Dieter Schröter, Joachim Ullrich, Robert Moshammer, Chii-Dong Lin, and J. Biegert, “Imaging an aligned polyatomic molecule with laser-induced electron diffraction,” Nat. Commun. 6, 7262 (2015). [52] Cosmin I Blaga, Junliang Xu, Anthony D DiChiara, Emily Sistrunk, Kaikai Zhang, Pierre Agostini, Terry A Miller, Louis F DiMauro, and CD Lin, “Imaging ultrafast molecular dynamics with laser- induced electron diffraction,” Nature 483, 194 (2012). [53] C. D. Lin, Anh-Thu Le, Zhangjin Chen, Toru Morishita, and Robert Lucchese, “Strong-field rescat- tering physics—self-imaging of a molecule by its own electrons,” J. Phys. B: At. Mol. Opt. Phys. 43, 122001 (2010). [54] Zeev Vager, “Coulomb explosion imaging of molecules,” Adv. At. Mol. Opt. Phy. 45, 203 (2001). [55] Arnab Khan, Till Jahnke, Stefan Zeller, Florian Trinter, Markus Schöffler, Lothar P. H. Schmidt, Reinhard Dörner, and Maksim Kunitski, “Visualizing the geometry of hydrogen dimers,” J. Phys. Chem. Lett. 11, 2457–2463 (2020). [56] Matteo McDonnell, Aaron C. LaForge, Juan Reino-Gonzalez, Martin Disla, Nora G. Kling, De- badarshini Mishra, Razib Obaid, Margaret Sundberg, Vit Svoboda, Sergio Diaz-Tendero, Fernando Martin, and Nora Berrah, “Ultrafast laser-induced isomerization dynamics in acetonitrile,” J. Phys. Chem. Lett. 11, 6724–6729 (2020). [57] Michael Burt, Rebecca Boll, Jason W. L. Lee, Kasra Amini, Hansjochen Köckert, Claire Val- lance, Alexander S. Gentleman, Stuart R. Mackenzie, Sadia Bari, Cédric Bomme, Stefan Düsterer, Benjamin Erk, Bastian Manschwetus, Erland Müller, Dimitrios Rompotis, Evgeny Savelyev, Nora 22
Schirmel, Simone Techert, Rolf Treusch, Jochen Küpper, Sebastian Trippel, Joss Wiese, Hen- rik Stapelfeldt, Barbara Cunha de Miranda, Renaud Guillemin, Iyas Ismail, Loı̈c Journel, Tatiana Marchenko, Jérôme Palaudoux, Francis Penent, Maria Novella Piancastelli, Marc Simon, Oksana Travnikova, Felix Brausse, Gildas Goldsztejn, Arnaud Rouzée, Marie Géléoc, Romain Geneaux, Thierry Ruchon, Jonathan Underwood, David M. P. Holland, Andrey S. Mereshchenko, Pavel K. Olshin, Per Johnsson, Sylvain Maclot, Jan Lahl, Artem Rudenko, Farzaneh Ziaee, Mark Brouard, and Daniel Rolles, “Coulomb-explosion imaging of concurrent CH2 BrI photodissociation dynamics,” Phys. Rev. A 96, 043415 (2017). [58] Maria E. Corrales, Gregory Gitzinger, Jesus Gonzalez-Vazquez, Vincent Loriot, Rebeca de Nalda, and Luis Bañares, “Velocity map imaging and theoretical study of the Coulomb explosion of CH3 I under intense femtosecond IR pulses,” J. Phys. Chem. A 116, 2669–2677 (2011). [59] Yanmei Wang, Song Zhang, Zhengrong Wei, and Bing Zhang, “Velocity map imaging of dissociative ionization and Coulomb explosion of CH3 I induced by a femtosecond laser,” J. Phys. Chem. A 112, 3846–3851 (2008). [60] Chelsea E. Liekhus-Schmaltz, Ian Tenney, Timur Osipov, Alvaro Sanchez-Gonzalez, Nora Berrah, Re- becca Boll, Cedric Bomme, Christoph Bostedt, John D. Bozek, Sebastian Carron, Ryan Coffee, Julien Devin, Benjamin Erk, Ken R. Ferguson, Robert W. Field, Lutz Foucar, Leszek J. Frasinski, James M. Glownia, Markus Gühr, Andrei Kamalov, Jacek Krzywinski, Heng Li, Jonathan P. Marangos, Todd J. Martinez, Brian K. McFarland, Shungo Miyabe, Brendan Murphy, Adi Natan, Daniel Rolles, Artem Rudenko, Marco Siano, Emma R. Simpson, Limor Spector, Michele Swiggers, Daniel Walke, Song Wang, Thorsten Weber, Philip H. Bucksbaum, and Vladimir S. Petrovic, “Ultrafast isomerization initiated by x-ray core ionization,” Nat. Commun. 6, 8199–8199 (2015). [61] Peter W. Higgs, “Broken symmetries and the masses of gauge bosons,” Phys. Rev. Lett. 13, 508–509 (1964). [62] Isaac Bersuker, The Jahn-Teller Effect (Cambridge University Press, 2006). [63] Caroline Arnold, Oriol Vendrell, Ralph Welsch, and Robin Santra, “Control of nuclear dynam- ics through conical intersections and electronic coherences,” Phys. Rev. Lett. 120, 123001–123001 (2018). [64] Min Li, Ming Zhang, Oriol Vendrell, Zhenning Guo, Qianru Zhu, Xiang Gao, Lushuai Cao, Keyu Guo, Qin-Qin Su, Wei Cao, Siqiang Luo, Jiaqing Yan, Yueming Zhou, Yunquan Liu, Zheng Li, and Peixiang Lu, “Ultrafast imaging of spontaneous symmetry breaking in a photoionized molecular sys- 23
tem,” Nat. Commun. 12, 4233–4233 (2021). [65] Ugo Jacovella, Hans J. Wörner, and Frédéric Merkt, “Jahn-teller effect and large-amplitude motion in CH+ 4 studied by high-resolution photoelectron spectroscopy of CH4 ,” J. Mol. Spectrosc. 343, 62–75 (2018). [66] Daria Popova-Gorelova, David A. Reis, and Robin Santra, “Theory of x-ray scattering from laser- driven electronic systems,” Phys. Rev. B 98, 224302 (2018). [67] L. S. Bartell and R. M. Gavin, “Effects of electron correlation in x-ray and electron diffraction,” J. Am. Chem. Soc. 86, 3493–3498 (1964). [68] S. Marchesini, “A unified evaluation of iterative projection algorithms for phase retrieval,” Rev. Sci. Instrum. 78 (2007). [69] S. Techert, F. Schotte, and M. Wulff, “Picosecond x-ray diffraction probed transient structural changes in organic solids,” Phys. Rev. Lett. 86, 2030–2033 (2001). [70] Jorg Hallmann, Wolfgang Morgenroth, Carsten Paulmann, Jav Davaasambuu, Qingyu Kong, Michael Wulff, and Simone Techert, “Time-resolved x-ray diffraction of the photochromic α-styrylpyrylium trifluoromethanesulfonate crystal films reveals ultrafast structural switching,” J. Am. Chem. Soc. 131, 15018–15025 (2009). [71] Haiwang Yong, Stefano M. Cavaletto, and Shaul Mukamel, “Ultrafast valence-electron dynamics in oxazole monitored by x-ray diffraction following a stimulated x-ray raman excitation,” J. Phys. Chem. Lett. 12, 9800–9806 (2021). [72] Daniel Healion, Yu Zhang, Jason D. Biggs, Niranjan Govind, and Shaul Mukamel, “Entangled va- lence electron–hole dynamics revealed by stimulated attosecond x-ray raman scattering,” J. Phys. Chem. Lett. 3, 2326–2331 (2012). [73] Zotev N. Stankus B., Yong H., “Ultrafast x-ray scattering reveals vibrational coherence following Rydberg excitation,” Nat. Chem. 11, 716–721 (2019). [74] Kochise Bennett, Markus Kowalewski, and Shaul Mukamel, “Comment on “self-referenced coherent diffraction x-ray movie of Angstrom- and femtosecond-scale atomic motion”,” Phys. Rev. Lett. 119, 069301 (2017). [75] Michael Woerner, Flavio Zamponi, Zunaira Ansari, Jens Dreyer, Benjamin Freyer, Mirabelle Prémont- Schwarz, and Thomas Elsaesser, “Concerted electron and proton transfer in ionic crystals mapped by femtosecond x-ray powder diffraction,” J. Chem. Phys. 133, 064509 (2010). 24
[76] Anton Barty, Jochen Küpper, and Henry N. Chapman, “Molecular imaging using x-ray free-electron lasers,” Annu. Rev. Phys. Chem. 64, 415–435 (2013). [77] Jochen Küpper, Stephan Stern, Lotte Holmegaard, Frank Filsinger, Arnaud Rouzée, Artem Rudenko, Per Johnsson, Andrew V. Martin, Marcus Adolph, Andrew Aquila, Sa ša Bajt, Anton Barty, Christoph Bostedt, John Bozek, Carl Caleman, Ryan Coffee, Nicola Coppola, Tjark Delmas, Sascha Epp, Benjamin Erk, Lutz Foucar, Tais Gorkhover, Lars Gumprecht, Andreas Hartmann, Robert Hart- mann, Günter Hauser, Peter Holl, Andre Hömke, Nils Kimmel, Faton Krasniqi, Kai-Uwe Kühnel, Jochen Maurer, Marc Messerschmidt, Robert Moshammer, Christian Reich, Benedikt Rudek, Robin Santra, Ilme Schlichting, Carlo Schmidt, Sebastian Schorb, Joachim Schulz, Heike Soltau, John C. H. Spence, Dmitri Starodub, Lothar Strüder, Jan Thøgersen, Marc J. J. Vrakking, Georg Weidenspoint- ner, Thomas A. White, Cornelia Wunderer, Gerard Meijer, Joachim Ullrich, Henrik Stapelfeldt, Daniel Rolles, and Henry N. Chapman, “X-ray diffraction from isolated and strongly aligned gas-phase molecules with a free-electron laser,” Phys. Rev. Lett. 112, 083002 (2014). [78] Thomas Kierspel, Andrew Morgan, Joss Wiese, Terry Mullins, Andy Aquila, Anton Barty, Richard Bean, Rebecca Boll, Sébastien Boutet, Philip Bucksbaum, et al., “X-ray diffractive imaging of con- trolled gas-phase molecules: Toward imaging of dynamics in the molecular frame,” J. Chem. Phys. 152, 084307 (2020). [79] Jie Yang, Xiaolei Zhu, Thomas J. A. Wolf, Zheng Li, J. P. F. Nunes, Ryan Coffee, James P. Cryan, Markus Gühr, Kareem Hegazy, Tony F. Heinz, Keith Jobe, Renkai Li, Xiaozhe Shen, Theodore Vec- cione, Stephen Weathersby, Kyle J. Wilkin, Charles Yoneda, Qiang Zheng, Todd J. Martinez, Martin Centurion, and Xijie Wang, “Imaging CF3 I conical intersection and photodissociation dynamics with ultrafast electron diffraction,” Science 361, 64 (2018). [80] Ivan S. Ufimtsev and Todd J. Martinez, “Quantum chemistry on graphical processing units. 3. ana- lytical energy gradients, geometry optimization, and first principles molecular dynamics,” J. Chem. Theory Comput. 5, 2619–2628 (2009). [81] Yoshiaki Amatatsu, Keiji Morokuma, and Satoshi Yabushita, “Ab initio potential energy surfaces and trajectory studies of a-band photodissociation dynamics: CH3 I∗ → CH3 + I and CH3 + I∗ ,” J. Chem. Phys. 94, 4858–4876 (1991). [82] Daiqian Xie, Hua Guo, Yoshiaki Amatatsu, and Ronnie Kosloff, “Three-dimensional photodissocia- tion dynamics of rotational state selected methyl iodide,” J. Phys. Chem. A 104, 1009–1019 (2000). 25
You can also read