Attosecond Diffraction Imaging of Electron Dynamics in Solids

Mingrui Yuan Wyant College of Optical Sciences, University of Arizona, Tucson, AZ 85721 Department of Physics, University of Arizona, Tucson, AZ 85721    Nikolay V. Golubev ngolubev@arizona.edu Department of Physics, University of Arizona, Tucson, AZ 85721
(July 3, 2024)
Abstract

Visualizing the electron dynamics in four dimensions of space and time is crucial to the understanding of several ubiquitous processes in nature. Hence, ultrafast X-ray and electron imaging tools have been developed to probe the dynamics of matter by means of the time-resolved diffraction imaging (TRDI). In this work, we report an extension of the theory underlying the TRDI to the case of the laser-driven electron dynamics in solid state systems. We demonstrate that the TRDI signal encodes an essential information about the time-dependent electron density of the system under study and thus makes it possible to decipher the ultrafast quantum dynamics and the electron transfer phenomena in solids. We apply the developed approach to image the laser-driven electron dynamics in neutral graphene showing that the predictions made by the fully quantum version of the TRDI deviate significantly from those obtained with the conventional semiclassical approach.

pacs:
Valid PACS appear here

Recent advancements in attosecond technologies [1] have facilitated the observation and eventually control of ultrafast electron dynamics in atoms [2], clusters [3], molecules [4, 5, 6], liquids [7], and solids [8, 9] with unprecedented resolution. However, the majority of previously utilized techniques, such as high-harmonic generation [5], attosecond transient absorption [6], photoelectron [10] and photofragmentation [11, 4] spectroscopies, are specifically designed to capture the temporal evolution of a system but often provide insufficient information about the electron dynamic in the space domain.

Established experimental techniques that have been used for over a century to image the spatial structure of matter are X-ray [12, 13] and electron diffraction [14, 15, 16, 17] spectroscopies. With the advent of free-electron lasers [18, 19] it is now possible to generate short coherent pulses in X-ray energy domain permitting thus direct observations of dynamics of matter with atomic spatial resolution. Even better temporal resolution has been recently demonstrated in attosecond electron diffraction microscopy experiment measuring the electron dynamics in the neutral multilayer graphene [20]. This attomicroscopy setup together with the free-electron laser facilities provide access to long-awaited real-space imaging of electron motion in matter on its natural time scales.

Theoretical description of X-ray and electron scattering from matter has been widely studied before [21, 22]. A traditional semiclassical (SC) approach to connect the scattering signal with the properties of a system under study is based on a simple relation between the electron density Q(𝐫)𝑄𝐫Q(\mathbf{r})italic_Q ( bold_r ), where 𝐫𝐫\mathbf{r}bold_r denotes the real-space coordinate, and the intensity of the scattered beam I(𝐒)𝐼𝐒I(\mathbf{S})italic_I ( bold_S ), where 𝐒𝐒\mathbf{S}bold_S is the scattering vector. Under the first Born approximation and assuming the scattering is elastic, the scattering intensity is proportional to the absolute value of the Fourier transform of the electron density [21, 22]: I(𝐒)|^𝐒[Q(𝐫)]|2proportional-to𝐼𝐒superscriptsubscript^𝐒delimited-[]𝑄𝐫2I(\mathbf{S})\propto|\hat{\mathcal{F}}_{\mathbf{S}}[Q(\mathbf{r})]|^{2}italic_I ( bold_S ) ∝ | over^ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q ( bold_r ) ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where ^^\hat{\mathcal{F}}over^ start_ARG caligraphic_F end_ARG denotes the Fourier transform operator. While in the past the computation of the electron density of a system Q(𝐫)𝑄𝐫Q(\mathbf{r})italic_Q ( bold_r ) was a complex problem by itself, the ever growing computational power made the simulations of the ground state electron density a routine task which, in turn, permitted the high-accuracy calculations of the scattering signal from complex systems.

Extending the SC approach to the case of a time-dependent target, one could assume that the intensities of the scattered beams will reflect the dynamics of the electron density in a similar manner [23, 24, 25]: I(𝐒,t)|^𝐒[Q(𝐫,t)]|2proportional-to𝐼𝐒𝑡superscriptsubscript^𝐒delimited-[]𝑄𝐫𝑡2I(\mathbf{S},t)\propto|\hat{\mathcal{F}}_{\mathbf{S}}[Q(\mathbf{r},t)]|^{2}italic_I ( bold_S , italic_t ) ∝ | over^ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q ( bold_r , italic_t ) ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Indeed, if the duration of the probe pulse is much shorter than the time scale of the electron motion underlying the evolution of the electron density Q(𝐫,t)𝑄𝐫𝑡Q(\mathbf{r},t)italic_Q ( bold_r , italic_t ), the SC formalism seems to be a natural choice. The SC approach has been widely used in the recent past to describe the time-resolved diffraction imaging (TRDI) in atoms [26, 27], molecules [26, 28], and solids [29].

A more careful consideration of the physics behind the time-resolved scattering process revealed fundamental issues with the applicability of the SC approach to the case of the scattering from a time-dependent system. In their pioneering work [30], Dixit and co-workers utilized a consistent description of light-matter interaction in the frame of quantum electrodynamics to demonstrate that the time-resolved scattering cross section deviates significantly from its SC analogue. It has been shown [30, 31] that the scattering signal at particular time instance t𝑡titalic_t is related to a special form of the so-called density-density correlation function Ψ(t+τ/2)|ρ^(𝐫)eiH^τρ^(𝐫)|Ψ(tτ/2)quantum-operator-productΨ𝑡𝜏2^𝜌superscript𝐫superscript𝑒𝑖^𝐻𝜏^𝜌𝐫Ψ𝑡𝜏2\langle\Psi(t+\tau/2)|\hat{\rho}(\mathbf{r}^{\prime})e^{-i\hat{H}\tau}\hat{% \rho}(\mathbf{r})|\Psi(t-\tau/2)\rangle⟨ roman_Ψ ( italic_t + italic_τ / 2 ) | over^ start_ARG italic_ρ end_ARG ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_e start_POSTSUPERSCRIPT - italic_i over^ start_ARG italic_H end_ARG italic_τ end_POSTSUPERSCRIPT over^ start_ARG italic_ρ end_ARG ( bold_r ) | roman_Ψ ( italic_t - italic_τ / 2 ) ⟩, where |ΨketΨ|\Psi\rangle| roman_Ψ ⟩ denotes the time-dependent wavefunction of the system evolving according to the Hamiltonian H^^𝐻\hat{H}over^ start_ARG italic_H end_ARG, ρ^(𝐫)^𝜌𝐫\hat{\rho}(\mathbf{r})over^ start_ARG italic_ρ end_ARG ( bold_r ) and ρ^(𝐫)^𝜌superscript𝐫\hat{\rho}(\mathbf{r}^{\prime})over^ start_ARG italic_ρ end_ARG ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) are density operators, and the parameter τ𝜏\tauitalic_τ accounts for the propagation of the probe pulse through the system. This fully quantum mechanical (QM) approach to the TRDI, QM-TRDI, has been applied to probe the pure electronic dynamics in atomic [30, 32, 33, 34, 31, 35, 36, 37, 38, 39] and molecular systems [40, 41, 42, 43, 44], and also non-adiabatic molecular transitions at avoided crossings and conical intersections [45, 46, 47, 48, 49]. Interestingly, the applications of the QM-TRDI to solid state systems have not been reported yet. This is especially surprising in view of the fact that the X-ray diffraction imaging has been initially invented [50] and then very successfully used [51] to investigate the structure and properties of crystals for more than a hundred years.

In this Letter, we extend the theory underlying the QM-TRDI to the case of the laser-driven electron dynamics in solid state systems. We demonstrate that the simulations based on the QM-TRDI deviate substantially from that performed with more conventional SC-TRDI approach. We apply the developed formalism to probe the ultrafast laser-driven electron dynamics in neutral graphene, showing the possibility to trace the dynamics of the electron density, and thus demonstrate that the QM-TRDI can be a very useful technique to study ultrafast electron motion in solids.

The choice of graphene is motivated by several reasons. First of all, graphene is one of the simplest solids which makes it computationally affordable to perform fully quantum modeling of the laser-driven electron dynamics and the subsequent simulations of the diffraction signal. Second, graphene has been extensively studied theoretically and thus both the electronic structure [52] and the laser dynamics [53, 54, 55, 56] in this system are well understood. Furthermore, Yakovlev and co-workers demonstrated the application of the SC-TRDI to the case of graphene and identified certain characteristic features in their calculated diffraction signal [29]. Finally, and this is our third reason to study graphene, the electron dynamics in this system has been recently recorded experimentally by means of the attosecond electron diffraction [20]. Importantly, the measured diffraction signal contradicts with the SC-TRDI simulations made by Yakovlev et al. [29]. Therefore, graphene turns out to be a perfect system to test the validity of the QM-TRDI approach and explain the difference between the previous theoretical simulations and the experimentally measured results.

Graphene is a monolayer material composed of carbon atoms arranged in a two-dimensional hexagonal honeycomb lattice structure [52]. Each carbon atom of graphene posses six electrons while four of them participate in the formation of valence bonds. Three of these (the so-called σ𝜎\sigmaitalic_σ-electrons) form tight bonds with the neighboring atoms in the plane while the fourth electron participates in the formation of π𝜋\piitalic_π (bonding) and πsuperscript𝜋\pi^{*}italic_π start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT (anti-bonding) bonds that lie above and below of the plane formed by atomic nuclei. The unique structure and interactions of these delocalized π𝜋\piitalic_π and πsuperscript𝜋\pi^{*}italic_π start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT orbitals of graphene made this material a holy grail of solid state physics attracting ever growing interest from the community.

The electronic structure of graphene can be described using a rather simple tight-binding model, leading to analytical solutions for the band energies, related eigenstates, and the corresponding electric dipole matrix elements connecting these states with each other. Taking only π𝜋\piitalic_π electrons into account, we form the two-band system consisting of the valence and the conduction bands of graphene (see Secs. I and II of the Supplemental Material (SM) for details). These bands are highly symmetric in reciprocal space and touch each other at the so-called Dirac points forming thus conical intersections [52]. In case of graphene, the intersection points lie exactly at the Fermi level making this material a semimetal (zero gap) character and leading to an enhanced transition probability from one band to the other at these points. Accordingly, in the following we will focus on the laser-driven electron dynamics occurring in the vicinity of the Dirac points of graphene.

We calculated the evolution of the electron density in reciprocal space of graphene driven by the interaction of the system with a laser field of the following waveform: 𝐄(t)=𝐞E0sin4(πt/τ)cos(ωt)𝐄𝑡𝐞subscript𝐸0superscript4𝜋𝑡𝜏𝜔𝑡\mathbf{E}(t)=\mathbf{e}E_{0}\sin^{4}(\pi t/\tau)\cos(\omega t)bold_E ( italic_t ) = bold_e italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ( italic_π italic_t / italic_τ ) roman_cos ( italic_ω italic_t ), where the peak field amplitude E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is chosen to be 2.5 V/nm, pulse duration τ𝜏\tauitalic_τ is 21 fs, the photon energy ω𝜔\omegaitalic_ω is 1.55 eV which corresponds to the wavelength of 800 nm, and 𝐞𝐞\mathbf{e}bold_e denotes the unit vector in the direction of the field polarization. The chosen laser pulse parameters are comparable to those utilized in the recent experiment Ref. [20]. We numerically solved the semiconductor Bloch equation employing the dipole approximation and utilizing the time-dependent crystal momentum frame evolving according to the Bloch acceleration theorem [57, 58]: 𝐤t=𝐤0e/𝐀(t)subscript𝐤𝑡subscript𝐤0𝑒Planck-constant-over-2-pi𝐀𝑡\mathbf{k}_{t}=\mathbf{k}_{0}-e/\hbar\mathbf{A}(t)bold_k start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = bold_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_e / roman_ℏ bold_A ( italic_t ), where e𝑒eitalic_e is the electron charge, 𝐤tsubscript𝐤𝑡\mathbf{k}_{t}bold_k start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and 𝐤0subscript𝐤0\mathbf{k}_{0}bold_k start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT denote the time-dependent and the field-free reciprocal space vectors, respectively, and 𝐀(t)=t𝐄(t)𝑑t𝐀𝑡superscriptsubscript𝑡𝐄superscript𝑡differential-dsuperscript𝑡\mathbf{A}(t)=-\int_{-\infty}^{t}\mathbf{E}(t^{\prime})dt^{\prime}bold_A ( italic_t ) = - ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT bold_E ( italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_d italic_t start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT is the vector potential associated with the applied laser pulse 𝐄(t)𝐄𝑡\mathbf{E}(t)bold_E ( italic_t ). The detailed description of the employed computational procedure can be found in Sec. V of SM.

Refer to caption
Figure 1: Light-driven electron dynamics in graphene. (a) The blue curve represents the vector potential of the pump laser field, while the red line depicts the evolution of the population of the conduction band of graphene. (b) The snapshots of the electron density distribution in the reciprocal space of graphene in the vicinity of the Dirac point placed at the origin. The four columns depict the time instances tI=9.8subscript𝑡𝐼9.8t_{I}=9.8italic_t start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 9.8 fs, tII=11.2subscript𝑡𝐼𝐼11.2t_{II}=11.2italic_t start_POSTSUBSCRIPT italic_I italic_I end_POSTSUBSCRIPT = 11.2 fs, tIII=11.8subscript𝑡𝐼𝐼𝐼11.8t_{III}=11.8italic_t start_POSTSUBSCRIPT italic_I italic_I italic_I end_POSTSUBSCRIPT = 11.8 fs, and tIV=20.0subscript𝑡𝐼𝑉20.0t_{IV}=20.0italic_t start_POSTSUBSCRIPT italic_I italic_V end_POSTSUBSCRIPT = 20.0 fs, indicated by the vertical grey dashed lines in (a). (c), (d) and (e) The corresponding integrated real space intraband, interband, and the total electron difference densities, respectively, for the same time instances shown in (a) and (b).

The action of the external field on graphene causes population transfer between the valence and conduction bands via two distinguishable mechanisms: (i) direct photon absorption occurring in the regions of reciprocal space where the energy gap between the bands matches the photon energy ω𝜔\omegaitalic_ω, and (ii) the so-called Landau–Zener transitions [59, 60] resulting from the transfer of electrons through conical intersections in the vicinity of Dirac points. Figure 1 summarizes the results of the electron dynamics driven by the laser pulse described above and which is polarized along C–C bonds of the graphene sample (x𝑥xitalic_x-direction in our simulations). The panel (a) of Fig. 1 depicts the vector potential 𝐀(t)𝐀𝑡\mathbf{A}(t)bold_A ( italic_t ) (blue solid line) and the corresponding population of the conduction band of graphene integrated over the unit cell in the reciprocal space (red dashed line). As one can see, the population dynamics occurs continuously along the action of the field leading to about 2% electron transfer between the valence and conduction bands.

The panel (b) of Fig. 1 shows the snapshots of the matrix element ρcc(𝐤,t)subscript𝜌𝑐𝑐𝐤𝑡\rho_{cc}(\mathbf{k},t)italic_ρ start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT ( bold_k , italic_t ), where the index c𝑐citalic_c denotes the conduction band, of the reciprocal space density matrix 𝝆(𝐤,t)𝝆𝐤𝑡\bm{\rho}(\mathbf{k},t)bold_italic_ρ ( bold_k , italic_t ) in the vicinity of the Dirac point located at the origin. It is seen that the electron density is displaced towards negative values of kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT direction when the vector potential is positive (snapshot I in Fig. 1(b), taken at tI=9.8subscript𝑡𝐼9.8t_{I}=9.8italic_t start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 9.8 fs as indicated by the vertical dashed line in the panel (a)) and towards positive values of kxsubscript𝑘𝑥k_{x}italic_k start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT when the sign of the vector potential is reversed (snapshot II, tII=11.2subscript𝑡𝐼𝐼11.2t_{II}=11.2italic_t start_POSTSUBSCRIPT italic_I italic_I end_POSTSUBSCRIPT = 11.2 fs). Importantly, at the moments of time when the vector potential is zero either during the action of the field (snapshot III, tIII=11.8subscript𝑡𝐼𝐼𝐼11.8t_{III}=11.8italic_t start_POSTSUBSCRIPT italic_I italic_I italic_I end_POSTSUBSCRIPT = 11.8 fs) or when the field is over (snapshot IV, tIV=20subscript𝑡𝐼𝑉20t_{IV}=20italic_t start_POSTSUBSCRIPT italic_I italic_V end_POSTSUBSCRIPT = 20 fs), the electron density remains symmetric with respect to the vertical line passing through the conical intersection.

The calculated reciprocal space density matrix 𝝆(𝐤,t)𝝆𝐤𝑡\bm{\rho}(\mathbf{k},t)bold_italic_ρ ( bold_k , italic_t ) gives access to the real space time-dependent electron density Q(𝐫,t)𝑄𝐫𝑡Q(\mathbf{r},t)italic_Q ( bold_r , italic_t ):

Q(𝐫,t)=1𝒩n,munit cellρn,m(𝐤,t)Qm,n(𝐤t,𝐫)𝑑𝐤,𝑄𝐫𝑡1𝒩subscript𝑛𝑚subscriptunit cellsubscript𝜌𝑛𝑚𝐤𝑡subscript𝑄𝑚𝑛subscript𝐤𝑡𝐫differential-d𝐤Q(\mathbf{r},t)=\frac{1}{\mathcal{N}}\sum_{n,m}\int_{\text{unit cell}}\rho_{n,% m}(\mathbf{k},t)Q_{m,n}(\mathbf{k}_{t},\mathbf{r})d\mathbf{k},italic_Q ( bold_r , italic_t ) = divide start_ARG 1 end_ARG start_ARG caligraphic_N end_ARG ∑ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT unit cell end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT ( bold_k , italic_t ) italic_Q start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( bold_k start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , bold_r ) italic_d bold_k , (1)

where the indices n𝑛nitalic_n and m𝑚mitalic_m iterate over the valence (v𝑣vitalic_v) and conduction (c𝑐citalic_c) bands, the integration is performed over the unit cell in the reciprocal space, 𝒩𝒩\mathcal{N}caligraphic_N is the normalization factor accounting for the volume of the unit cell and the number of active electrons (𝒩=(2π)2/3a2𝒩superscript2𝜋23superscript𝑎2\mathcal{N}=(2\pi)^{2}/\sqrt{3}a^{2}caligraphic_N = ( 2 italic_π ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / square-root start_ARG 3 end_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for graphene, where a=2.46Å𝑎2.46italic-Åa=2.46\AAitalic_a = 2.46 italic_Å is the lattice constant), and Qm,n(𝐤,𝐫)subscript𝑄𝑚𝑛𝐤𝐫Q_{m,n}(\mathbf{k},\mathbf{r})italic_Q start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( bold_k , bold_r ) denote the 𝐤𝐤\mathbf{k}bold_k-resolved electron densities of π𝜋\piitalic_π/πsuperscript𝜋\pi^{*}italic_π start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT bands (m=n=v/c𝑚𝑛𝑣𝑐m=n=v/citalic_m = italic_n = italic_v / italic_c, respectively), or the corresponding transition densities connecting the bands with each other (mn𝑚𝑛m\neq nitalic_m ≠ italic_n). The total real-space electron density Q(𝐫,t)𝑄𝐫𝑡Q(\mathbf{r},t)italic_Q ( bold_r , italic_t ) can be decomposed in two contributions: Q(𝐫,t)=Qintra(𝐫,t)+Qinter(𝐫,t)𝑄𝐫𝑡subscript𝑄intra𝐫𝑡subscript𝑄inter𝐫𝑡Q(\mathbf{r},t)=Q_{\text{intra}}(\mathbf{r},t)+Q_{\text{inter}}(\mathbf{r},t)italic_Q ( bold_r , italic_t ) = italic_Q start_POSTSUBSCRIPT intra end_POSTSUBSCRIPT ( bold_r , italic_t ) + italic_Q start_POSTSUBSCRIPT inter end_POSTSUBSCRIPT ( bold_r , italic_t ), where the first term Qintra(𝐫,t)subscript𝑄intra𝐫𝑡Q_{\text{intra}}(\mathbf{r},t)italic_Q start_POSTSUBSCRIPT intra end_POSTSUBSCRIPT ( bold_r , italic_t ) represents the diagonal portion of Eq. (1) (n=m{v,c}𝑛𝑚𝑣𝑐n=m\in\{v,c\}italic_n = italic_m ∈ { italic_v , italic_c }) and thus is referred to as the intraband electron density, while the second term Qinter(𝐫,t)subscript𝑄inter𝐫𝑡Q_{\text{inter}}(\mathbf{r},t)italic_Q start_POSTSUBSCRIPT inter end_POSTSUBSCRIPT ( bold_r , italic_t ) is composed from the off-diagonal elements responsible for the coherent interband electron motion (see Sec. I of SM for the detailed derivation). In the follow-up discussion we will use the real space electron difference densities (EDDs) integrated over z𝑧zitalic_z axis for convenience: Q(x,y,t)=[Q(𝐫,t)Q(𝐫,0)]𝑑z𝑄𝑥𝑦𝑡delimited-[]𝑄𝐫𝑡𝑄𝐫0differential-d𝑧Q(x,y,t)=\int[Q(\mathbf{r},t)-Q(\mathbf{r},0)]dzitalic_Q ( italic_x , italic_y , italic_t ) = ∫ [ italic_Q ( bold_r , italic_t ) - italic_Q ( bold_r , 0 ) ] italic_d italic_z, where we removed, in addition, the static portion of the density to make the corresponding figures more representative.

The panels (c), (d), and (e) of Fig. 1 depict the snapshots of the intraband Qintra(x,y,t)subscript𝑄intra𝑥𝑦𝑡Q_{\text{intra}}(x,y,t)italic_Q start_POSTSUBSCRIPT intra end_POSTSUBSCRIPT ( italic_x , italic_y , italic_t ), interband Qinter(x,y,t)subscript𝑄inter𝑥𝑦𝑡Q_{\text{inter}}(x,y,t)italic_Q start_POSTSUBSCRIPT inter end_POSTSUBSCRIPT ( italic_x , italic_y , italic_t ), and the total Q(x,y,t)𝑄𝑥𝑦𝑡Q(x,y,t)italic_Q ( italic_x , italic_y , italic_t ) EDDs for the same time instances I, II, III, and IV discussed before. As one can see, the intraband EDDs for time instances I and II (columns I and II in Fig. 1(c)) are nearly identical to each other up to a scaling factor reflecting the difference in the overall population of the conduction band. At the same time, the corresponding reciprocal space densities (Fig. 1(b), I and II) mirror each other with respect to the y𝑦yitalic_y axis passing through the origin as we already discussed in the previous paragraph. The identical real-space EDDs for the distinguishable reciprocal space densities arise due to the symmetry of the Qm,n(𝐤,𝐫)subscript𝑄𝑚𝑛𝐤𝐫Q_{m,n}(\mathbf{k},\mathbf{r})italic_Q start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( bold_k , bold_r ) densities with respect to the Dirac point in graphene. In a difference, the corresponding interband EDDs, shown in Fig. 1(d), I and II, are antisymmetric with respect to each other. Accordingly, the total EDDs for time instances I and II (Fig. 1(e), I and II), which are combinations of the intra- and interband contributions, are also antisymmetric yet qualitatively identical to each other. The intraband EDDs for time instances III and IV are also similar to each other (Fig. 1(c), III and IV), same as the corresponding reciprocal space densities (Fig. 1(b), III and IV). We note, however, a significant difference between the intraband EDDs and the reciprocal space densities for time instances I/II and III/IV, respectively. The interband EDD at time instance IV (Fig. 1(d), IV) has nearly disappeared due to the fast decoherence of the matrix element ρvc(𝐤,t)subscript𝜌𝑣𝑐𝐤𝑡\rho_{vc}(\mathbf{k},t)italic_ρ start_POSTSUBSCRIPT italic_v italic_c end_POSTSUBSCRIPT ( bold_k , italic_t ) and its complex conjugate. As the result, the total EDD is composed almost entirely from the intraband component shortly after the action of the pulse on the system is over (see Fig. 1(e), IV). The complete movie of the laser-driven electron dynamics in graphene can be found in SM Video I.

Refer to caption
Figure 2: Comparison of the time-resolved Bragg diffraction intensities in graphene computed with the semiclassical (SC, red dashed lines) and quantum mechanical (QM, green solid lines) approaches. Top panel: Vector potential of the pump field (blue solid line). Middle and bottom panels: SC and QM diffraction intensities measured at [1,0]10[1,0][ 1 , 0 ] and [1,1]11[1,1][ 1 , 1 ] Bragg spots, respectively. Note that SC intensity is scaled by a factor of four for better visibility. The vertical grey dashed lines indicate the four time instances depicted in Fig. 1. The inset schematically shows the first Brillouin zone of graphene and the positions of Bragg spots.

Utilizing the formalism discussed above, the intensities of Bragg diffraction peaks in the SC-TRDI approximation can be expressed as

ISC(𝐒,t)=|^𝐒[Q(𝐫,t)]|2=1𝒩|n,munit cellρn,m(𝐤,t)^𝐒[Qm,n(𝐤t,𝐫)]𝑑𝐤|2,subscript𝐼SC𝐒𝑡superscriptsubscript^𝐒delimited-[]𝑄𝐫𝑡21𝒩superscriptsubscript𝑛𝑚subscriptunit cellsubscript𝜌𝑛𝑚𝐤𝑡subscript^𝐒delimited-[]subscript𝑄𝑚𝑛subscript𝐤𝑡𝐫differential-d𝐤2\begin{split}I_{\text{SC}}&(\mathbf{S},t)=\left|\mathcal{\hat{F}}_{\mathbf{S}}% [Q(\mathbf{r},t)]\right|^{2}\\ &=\frac{1}{\mathcal{N}}\left|\sum_{n,m}\int_{\text{unit cell}}\rho_{n,m}(% \mathbf{k},t)\mathcal{\hat{F}}_{\mathbf{S}}[Q_{m,n}(\mathbf{k}_{t},\mathbf{r})% ]d\mathbf{k}\right|^{2},\end{split}start_ROW start_CELL italic_I start_POSTSUBSCRIPT SC end_POSTSUBSCRIPT end_CELL start_CELL ( bold_S , italic_t ) = | over^ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q ( bold_r , italic_t ) ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = divide start_ARG 1 end_ARG start_ARG caligraphic_N end_ARG | ∑ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT unit cell end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT ( bold_k , italic_t ) over^ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( bold_k start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , bold_r ) ] italic_d bold_k | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , end_CELL end_ROW (2)

where the scattering vector 𝐒=h2πa(1/3,1)+l2πa(1/3,1)𝐒2𝜋𝑎131𝑙2𝜋𝑎131\mathbf{S}=h\frac{2\pi}{a}(1/\sqrt{3},1)+l\frac{2\pi}{a}(1/\sqrt{3},-1)bold_S = italic_h divide start_ARG 2 italic_π end_ARG start_ARG italic_a end_ARG ( 1 / square-root start_ARG 3 end_ARG , 1 ) + italic_l divide start_ARG 2 italic_π end_ARG start_ARG italic_a end_ARG ( 1 / square-root start_ARG 3 end_ARG , - 1 ) is composed from the corresponding reciprocal space vectors of graphene, and hhitalic_h and l𝑙litalic_l are integers (see Sec. III of SM for details). We simulated the SC-TRDI for the first-order Bragg peaks of graphene. Due to the high symmetry of the graphene structure, which is reflected in the corresponding symmetries of the Fourier transforms ^𝐒[Qm,n(𝐤t,𝐫)]subscript^𝐒delimited-[]subscript𝑄𝑚𝑛subscript𝐤𝑡𝐫\mathcal{\hat{F}}_{\mathbf{S}}[Q_{m,n}(\mathbf{k}_{t},\mathbf{r})]over^ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q start_POSTSUBSCRIPT italic_m , italic_n end_POSTSUBSCRIPT ( bold_k start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , bold_r ) ] contributing to Eq. (2), only two signals ISC(𝐒1,t)subscript𝐼SCsubscript𝐒1𝑡I_{\text{SC}}(\mathbf{S}_{1},t)italic_I start_POSTSUBSCRIPT SC end_POSTSUBSCRIPT ( bold_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_t ) ([h=1,l=0]delimited-[]formulae-sequence1𝑙0[h=1,l=0][ italic_h = 1 , italic_l = 0 ], [0,1]01[0,1][ 0 , 1 ], [0,1]01[0,-1][ 0 , - 1 ], [1,0]10[-1,0][ - 1 , 0 ]) and ISC(𝐒2,t)subscript𝐼SCsubscript𝐒2𝑡I_{\text{SC}}(\mathbf{S}_{2},t)italic_I start_POSTSUBSCRIPT SC end_POSTSUBSCRIPT ( bold_S start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_t ) ([1,1]11[1,1][ 1 , 1 ], [1,1]11[-1,-1][ - 1 , - 1 ]) are distinguishable from each other. Figure 2 depicts the correspondence between the applied vector potential (blue solid line in the top panel) and the simulated “normalized” (excluding the static background and the part responsible for the population transfer) SC-TRDI signal for the two families of Bragg spots ISC(𝐒1,t)subscript𝐼SCsubscript𝐒1𝑡I_{\text{SC}}(\mathbf{S}_{1},t)italic_I start_POSTSUBSCRIPT SC end_POSTSUBSCRIPT ( bold_S start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_t ) and ISC(𝐒2,t)subscript𝐼SCsubscript𝐒2𝑡I_{\text{SC}}(\mathbf{S}_{2},t)italic_I start_POSTSUBSCRIPT SC end_POSTSUBSCRIPT ( bold_S start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , italic_t ), shown by the red dashed lines in the middle and bottom panels, respectively. As one can see, the diffraction intensities at the maximum (tI=9.8subscript𝑡𝐼9.8t_{I}=9.8italic_t start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = 9.8 fs) and minimum (tII=11.2subscript𝑡𝐼𝐼11.2t_{II}=11.2italic_t start_POSTSUBSCRIPT italic_I italic_I end_POSTSUBSCRIPT = 11.2 fs) values of the vector potential have the same sign and comparable magnitudes with each other. This is again do to the symmetry of the Fourier transform: |^𝐒[Q(𝐫,t)]|2=|^𝐒[Q(𝐫,t)]|2superscriptsubscript^𝐒delimited-[]𝑄𝐫𝑡2superscriptsubscript^𝐒delimited-[]𝑄𝐫𝑡2\left|\mathcal{\hat{F}}_{\mathbf{S}}[Q(\mathbf{r},t)]\right|^{2}=\left|% \mathcal{\hat{F}}_{\mathbf{S}}[Q(-\mathbf{r},t)]\right|^{2}| over^ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q ( bold_r , italic_t ) ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = | over^ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q ( - bold_r , italic_t ) ] | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The total EDDs, as well as their components, are either fully symmetric (Qintra(x,y,t)subscript𝑄intra𝑥𝑦𝑡Q_{\text{intra}}(x,y,t)italic_Q start_POSTSUBSCRIPT intra end_POSTSUBSCRIPT ( italic_x , italic_y , italic_t ), Fig. 1(c), I and II) or antisymmetric (Qinter(x,y,t)subscript𝑄inter𝑥𝑦𝑡Q_{\text{inter}}(x,y,t)italic_Q start_POSTSUBSCRIPT inter end_POSTSUBSCRIPT ( italic_x , italic_y , italic_t ) and Q(x,y,t)𝑄𝑥𝑦𝑡Q(x,y,t)italic_Q ( italic_x , italic_y , italic_t ), Fig. 1(d) and (e), I and II) with respect to each other at time instances I and II. Accordingly, the absolute values of their Fourier transforms are comparable to each other which makes the SC-TRDI signal oscillating twice the frequency of the applied external field.

Let us turn to the simulations employing the QM-TRDI approach. Assuming the duration of the probe pulse is short and that the photon energy transfer during the scattering event is small, such that ω𝐤sω𝐤insubscript𝜔subscript𝐤ssubscript𝜔subscript𝐤in\omega_{\mathbf{k}_{\text{s}}}\approx\omega_{\mathbf{k}_{\text{in}}}italic_ω start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT s end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≈ italic_ω start_POSTSUBSCRIPT bold_k start_POSTSUBSCRIPT in end_POSTSUBSCRIPT end_POSTSUBSCRIPT, the QM-TRDI intensities can be calculated as (see Sec. IV of SM for details)

IQM(𝐒,t)=1𝒩n,m,funit cellρn,m(𝐤,t)^𝐒[Qf,m(𝐤t,𝐫)]^𝐒[Qf,n(𝐤t,𝐫)]𝑑𝐤,subscript𝐼QM𝐒𝑡1𝒩subscript𝑛𝑚𝑓subscriptunit cellsubscript𝜌𝑛𝑚𝐤𝑡subscriptsuperscript^𝐒delimited-[]subscript𝑄𝑓𝑚subscript𝐤𝑡𝐫subscript^𝐒delimited-[]subscript𝑄𝑓𝑛subscript𝐤𝑡𝐫differential-d𝐤I_{\text{QM}}(\mathbf{S},t)=\frac{1}{\mathcal{N}}\sum_{n,m,f}\int_{\text{unit % cell}}\rho_{n,m}(\mathbf{k},t)\mathcal{\hat{F}}^{*}_{\mathbf{S}}[Q_{f,m}(% \mathbf{k}_{t},\mathbf{r})]\mathcal{\hat{F}}_{\mathbf{S}}[Q_{f,n}(\mathbf{k}_{% t},\mathbf{r})]d\mathbf{k},italic_I start_POSTSUBSCRIPT QM end_POSTSUBSCRIPT ( bold_S , italic_t ) = divide start_ARG 1 end_ARG start_ARG caligraphic_N end_ARG ∑ start_POSTSUBSCRIPT italic_n , italic_m , italic_f end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT unit cell end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT ( bold_k , italic_t ) over^ start_ARG caligraphic_F end_ARG start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q start_POSTSUBSCRIPT italic_f , italic_m end_POSTSUBSCRIPT ( bold_k start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , bold_r ) ] over^ start_ARG caligraphic_F end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT [ italic_Q start_POSTSUBSCRIPT italic_f , italic_n end_POSTSUBSCRIPT ( bold_k start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT , bold_r ) ] italic_d bold_k , (3)

where the index f𝑓fitalic_f runs over the final electronic states reached by the scattering process. As one can see from Eq. (3), the resulting signal depends linearly on the reciprocal space electron densities ρn,m(𝐤,t)subscript𝜌𝑛𝑚𝐤𝑡\rho_{n,m}(\mathbf{k},t)italic_ρ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT ( bold_k , italic_t ), while in the case of the SC-TRDI, Eq. (2), the intensity is obtained as an absolute value of the sum of the corresponding quantities. Therefore, the negative (Fig. 1(b), I) and positive (Fig. 1(b), II) displacements of the electron density in the reciprocal space become spectroscopically distinguishable from each other in the frame of the QM-TRDI approach. Furthermore, the linearity of Eq. (3) makes the total diffraction signal a simple sum of the intra- and interband contributions which is not the case for the SC-TRDI, Eq. (2). Interestingly, we found that the major portion of the QM-TRDI signal comes from the intraband electron dynamics which is again resulting from the high symmetry of graphene structure (see Sec. V of SM for details). Another consequence of the symmetries of the Fourier transforms contributing to Eq. (3) in comparison to those present in the SC formalism is different relation between the diffraction signals measured at the corresponding Bragg spots: [1,0]=[0,1]=[0,1]=[1,0]10010110[1,0]=[0,1]=-[0,-1]=-[-1,0][ 1 , 0 ] = [ 0 , 1 ] = - [ 0 , - 1 ] = - [ - 1 , 0 ] and [1,1]=[1,1]1111[1,1]=-[-1,-1][ 1 , 1 ] = - [ - 1 , - 1 ]. The calculated normalized QM-TRDI signals for [1,0]10[1,0][ 1 , 0 ] and [1,1]11[1,1][ 1 , 1 ] Bragg peaks of graphene are shown by the green solid lines in the middle and the bottom panels of Fig. 2, respectively. In contrast to the SC-TRDI results, the QM-TRDI signal oscillates with the same frequency as the applied external field.

Before concluding, we emphasize that Eqs. (2) and (3) become identical in case of the elastic (f=v𝑓𝑣f=vitalic_f = italic_v) scattering from the stationary target present in the ground electronic state (ρn,m(𝐤,t)=δn=m=vsubscript𝜌𝑛𝑚𝐤𝑡subscript𝛿𝑛𝑚𝑣\rho_{n,m}(\mathbf{k},t)=\delta_{n=m=v}italic_ρ start_POSTSUBSCRIPT italic_n , italic_m end_POSTSUBSCRIPT ( bold_k , italic_t ) = italic_δ start_POSTSUBSCRIPT italic_n = italic_m = italic_v end_POSTSUBSCRIPT). In general, however, the QM-TRDI formalism is not equivalent to a straightforward extension of the SC approach to the case of the time-dependent electron density. Most importantly, the previous calculations based on the SC-TRDI approach report the diffraction intensities to be oscillating with the double the frequency of the driver laser field [29]. Our new results demonstrate, and this agrees well with the recent experimental measurements of the electron diffraction in graphene [20], that the diffraction signal should oscillate with the same frequency as the driver field. We would like to note that our simulations of the QM-TRDI in graphene are performed in the assumption that the scattering can only happen to the same electronic states forming the electronic wave packet. While this simplification could potentially lead to certain underestimation of the diffraction intensities, the quantitative conclusions about the frequency of the oscillations we made should remain valid. Furthermore, it has been shown in the recent studies of the QM-TRDI in molecules [41] that the resulting signal converges rapidly with the number of final states included in the simulations.

In conclusion, we have demonstrated the application of QM-TRDI to probe the ultrafast laser-driven electron dynamics in solid state systems. We presented the extension of the theory underlying QM-TRDI to the case of solids and derived a simple expression that connects the dynamics of electron density in the reciprocal space with the intensities of the scattered beams. We performed fully quantum simulations of the electron dynamics in neutral graphene sample and demonstrated that QM-TRDI signal deviates significantly from that obtained with the SC-TRDI approach. Finally, we identified more complicated symmetry relations between the diffraction signals measured at different Bragg spots in comparison to those present in the SC-TRDI simulations. This, in turn, could be beneficial for the corresponding experimental measurements since more information about the dynamics of the system under study is fingerprinted in the spectra. The investigation of possible computational schemes capable of recovering the dynamics of electron density from diffraction images is a promising direction of further research. We hope that our work will motivate such studies.

Acknowledgements.
The authors wish to express their sincere gratitude to Mohammed Th. Hassan for the excellent criticisms of the manuscript, and for innumerable discussions during the development of this work. The computational part of this research is based upon High Performance Computing (HPC) resources supported by the University of Arizona TRIF, UITS, and Research, Innovation, and Impact (RII) and maintained by the UArizona Research Technologies department. NVG acknowledges the financial support by the Branco Weiss Fellowship—Society in Science, administered by the ETH Zürich.

References

  • Krausz and Ivanov [2009] F. Krausz and M. Ivanov, Rev. Mod. Phys. 81, 163 (2009).
  • Goulielmakis et al. [2010] E. Goulielmakis, Z.-H. Loh, A. Wirth, R. Santra, N. Rohringer, V. S. Yakovlev, S. Zherebtsov, T. Pfeifer, A. M. Azzeer, M. F. Kling, S. R. Leone, and F. Krausz, Nature 466, 739 (2010).
  • Gong et al. [2022] X. Gong, S. Heck, D. Jelovina, C. Perry, K. Zinchenko, R. Lucchese, and H. J. Wörner, Nature 609, 507 (2022).
  • Calegari et al. [2014] F. Calegari, D. Ayuso, A. Trabattoni, L. Belshaw, S. De Camillis, S. Anumula, F. Frassetto, L. Poletto, A. Palacios, P. Decleva, J. B. Greenwood, F. Martin, and M. Nisoli, Science 346, 336 (2014).
  • Kraus et al. [2015] P. M. Kraus, B. Mignolet, D. Baykusheva, A. Rupenyan, L. Horny, E. F. Penka, G. Grassi, O. I. Tolstikhin, J. Schneider, F. Jensen, L. B. Madsen, A. D. Bandrauk, F. Remacle, and H. J. Wörner, Science 350, 790 (2015).
  • Matselyukh et al. [2022] D. T. Matselyukh, V. Despré, N. V. Golubev, A. I. Kuleff, and H. J. Wörner, Nat. Phys. 18, 1206 (2022).
  • Jordan et al. [2020] I. Jordan, M. Huppert, D. Rattenbacher, M. Peper, D. Jelovina, C. Perry, A. Von Conta, A. Schild, and H. J. Wörner, Science 369, 974 (2020).
  • Borrego-Varillas et al. [2022] R. Borrego-Varillas, M. Lucchini, and M. Nisoli, Rep. Prog. Phys. 85, 066401 (2022).
  • Hui et al. [2022] D. Hui, H. Alqattan, S. Yamada, V. Pervak, K. Yabana, and M. T. Hassan, Nat. Photonics 16, 33 (2022).
  • Gruson et al. [2016] V. Gruson, L. Barreau, Á. Jiménez-Galan, F. Risoud, J. Caillat, A. Maquet, B. Carré, F. Lepetit, J.-F. Hergott, T. Ruchon, L. Argenti, R. Taïeb, F. Martín, and P. Salières, Science 354, 734 (2016).
  • Sansone et al. [2010] G. Sansone, F. Kelkensberg, J. F. Pérez-Torres, F. Morales, M. F. Kling, W. Siu, O. Ghafur, P. Johnsson, M. Swoboda, E. Benedetti, F. Ferrari, F. Lépine, J. L. Sanz-Vicario, S. Zherebtsov, I. Znakovskaya, A. L’Huillier, M. Yu. Ivanov, M. Nisoli, F. Martín, and M. J. J. Vrakking, Nature 465, 763 (2010).
  • Cowley [1995] J. M. Cowley, Diffraction Physics, 3rd ed., North-Holland Personal Library (Elsevier Science B.V, Amsterdam ; New York, 1995).
  • Schlichting and Miao [2012] I. Schlichting and J. Miao, Curr. Opin. Struc. Biol. 22, 613 (2012).
  • Colliex et al. [2006] C. Colliex, J. M. Cowley, S. L. Dudarev, M. Fink, J. Gjønnes, R. Hilderbrandt, A. Howie, D. F. Lynch, L. M. Peng, G. Ren, A. W. Ross, V. H. Smith, J. C. H. Spence, J. W. Steeds, J. Wang, M. J. Whelan, and B. B. Zvyagin, in International Tables for Crystallography, Vol. C, edited by H. Fuess, Th. Hahn, H. Wondratschek, U. Müller, U. Shmueli, E. Prince, A. Authier, V. Kopský, D. B. Litvin, M. G. Rossmann, E. Arnold, S. Hall, B. McMahon, and E. Prince (International Union of Crystallography, Chester, England, 2006) 1st ed., pp. 259–429.
  • Carter [2016] C. B. Carter, Transmission Electron Microscopy: Diffraction, Imaging, and Spectrometry, 1st ed. (Springer Berlin Heidelberg, New York, NY, 2016).
  • Hassan et al. [2017] M. Th. Hassan, J. S. Baskin, B. Liao, and A. H. Zewail, Nature Photonics 11, 425 (2017).
  • Hassan [2018] M. T. Hassan, Journal of Physics B: Atomic, Molecular and Optical Physics 51, 032005 (2018).
  • Ishikawa et al. [2012] T. Ishikawa, H. Aoyagi, T. Asaka, Y. Asano, N. Azumi, T. Bizen, H. Ego, K. Fukami, T. Fukui, Y. Furukawa, S. Goto, H. Hanaki, T. Hara, T. Hasegawa, T. Hatsui, A. Higashiya, T. Hirono, N. Hosoda, M. Ishii, T. Inagaki, Y. Inubushi, T. Itoga, Y. Joti, M. Kago, T. Kameshima, H. Kimura, Y. Kirihara, A. Kiyomichi, T. Kobayashi, C. Kondo, T. Kudo, H. Maesaka, X. M. Maréchal, T. Masuda, S. Matsubara, T. Matsumoto, T. Matsushita, S. Matsui, M. Nagasono, N. Nariyama, H. Ohashi, T. Ohata, T. Ohshima, S. Ono, Y. Otake, C. Saji, T. Sakurai, T. Sato, K. Sawada, T. Seike, K. Shirasawa, T. Sugimoto, S. Suzuki, S. Takahashi, H. Takebe, K. Takeshita, K. Tamasaku, H. Tanaka, R. Tanaka, T. Tanaka, T. Togashi, K. Togawa, A. Tokuhisa, H. Tomizawa, K. Tono, S. Wu, M. Yabashi, M. Yamaga, A. Yamashita, K. Yanagida, C. Zhang, T. Shintake, H. Kitamura, and N. Kumagai, Nat. Photonics 6, 540 (2012).
  • Pellegrini et al. [2016] C. Pellegrini, A. Marinelli, and S. Reiche, Rev. Mod. Phys. 88, 015006 (2016).
  • Hui et al. [2023] D. Hui, H. Alqattan, M. Sennary, N. V. Golubev, and M. T. Hassan, Attosecond Electron Microscopy (2023), arXiv:2305.03014 [physics] .
  • Coppens [1992] P. Coppens, Ann. Phys. Chem. 43, 663 (1992).
  • Coppens [1997] P. Coppens, X-Ray Charge Densities and Chemical Bonding, International Union of Crystallography Texts on Crystallography No. 4 (International Union of Crystallography ; Oxford University Press, [Chester, England] : Oxford ; New York, 1997).
  • Cao and Wilson [1998] J. Cao and K. R. Wilson, J. Phys. Chem. A 102, 9523 (1998).
  • Authier [2008] A. Authier, Dynamical Theory of X-ray Diffraction, repr ed., Monographs on Crystallography No. 11 (Oxford Univ. Press, Oxford, 2008).
  • Centurion et al. [2022] M. Centurion, T. J. Wolf, and J. Yang, Ann. Phys. Chem. 73, 21 (2022).
  • Shao and Starace [2010] H.-C. Shao and A. F. Starace, Phys. Rev. Lett. 105, 263201 (2010).
  • Suominen and Kirrander [2014] H. J. Suominen and A. Kirrander, Phys. Rev. Lett. 112, 043002 (2014).
  • Baum et al. [2010] P. Baum, J. Manz, and A. Schild, Sci. China Phys. Mech. Astron. 53, 987 (2010).
  • Yakovlev et al. [2015] V. S. Yakovlev, M. I. Stockman, F. Krausz, and P. Baum, Sci. Rep. 5, 14581 (2015).
  • Dixit et al. [2012] G. Dixit, O. Vendrell, and R. Santra, Proc. Natl. Acad. Sci. 109, 11636 (2012).
  • Dixit et al. [2014] G. Dixit, J. M. Slowik, and R. Santra, Phys. Rev. A 89, 043409 (2014).
  • Dixit et al. [2013] G. Dixit, J. M. Slowik, and R. Santra, Phys. Rev. Lett. 110, 137403 (2013).
  • Dixit and Santra [2013] G. Dixit and R. Santra, J. Chem. Phys. 138, 134311 (2013).
  • Shao and Starace [2013] H.-C. Shao and A. F. Starace, Phys. Rev. A 88, 062711 (2013).
  • Shao and Starace [2014] H.-C. Shao and A. F. Starace, Phys. Rev. A 90, 032710 (2014).
  • Shao and Starace [2016] H.-C. Shao and A. F. Starace, Phys. Rev. A 94, 030702 (2016).
  • Dixit and Santra [2017] G. Dixit and R. Santra, Phys. Rev. A 96, 053413 (2017).
  • Grosser et al. [2017] M. Grosser, J. M. Slowik, and R. Santra, Phys. Rev. A 95, 062107 (2017).
  • Shao and Starace [2017] H.-C. Shao and A. F. Starace, in Ultrafast Nonlinear Imaging and Spectroscopy V, edited by Z. Liu (SPIE, San Diego, United States, 2017) p. 12.
  • Bennett et al. [2014] K. Bennett, J. D. Biggs, Y. Zhang, K. E. Dorfman, and S. Mukamel, J. Chem. Phys. 140, 204311 (2014).
  • Hermann et al. [2020] G. Hermann, V. Pohl, G. Dixit, and J. C. Tremblay, Phys. Rev. Lett. 124, 013002 (2020).
  • Giri et al. [2021] S. Giri, J. C. Tremblay, and G. Dixit, Phys. Rev. A 104, 053115 (2021).
  • Rouxel et al. [2021] J. R. Rouxel, D. Keefer, and S. Mukamel, Struct. Dyn. 8, 014101 (2021).
  • Tremblay et al. [2021] J. C. Tremblay, V. Pohl, G. Hermann, and G. Dixit, Faraday Discuss. 228, 82 (2021).
  • Kowalewski et al. [2017] M. Kowalewski, K. Bennett, and S. Mukamel, Struct. Dyn. 4, 054101 (2017).
  • Bennett et al. [2018] K. Bennett, M. Kowalewski, J. R. Rouxel, and S. Mukamel, Proc. Natl. Acad. Sci. 115, 6538 (2018).
  • Simmermacher et al. [2019] M. Simmermacher, A. Moreno Carrascosa, N. E. Henriksen, K. B. Møller, and A. Kirrander, J. Chem. Phys. 151, 174302 (2019).
  • Giri et al. [2022] S. Giri, J. C. Tremblay, and G. Dixit, Phys. Rev. A 106, 033120 (2022).
  • Tremblay et al. [2023] J. C. Tremblay, A. Blanc, P. Krause, S. Giri, and G. Dixit, ChemPhysChem 24, e202200463 (2023).
  • Friedrich et al. [1912] W. Friedrich, P. Knipping, and M. von Laue, Interferenz-Erscheinungen Bei Röntgenstrahlen (Sitzungsberichte der Mathematisch-Physikalischen Classe der Königlich-Bayerischen Akademie der Wissenschaften zu München, München, 1912).
  • Jones [2014] N. Jones, Nature 505, 602 (2014).
  • Castro Neto et al. [2009] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, Rev. Mod. Phys. 81, 109 (2009).
  • Ishikawa [2010] K. L. Ishikawa, Phys. Rev. B 82, 201402 (2010).
  • Kelardeh et al. [2015] H. K. Kelardeh, V. Apalkov, and M. I. Stockman, Phys. Rev. B 91, 045439 (2015).
  • Liu et al. [2018] C. Liu, J. Manz, K. Ohmori, C. Sommer, N. Takei, J. C. Tremblay, and Y. Zhang, Phys. Rev. Lett. 12110.1103/PhysRevLett.121.173201 (2018).
  • Morimoto et al. [2022] Y. Morimoto, Y. Shinohara, K. L. Ishikawa, and P. Hommelhoff, New J. Phys. 24, 033051 (2022).
  • Houston [1940] W. V. Houston, Phys. Rev. 57, 184 (1940).
  • Krieger and Iafrate [1986] J. B. Krieger and G. J. Iafrate, Phys. Rev. B 33, 5494 (1986).
  • Higuchi et al. [2017] T. Higuchi, C. Heide, K. Ullmann, H. B. Weber, and P. Hommelhoff, Nature 550, 224 (2017).
  • Ivakhnenko et al. [2023] O. V. Ivakhnenko, S. N. Shevchenko, and F. Nori, Phys. Rep. 995, 1 (2023).