High order harmonic generation in semiconductors
driven at
near- and mid-IR wavelengths
Abstract
We study high order harmonics generation (HHG) in crystalline silicon and diamond subjected to near and mid-infrared laser pulses. We employ time-dependent density functional theory and solve the time-dependent Kohn-Sham equation in the single-cell geometry. We demonstrate that clear and clean HHG spectra can be generated with careful selection of the pulse duration. In addition, we simulate dephasing effects in a large silicon super-cell through a displacement of atomic positions prepared by a molecular dynamics simulation. We compare our results with the previous calculations by Floss et al. [Phys. Rev. A 97, 011401(R) (2018)] on Diamond at 800 nm and by Tancogne-Dejean et al. [Phys. Rev. Lett. 118, 087403 (2017)] on Si at 3000 nm.
I Introduction
High order harmonics generation (HHG) in solids has attracted efforts of many experimental and theoretical groups. The ultimate goal of these efforts is producing a compact and tunable source of coherent XUV and soft X-rays on a chip that can be integrated into a microelectronic device. The recent advances of HHG in solids has been reviewed in the experiment Ghimire and Reis (2019) and theory Yu et al. (2019). Previous efforts on HHG in condensed matter have been driven by improved harmonic intensities offered by the higher atomic densities of solids compared to those offered by atomic gases. Experimental studies into HHG for thin films of bulk crystal targets Ghimire et al. (2011); Schubert et al. (2014); Hohenleutner et al. (2015); Luu et al. (2015); Vampa et al. (2015a); Langer et al. (2017); Liu et al. (2017); You et al. (2017); Shirai et al. (2018); Vampa et al. (2018); Orenstein et al. (2019) have shown that improved intensities are achievable as demonstrated theoretically Ghimire and Reis (2019). Targets have also recently been expanded to structured metasurfaces Krasnok et al. (2018); Liu et al. (2018) and their 2D component layers like graphene Yoshikawa et al. (2017, 2019); Hafez et al. (2018).
Theoretical studies currently employ methods of varying quantitative power to model the interaction of solid state targets with the pulsed electric fields driving HHG. These methods include solution of the time-dependent Schrödinger equation Wu et al. (2015); Apostolova and Obreshkov (2018); Byun et al. (2021), one-dimensional models Hansen et al. (2017); Ikemachi et al. (2017); Bauer and Hansen (2018); Hansen et al. (2018); Jin et al. (2019), density matrix models Ghimire et al. (2012); Vampa et al. (2014, 2015b) and ab initio descriptions like time-dependent density functional theory (TDDFT) Otobe (2012); Floss et al. (2018, 2019); Otobe (2016); Tancogne-Dejean et al. (2017). Recently, theories have also been developed to describe HHG in more complex targets such as bulk surfaces Georges and Karatzas (2005); Golde et al. (2008); Breton et al. (2018) and metasurfaces Shcherbakov et al. (2019). Investigations using these methods have revealed the contribution of inter- and intraband mechanisms of the HHG process which are visualized in the energy band picture Vampa et al. (2014); Ghimire and Reis (2019). Similarly, the impact of band structure, band selection and laser polarization have been studied at length theoretically Langer et al. (2017); You et al. (2017); Yoshikawa et al. (2019).
Theoretical description of HHG in bulk crystalline targets requires dealing with light propagation effects. Indeed, in the experiment, a strong dephasing effect is often observed for condensed matter targets Hohenleutner et al. (2015); Luu et al. (2015); Vampa et al. (2014); Luu and Wörner (2016); Yu et al. (2016); You et al. (2017). This effect has been explored in theoretical studies Floss et al. (2018, 2019); Orlando et al. (2020) and found to be prominent in bulk solids due to the higher atomic density than found in atomic gases. However, ultra-fast dephasing times of the order of 1 fs that has been proposed in theory Kim et al. (2019); Du (2019) have posed a problem for the strong-field community. Recently, these dephasing times have been reconsidered through simulation of the linear response of the current density to broadband excitation Floss et al. (2018). This analysis proposed dephasing times of the order of 10 fs and rectified the experimental and theoretical spectra generated using TDDFT and semiconductor Bloch equations (SBE) through simulation of the laser focus propagation effect.
One of the major goals of theoretical investigations of HHG in solids is to find the optimum conditions to generate clear and strong harmonic signals. In the Maxwell-SBE calculations Floss et al. (2018), it was shown that the propagation effect is essential to obtain a clear HHG signal of high order due to the dephasing effect. Using the single-scale Maxwell-TDDFT scheme Yamada and Yabana (2021), an effort was made to find the optimal thickness of a thin film that produces the most intense HHG signals. It was shown that a very thin film of thickness of 2 - 15 nm is the optimum choice to produce intense HHG signals. In the present work, we investigate yet another unexplored factor, a time duration of the pulse that produces clear and intense HHG signals of high order. For longer driving pulses, more cycles near the maximum field strength contribute to the HHG process with the matter current getting closer to a quasi-periodic behavior thus contributing to cleaner harmonic peaks. For the purpose of our investigation, we study HHG in crystalline diamond (Di) and Si utilizing the TDDFT implementation within the computational SALMON framework Noda et al. (2019). We limit ourselves with the single cell (SC) calculations, which solve the time-dependent Kohn-Sham (TDKS) equation for the electron dynamics in TDDFT. These calculations consider the interaction between the electrons inside the bulk crystal targets and a pre-specified pulsed electric field. Within this scheme, we also investigate a dephasing effect by considering a large supercell with thermally distorted atomic positions. Calculation of HHG signals from the supercell will clarify the dephasing effect coming from electron-phonon coupling.
The paper is organized as follows: In Sec. II we describe the theoretical techniques for both time-dependent HHG and dephasing calculations. In Sec. III, we present numerical results and examine the HHG spectra. We conclude in Sec. IV by outlining the future of this method through extension to other condensed-matter systems.
II Methods
II.1 Single unit-cell method
The SC-TDDFT method considers the interaction of a bulk crystal unit cell with a spatially uniform electric field Bertsch et al. (2000); Otobe et al. (2008). This method has been used in the study of nonlinear and ultrafast dynamics Floss et al. (2018, 2019); Otobe (2012, 2016); Tancogne-Dejean et al. (2017); Yamada and Yabana (2020); Uemoto et al. (2019); Yamada and Yabana (2019); Schultze et al. (2014); Wachter et al. (2014). Utilizing the dipole approximation, the electronic motion can be described through the Bloch orbitals of the bulk, , where the wave vector is contained within the three dimensional (3D) Brillouin zone of the unit cell. The TDKS equation for such orbitals is written as:
(1) | |||
The applied electric field in this approach is defined by the vector potential of the incident pulse, . The electric potential includes the Hartree potential from the electrons and the local contribution of the ionic potential. The terms and describe the nonlocal part of the ionic pseudopotential Troullier and Martins (1991) and the exchange-correlation potential Perdew and Zunger (1981), respectively. In our calculations, this exchange-correlation potential utilizes the adiabatic local-density approximation (LDA) Onida et al. (2002). Note that the LDA underestimates the bandgap of Si, as it does for diamond and other dielectrics. The experimental direct bandgap of Si is 3.4 eV while LDA gives 2.4 eV. We neglect the exchange-correlation term in the vector potential for simplicity and do not rigorously model the exchange-correlation effects of these types for infinite periodic systems Ullrich (2012); Vignale and Kohn (1996).
The averaged electric current density, used in generating HHG spectra, is calculated as follows:
(2) | ||||
Here is the volume of the unit cell and is the band index restricted to occupied bands. The term denotes the contribution to the current density from the nonlocal part of the pseudopotential Bertsch et al. (2000) and is given as follows:
(3) |
II.2 Simulation details
We utilize the open-source software package SALMON (Scalable Ab initio Light-Matter simulator for Optics and Nanoscience) Noda et al. (2019). In this code, a uniform 3D spatial grid is defined for the electron orbitals. The time evolution of the electron orbitals is carried out using the Taylor expansion method Yabana and Bertsch (1996).
For TDDFT calculations without dephasing, we define a cubic unit cell of the diamond structure containing eight atoms with varying side lengths depending on the specific target. For Di and Si calculations, the unit cell dimensions are defined as nm and nm. A spatial grid of points is used in Si and in Di calculations to enable an accurate description of the Bloch orbitals. Convergence over spatial (r-point) grids is observed at these grid sizes. Similarly, the 3D Brillouin zone is sampled by a grid made up of -points. Note that convergence over the Brillouin zone is not observed at this grid size for diamond although qualitatively this size appears reasonable where we restrict ourselves to below the order harmonic. Otobe (2012) finds similar grid parameters for diamond but with convergence observed over the Brillouin zone. The timestep is set for stability of the calculation to be fs for Si and fs for Di. Note that tests for convergence are important in these calculations due to the discretization of the spatial, -space and time parameters discussed here.
For the Si calculation with dephasing, a supercell of size describing a atom system is used. The Brillouin zone is sampled by a grid made up of -points. Atomic positions are shifted from the equilibrium position with the diamond structure while are frozen during the time evolution of electron orbitals. They are generated from molecular dynamics simulations. The detail of the preparation will be described below.
II.3 Generating HHG spectra
In our SC-TDDFT calculations, we utilize the following incident pulse profile:
(4) |
Here is the peak amplitude of the electric field, is the carrier frequency and is the full duration of the pulse. For the envelope, the FWHM duration is related to the full duration T by . In our calculations, we consider the photon energies eV ( nm), eV ( nm) or eV ( nm). We select typical pulse durations fs which are considerably larger than in previous HHG studies ( fs in Floss et al. (2018, 2019) and 30 fs in Yamada and Yabana (2021)).
In the SC-TDDFT formalism, the HHG spectrum is calculated as a Fourier transform of the matter current
(5) | ||||
(6) | ||||
(7) |
where is the window smoothing function applied to remove spurious peaks in the spectra. A Gaussian convolution of the form
(8) |
is applied to carry out frequency averaging as often the case in experimental HHG measurements.
III Results
III.1 Bulk Si at mid-IR wavelengths




We first consider bulk Si driven by a 2000 nm laser pulse with 200 fs full duration with a peak intensity of . In Fig. 1, the time profile of the vector potential is shown in the upper panel, and the induced matter current is shown in the lower panel. Here the simulated matter current displays large noise contributions that become particularly prominent after around fs. Note also that the current is very asymmetric compared to the symmetric driving pulse centered around fs. We thus find that the SC-TDDFT method assuming a diamond structure without any distortion effects records unphysically oscillating current towards the end of the specified pulse length. In Sec. III.C, we will show that the oscillating current mostly disappears once we introduce the dephasing effect caused by the thermal motion of atoms.
Fig. 2 displays the corresponding HHG spectrum (in green) generated from the matter current shown in Fig. 1 using Eqs. (5-7). We confirmed the simulation’s convergence over spatial (r-point) and momentum (k-point) grids up to order harmonic for Si at 2000 nm. Nevertheless we show the spectrum up to order to demonstrate both the presence of, and our methods application to, higher order harmonics. Although HHG signals can be identified in the unaltered spectrum (green) even around the highest order shown in the figure, the harmonic structure appears noisy and some harmonics cannot be clearly seen in this spectrum. Gaussian convolution using Eq. (8) overcomes this problem. Figure 2 shows the spectrum with the Gaussian filter (in red).
Fig. 3 demonstrates the major effect that pulse duration has on the clarity and extent of HHG spectra, considering a nm pulse interacting with Si. The fs calculation (in green) shows clear harmonic peaks beyond the harmonic order where the fs calculation generally shows unclear harmonics until the order. In addition to this, we note that the harmonic cutoff, comparing the simulations for the fs and fs pulses, appears dependent on pulse duration. We further examined this dependence through the application of an inverse Gabor transform focused to reveal the underlying current contributions associated with the highest order harmonics. Our analysis demonstrates, as expected, the importance of multi-cycle current accumulation in the formation of harmonic structure, where we observe that a fs, nm pulse possesses too few cycles to capture the higher order returns.
Our Si results at 2000 nm (Fig. 2) and 3000 nm (Fig. 3) need to be contrasted with the earlier calculation on Si at 3000 nm presented by Tancogne-Dejean et al. (2017). In particular, using equivalent parameters to those reported in Tancogne-Dejean et al. (2017), a direct comparison is achieved for 3000 nm in Fig. 3. Tancogne-Dejean et al. (2017) recorded noisy harmonics of the lower order while their higher order harmonics appeared clean. They related this effect with the joint density of states (JDOS) as a measure of the overlap between the valence and conduction bands. A large overlap and high JDOS would promote the recombination and inter-band HHG. Somewhat counter-intuitively, Tancogne-Dejean et al. (2017) discovered that a high JDOS was in fact detrimental to HHG emission making the corresponding harmonic peaks noisy. Conversely, those harmonics falling into a low JDOS region of the HHG spectrum stayed clean. This observation led the authors to conclude that it was the intra-band mechanism that was largely responsible for HHG in silicon driven at MIR. However, our HHG spectra at both nm (Fig. 2) and nm (Fig. 3) at fs total duration are both entirely clean even without averaging. Note that we do not observe a clear and obvious JDOS effect in our calculations at both 2000 nm and 3000 nm when the pulse duration is sufficiently large.
III.2 Bulk diamond at near IR wavelengths


In Fig. 4 we display the HHG spectra of Di produced by the SC-TDDFT. The primary calculation is driven by an nm incident pulse of a fs full duration and the peak intensity of . The Gaussian convoluted spectrum again reveals clear harmonics up to order of the fundamental frequency , shown in the top panel of Fig. 4. The lower order harmonics of this spectrum are zoomed in on the bottom panel. The HHG spectrum is compared with an analogous calculation of Floss et al. (2019). The three sets of the SC-TDDFT calculations displayed in Fig. 4 are clearly differentiated by the pulse duration.
These authors employed the SC-TDDFT driven by a pulse with a total duration of fs Floss et al. (2018). Such a calculation did not produce a particularly clear HHG spectrum as is seen in Fig. 4. It is only the Maxwell propagation technique coupled with the SBE that allowed Floss et al. (2018) to generate a clear spectrum. We thus conclude that, with refined parameters and a longer pulse duration, one can obtain distinguishable high harmonics within the SC-TDDFT alone, while in previous calculations Floss et al. (2019) the Maxwell propagation or an inclusion of the explicit dephasing effect was required to obtain clear spectra that are often observed in measurements.
III.3 Dephasing Effect



We expand the SC-TDDFT method to consider the atomic system at a finite temperature by allowing the atomic positions to be displaced from equilibrium. For this purpose, we consider bulk Si and first carry out a molecular dynamics simulation with an empirical force field at a specified temperature using a thermostat. We use LAMMPS software Plimpton (1995) to carry out the molecular dynamics simulation where a three-body Stillinger-Weber potential Stillinger and Weber (1985) is used. We pick up several atomic configurations with sufficiently long intervals, and use these configurations in the SC-TDDFT. This represents a first-principles approach introducing the dephasing effect into TDDFT for the first time without relying on phenomenological parameter . This is a convenient approach where we consider that previously ultrafast dephasing times fs had to be introduced to reconcile theoretical treatments and experiment. During the time evolution stage in the SC-TDDFT, the atomic positions are fixed. We carry out calculations using supercells of different sizes, and have found that results of different configurations are quite similar to each other if we choose sufficiently large supercell. We will show below the calculation results using a supercell containing 512 atoms. Employing -points, the total number of atoms is equal to , that is equal to the total number of atoms in a cell containing 8 atoms with the -points that we showed in Fig. 2. In fact, we numerically confirmed that calculations without any distortion coincide with each other accurately between the two unit cells, 8 atoms with -points and 512 atoms with -points.
We show electron current density in bulk Si under the applied electric field in Fig. 5, caused by the same applied pulse as shown in Fig. 1. The black (blue) calculation displays the current under the atomic configuration at a temperature of 300K (180K). The matter current in disordered Si with distortion exhibits the desired dephasing after the pulse has ceased interacting. With dephasing present, signals beyond fs appear significantly suppressed and the noise contributions seen in Fig. 1 (red in Fig. 5) are removed.
Figure 6 demonstrates the effect of incorporating dephasing into the HHG calculations in Si where Gaussian convolutions are also applied. The black calculation shows the HHG spectrum using atomic positions from a molecular dynamics calculation at 300K. While signals up to order can be seen in Fig. 2 without atomic distortion, here we can see HHG signals up to order, and higher order signals cannot be identified. We also note that the HHG signals themselves become deeper between peaks and cleaner than those without atomic distortion, even when not incorporating a Gaussian convolution. Therefore, the dephasing effect modified the HHG spectrum in two ways, removing signals of HHG signals higher than order, and increasing peak-to-peak depth for those signals that survive.
In Fig. 6, the blue calculation shows the HHG spectrum using atomic positions from a molecular dynamics simulation at 180 K. Now we find HHG signals up to the order, where higher order signals are suppressed. This result at K clearly indicates that it is important to keep the material cold in measuring HHG to observe higher order signals. We however note that the distortion of atomic positions is generated by the classical molecular dynamics calculation. At very low temperatures, quantum zero-point motion will contribute significantly and should be included in discussing the HHG spectra.
The results with the effective thermally-induced dephasing effect, incorporated through atomic positions generated by molecular dynamics simulations, suggest that sample cooling could be used to greatly improve the high-order harmonic signal intensity observed in experiment. In principle, this should not pose a problem to current experiment as liquid nitrogen or helium cooling is commonly utilized in a variety of condensed matter physics contexts.
It should be noted that we have opted to consider fixed ionic coordinates in our dephasing simulations instead of performing evolution through classical mechanics. We do not incorporate an Ehrenfest molecular dynamics approach into TDDFT as we consider that spurious coherence may occur in the classical approximation when considering solids as, when one considers molecules, the coherence between two states will soon disappear as the ionic wave packets evolve differently between ionized and non-ionized electronic states and the spurious coherence that may occur in molecular simulations could extend to simulations concerning solids.
IV Concluding Remarks
We have presented here clear HHG spectra for diamond-like semiconductors interacting with near- and mid-IR pulsed electric fields. We obtain clear harmonic spectra for sufficiently long pulses with durations fs and these spectra are further clarified through Gaussian convolution. We therefore review the suggestions made in previous literature considering diamond Floss et al. (2018, 2019) and silicon Tancogne-Dejean et al. (2017). In crystalline Si, we do not observe the expected joint density of states effect but do demonstrate a prominent dephasing effect on the HHG spectra. Using a first-principles approach introducing dephasing into TDDFT we observe that higher-order harmonics are strongly suppressed by the decoherence effects in the target whilst lower order harmonics appear clearer in the spectra. On the basis of the strength of the decoherence effects for Si at nm laser wavelengths, we suggest that helium or liquid nitrogen cooling of the target will improve higher order harmonic intensity returns.
V Acknowledgements
The authors would like to acknowledge the two super computer resources employed for TDDFT calculations: the National Computational Infrastructure (NCI Australia) and Fugaku at RIKEN R-CCS, Japan, with support through the HPCI System Research Project (Project ID: hp210137). This research was partially supported by JST-CREST under Grant No. JP-MJCR16N5, by MEXT Quantum Leap Flagship Program (MEXT Q-LEAP) under Grant No. JPMXS0118068681, and by JSPS KAKENHI Grant No. 20H02649.
References
- Ghimire and Reis (2019) S. Ghimire and D. A. Reis, High-harmonic generation from solids, Nature Physics 15, 10 (2019).
- Yu et al. (2019) C. Yu, S. Jiang, and R. Lu, High order harmonic generation in solids: a review on recent numerical methods, Advances in Physics: X 4, 1562982 (2019).
- Ghimire et al. (2011) S. Ghimire, A. D. DiChiara, E. Sistrunk, P. Agostini, L. F. DiMauro, and D. A. Reis, Observation of high-order harmonic generation in a bulk crystal, Nat. Phys. 7, 138 (2011).
- Schubert et al. (2014) O. Schubert, M. Hohenleutner, F. Langer, B. Urbanek, C. Lange, U. Huttner, D. Golde, T. Meier, M. Kira, S. W. Koch, et al., Sub-cycle control of terahertz high-harmonic generation by dynamical bloch oscillations, Nat. Photonics 8, 119 (2014).
- Hohenleutner et al. (2015) M. Hohenleutner, F. Langer, O. Schubert, M. Knorr, U. Huttner, S. W. Koch, M. Kira, and R. Huber, Real-time observation of interfering crystal electrons in high-harmonic generation, Nature (London) 523, 572 (2015).
- Luu et al. (2015) T. T. Luu, M. Garg, S. Y. Kruchinin, A. Moulet, M. T. Hassan, and E. Goulielmakis, Extreme ultraviolet high-harmonic spectroscopy of solids, Nature (London) 521, 498 (2015).
- Vampa et al. (2015a) G. Vampa, T. J. Hammond, N. Thiré, B. E. Schmidt, F. Légaré, C. R. McDonald, T. Brabec, and P. B. Corkum, Linking high harmonics from gases and solids, Nature (London) 522, 462 (2015a).
- Langer et al. (2017) F. Langer, M. Hohenleutner, U. Huttner, S. W. Koch, M. Kira, and R. Huber, Symmetry-controlled time structure of high-harmonic carrier fields from a solid, Nat. Photonics 11, 227 (2017).
- Liu et al. (2017) H. Liu, Y. Li, Y. S. You, S. Ghimire, T. F. Heinz, and D. A. Reis, High-harmonic generation from an atomically thin semiconductor, Nat. Phys. 13, 262 (2017).
- You et al. (2017) Y. S. You, D. A. Reis, and S. Ghimire, Anisotropic high-harmonic generation in bulk crystals, Nat. Phys. 13, 345 (2017).
- Shirai et al. (2018) H. Shirai, F. Kumaki, Y. Nomura, and T. Fuji, High-harmonic generation in solids driven by subcycle midinfrared pulses from two-color filamentation, Opt. Lett. 43, 2094 (2018).
- Vampa et al. (2018) G. Vampa, T. J. Hammond, M. Taucer, X. Ding, X. Ropagnol, T. Ozaki, S. Delprat, M. Chaker, N. Thiré, B. E. Schmidt, et al., Strong-field optoelectronics in solids, Nat. Photonics 12, 465 (2018).
- Orenstein et al. (2019) G. Orenstein, A. J. Uzan, S. Gadasi, T. Arusi-Parpar, M. Krüger, R. Cireasa, B. D. Bruner, and N. Dudovich, Shaping electron-hole trajectories for solid-state high harmonic generation control, Opt. Express 27, 37835 (2019).
- Krasnok et al. (2018) A. Krasnok, M. Tymchenko, and A. Alù, Nonlinear metasurfaces: a paradigm shift in nonlinear optics, Materials Today 21, 8 (2018).
- Liu et al. (2018) H. Liu, C. Guo, G. Vampa, J. L. Zhang, T. Sarmiento, M. Xiao, P. H. Bucksbaum, J. Vukovi, S. Fan, and D. A. Reis, Enhanced high-harmonic generation from an all-dielectric metasurface, Nat. Phys. 14, 1006 (2018).
- Yoshikawa et al. (2017) N. Yoshikawa, T. Tamaya, and K. Tanaka, High-harmonic generation in graphene enhanced by elliptically polarized light excitation, Science 356, 736 (2017).
- Yoshikawa et al. (2019) N. Yoshikawa, K. Nagai, K. Uchida, Y. Takaguchi, S. Sasaki, Y. Miyata, and K. Tanaka, Interband resonant high-harmonic generation by valley polarized electron–hole pairs, Nat. Commun. 10, 3709 (2019).
- Hafez et al. (2018) H. A. Hafez, S. Kovalev, J.-C. Deinert, Z. Mics, B. Green, N. Awari, M. Chen, S. Germanskiy, U. Lehnert, J. Teichert, et al., Extremely efficient terahertz high-harmonic generation in graphene by hot dirac fermions, Nature 561, 507 (2018).
- Wu et al. (2015) M. Wu, S. Ghimire, D. A. Reis, K. J. Schafer, and M. B. Gaarde, High-harmonic generation from bloch electrons in solids, Phys. Rev. A 91, 043839 (2015).
- Apostolova and Obreshkov (2018) T. Apostolova and B. Obreshkov, Ultrafast photoionization and excitation of surface-plasmon-polaritons on diamond surfaces, Diam. Relat. Mater. 82, 165 (2018).
- Byun et al. (2021) C. W. Byun, M.-H. Lee, and N. N. Choi, High-order harmonic generation from solids using houston states, J. of Kor. Phys. Soc. 78, 662 (2021).
- Hansen et al. (2017) K. K. Hansen, T. Deffge, and D. Bauer, High-order harmonic generation in solid slabs beyond the single-active-electron approximation, Phys. Rev. A 96, 053418 (2017).
- Ikemachi et al. (2017) T. Ikemachi, Y. Shinohara, T. Sato, J. Yumoto, M. KuwataGonokami, and K. L. Ishikawa, Trajectory analysis of high-order-harmonic generation from periodic crystals, Phys. Rev. A 95, 043416 (2017).
- Bauer and Hansen (2018) D. Bauer and K. K. Hansen, High-harmonic generation in solids with and without topological edge states, Phys. Rev. Lett. 120, 177401 (2018).
- Hansen et al. (2018) K. K. Hansen, D. Bauer, and L. B. Madsen, Finite-system effects on high-order harmonic generation: From atoms to solids, Phys. Rev. A 97, 043424 (2018).
- Jin et al. (2019) J.-Z. Jin, H. Liang, X.-R. Xiao, M.-X. Wang, S.-G. Chen, X.-Y. Wu, Q. Gong, and L.-Y. Peng, Contribution of floquet-bloch states to high-order harmonic generation in solids, Phys. Rev. A 100, 013412 (2019).
- Ghimire et al. (2012) S. Ghimire, A. D. DiChiara, E. Sistrunk, G. Ndabashimiye, U. B. Szafruga, P. A. A. Mohammad, L. F. DiMauro, and D. A. Reis, Generation and propagation of high-order harmonics in crystals, Phys. Rev. A 85, 043836 (2012).
- Vampa et al. (2014) G. Vampa, C. R. McDonald, G. Orlando, D. D. Klug, P. B. Corkum, and T. Brabec, Theoretical analysis of high-harmonic generation in solids, Phys. Rev. Lett. 113, 073901 (2014).
- Vampa et al. (2015b) G. Vampa, C. R. McDonald, G. Orlando, P. B. Corkum, and T. Brabec, Semiclassical analysis of high harmonic generation in bulk crystals, Phys. Rev. B 91, 064302 (2015b).
- Otobe (2012) T. Otobe, First-principle description for the high-harmonic generation in a diamond by intense short laser pulse, J. App. Phys. 111, 093112 (2012).
- Floss et al. (2018) I. Floss, C. Lemell, G. Wachter, V. Smejkal, S. A. Sato, X.-M. Tong, K. Yabana, and J. Burgdörfer, Ab initio multiscale simulation of high-order harmonic generation in solids, Phys. Rev. A 97, 011401(R) (2018).
- Floss et al. (2019) I. Floss, C. Lemell, K. Yabana, and J. Burgdörfer, Incorporating decoherence into solid-state time-dependent density functional theory, Phys. Rev. B 99, 224301 (2019).
- Otobe (2016) T. Otobe, High-harmonic generation in alpha-quartz by electron-hole recombination, Phys. Rev. B 94, 235152 (2016).
- Tancogne-Dejean et al. (2017) N. Tancogne-Dejean, O. D. Mücke, F. X. Kärtner, and A. Rubio, Impact of the electronic band structure in high-harmonic generation spectra of solids, Phys. Rev. Lett. 118, 087403 (2017).
- Georges and Karatzas (2005) A. T. Georges and N. E. Karatzas, Theory of multiple harmonic generation in reflection from a metal surface, Appl. Phys. B 81, 479 (2005).
- Golde et al. (2008) D. Golde, T. Meier, and S. W. Koch, High harmonics generated in semiconductor nanostructures by the coupled dynamics of optical inter- and intraband excitations, Phys. Rev. B 77, 075330 (2008).
- Breton et al. (2018) G. L. Breton, A. Rubio, and N. Tancogne-Dejean, High-harmonic generation from few-layer hexagonal boron nitride: Evolution from monolayer to bulk response, Phys. Rev. B 98, 165308 (2018).
- Shcherbakov et al. (2019) M. R. Shcherbakov, K. Werner, Z. Fan, N. Talisa, E. Chowdhury, and G. Shvets, Photon acceleration and tunable broadband harmonics generation in nonlinear time-dependent metasurfaces, Nature Comm. 10, 1345 (2019).
- Luu and Wörner (2016) T. T. Luu and H. J. Wörner, High-order harmonic generation in solids: a unifying approach, Phys. Rev. B 94, 115164 (2016).
- Yu et al. (2016) C. Yu, X. Zhang, S. Jiang, X. Cao, G. Yuan, L. B. T. Wu, and R. Lu, Dependence of high-order-harmonic generation on dipole moment in sio2 crystals, Phys. Rev. A 94, 013846 (2016).
- Orlando et al. (2020) G. Orlando, T.-S. Ho, and S.-I. Chu, Simple model of dephasing for high-order harmonic generation in solids, J. of Opt. Soc. of Amer. B 37, 1540 (2020).
- Kim et al. (2019) Y. W. Kim, T.-J. Shao, H. Kim, S. Han, S. Kim, M. Ciappina, X.-B. Bian, and S.-W. Kim, Spectral interference in high harmonic generation from solids, ACS Phot. 6, 851 (2019).
- Du (2019) T.-Y. Du, Probing the dephasing time of crystals via spectral properties of high-order harmonic generation, Phys. Rev. A 100, 053401 (2019).
- Yamada and Yabana (2021) S. Yamada and K. Yabana, Determining the optimum thickness for high harmonic generation from nanoscale thin films: An ab initio computational study, Phys. Rev. B 103, 155426 (2021).
- Noda et al. (2019) M. Noda, S. A. Sato, Y. Hirokawa, M. Uemoto, T. Takeuchi, S. Yamada, A. Yamada, Y. Shinohara, M. Yamaguchi, K. Iida, et al., Salmon: Scalable ab-initio light-matter simulator for optics and nanoscience, Comput. Phys. Commun. 235, 356 (2019).
- Bertsch et al. (2000) G. F. Bertsch, J.-I. Iwata, A. Rubio, and K. Yabana, Real-space, real-time method for the dielectric function, Phys. Rev. B 62, 7998 (2000).
- Otobe et al. (2008) T. Otobe, M. Yamagiwa, J.-I. Iwata, K. Yabana, T. Nakatsukasa, and G. F. Bertsch, First-principles electron dynamics simulation for optical breakdown of dielectrics under an intense laser field, Phys. Rev. B 77, 165104 (2008).
- Yamada and Yabana (2020) S. Yamada and K. Yabana, Symmetry properties of attosecond transient absorption spectroscopy in crystalline dielectrics, Phys. Rev. B 101, 165128 (2020).
- Uemoto et al. (2019) M. Uemoto, Y. Kuwabara, S. A. Sato, and K. Yabana, Nonlinear polarization evolution using time-dependent density functional theory, J. Chem. Phys. 150, 094101 (2019).
- Yamada and Yabana (2019) A. Yamada and K. Yabana, Energy transfer from intense laser pulse to dielectrics in time-dependent density functional theory, Eur. Phys. J. D 73, 87 (2019).
- Schultze et al. (2014) M. Schultze, K. Ramasesha, C. Pemmaraju, S. Sato, D. Whitmore, A. Gandman, J. S. Prell, L. J. Borja, D. Prendergast, K. Yabana, et al., Attosecond band-gap dynamics in silicon, Science 346, 1348 (2014).
- Wachter et al. (2014) G. Wachter, C. Lemell, J. Burgdörfer, S. A. Sato, X.-M. Tong, and K. Yabana, Ab initio simulation of electrical currents induced by ultrafast laser excitation of dielectric materials, Phys. Rev. Lett. 113, 087401 (2014).
- Troullier and Martins (1991) N. Troullier and J. L. Martins, Efficient pseudopotentials for plane-wave calculations, Phys. Rev. B 43, 1993 (1991).
- Perdew and Zunger (1981) J. P. Perdew and A. Zunger, Self-interaction correction to density-functional approximations for many-electron systems, Phys. Rev. B 23, 5048 (1981).
- Onida et al. (2002) G. Onida, L. Reining, and A. Rubio, Electronic excitations: density-functional versus many-body green’s-function approaches, Rev. Mod. Phys. 74, 601 (2002).
- Ullrich (2012) C. A. Ullrich, Time-Dependent Density-Functional Theory: Concepts and Applications, Oxford Graduate Texts (Oxford University Press, New York, 2012).
- Vignale and Kohn (1996) G. Vignale and W. Kohn, Current-dependent exchange-correlation potential for dynamical linear response theory, Phys. Rev. Lett. 77, 2037 (1996).
- Yabana and Bertsch (1996) K. Yabana and G. F. Bertsch, Time-dependent local-density approximatin in real time, Phys. Rev. B 54, 4484 (1996).
- Plimpton (1995) S. Plimpton, Fast parallel algorithms for short-range molecular dynamics, J. Comp. Phys. 117(1), 1 (1995).
- Stillinger and Weber (1985) F. H. Stillinger and T. A. Weber, Computer simulation of local order in condensed phases of silicon, Phys. Rev. B 31, 5262 (1985).