Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Quasi-classical analysis of the dynamics of the high-order harmonic generation from solids

Open Access Open Access

Abstract

We introduce a quasi-classical model in the k space combined with the energy band structure of solids to understand the mechanisms of high-order harmonic generation (HHG) process occurring in a subcycle timescale. This model interprets the multiple plateau structure in HHG spectra well and the linear dependence of cutoff energies on the amplitude of vector potential A0 of the laser fields. It also predicts the emission time of HHG, which agrees well with the results by solving the time-dependent Schrödinger equation (TDSE). It provides a scheme to reconstruct the energy dispersion relations in Brillouin zone and to control the trajectories of HHG by varying the shape of laser pulses. This model is instructive for experimental measurements.

© 2017 Optical Society of America

Corrections

Tao-Yuan Du and Xue-Bin Bian, "Quasi-classical analysis of the dynamics of the high-order harmonic generation from solids: erratum," Opt. Express 25, 7731-7731 (2017)
https://opg.optica.org/oe/abstract.cfm?uri=oe-25-7-7731

1. Introduction

High-order harmonic generation (HHG) in atomic and molecular systems in the gas phase has been well studied theoretically and experimentally [1,2]. It can be understood by a semi-classical three-step model [3]. The bound electron may be ionized by tunneling, then driven by the external laser field. When it recombines with the parent ion, harmonics are emitted. The cutoff energy is around Ip + 3.17Up (Up is the ponderomotive energy, which is proportional to A02, where A0 is the amplitude of the vector potential A(t) of the laser fields). The HHG has resulted in the birth of attosecond (1 as = 10−18 s) pulses [4,5] and new imaging tools, such as molecular tomography [6] and spectroscopy [7]. Recent experiments have demonstrated that light-solid interactions offer a wide range of other phenomena [8–10], including HHG from solid materials [11–16]. Experimental results present a recollision feature in two-color laser fields [17], which is similar to HHG from the gas phase. However, it is also found that the cutoff of HHG from the solid phase depends on the strength E0 of the laser fields linearly [11], rather than quadratically in the gas phase. Theoretical studies [18,19] show that the cutoff of HHG from the solid phase also depends on the wavelength λ linearly, rather than λ2 in the gas phase. Multi-plateau structure in solid HHG has also been found theoretically [18,19] and experimentally [20] recently, which is quite different from HHG in the gas phase. Two-band [21] and multiband models [18,19,22], intraband [23,24] and interband transitions [25] are used to explain the mechanism behind HHG. However, they are still much debated. Simple quantitative universal models are still required to reveal the mechanisms of HHG from solids.

2. Quasi-classical model

This work introduces a quasi-classical model to investigate the electron dynamic processes under the laser fields in the wave vector k space, which is similar to the three-step model for HHG generated from the atomic and molecular systems [3]. It is also described into three steps: Zener tunneling, electron wave packet oscillation in conduction bands, and an instantaneous electron-hole pair recombination. A similar three-step picture in the coordinate space for HHG from solids is also presented in [21,26]. It requires that the displacements of the electron and the hole are the same, i.e., xc-xv = 0. However, the electron and hole are totally delocalized in solids, we replace the above condition by requiring that they have the same wave vector k in our proposed model, i.e., kc-kv = 0. Here, we make an assumption that the durations of the tunnel and recombination processes are neglected.

Due to the drive of laser fields, electrons in the valence band have probabilities to tunnel to conduction bands, i.e., Zener tunneling [23,27–29]. But the tunneling probabilities exponentially decay with the increase of energy gap. Only a small portion of electrons populated near the wave vector k = 0 on top of the valence band with minimal band gap can tunnel to conduction bands with the laser parameters used in the current work. So we choose an initial state with k0 = 0. The motion of Bloch electron wave-packet in the k space within conduction bands is given by

k(t)=k0+eA(t),

After tunneling into the conduction bands, the electrons are still attracted by the long-range periodic potentials, i.e., the energy of the electron is notp22, but  εc(k(t)) conduction band energy computed based on Bloch theorem. The HHG emissions synchronize with the Bloch electron wave packet motions. The HHG order η is determined by

η(t)=εc(k(t))εv(k(t))ω0
where ω0 is the laser field frequency, εv represents the energies of the hole in the valence band. Based on this quasi-classical model, the cutoff energy ηcutoff is determined by the maximal band gap between the conduction (including higher bands) and valence bands in the range of k ∈[0, kmax]. For ZnO, its energy band structure has been well studied as illustrated in Fig. 1(a). One may find that the energy gap ΔE between the first conduction band and the valence band is linearly dependent on k approximately as shown in Fig. 1(b). Due to the fact that the Bloch electron wave packet motion has a maximal value kmax, which is determined by the amplitude A0 from Eq. (1) if we adopt the initial k0 = 0. As a result,

 figure: Fig. 1

Fig. 1 (a) The first conduction (red line) and the valence (blue line) bands of ZnO along the Γ-M direction of the Brillouin zone from [30]. (b) The band gap between the conduction and valence bands along the Γ-M direction as a function of wave vector k. (c) Comparison of the cutoff energies of HHG from the experimental results [11] and the proposed quasi-classical model.

Download Full Size | PDF

ηcutoff=ΔεkmaxkmaxA0E0λ

In turn, once the parameters of driving fields and ηcutoff are given, the energy bands can be reconstructed by this model.

2.1 Comparison of quasi-classical model and experimental measurements

The cutoff energies by the above model compared with experimental data [11] are presented in Fig. 1(c). The energy bands of bulk crystal ZnO are reproduced from [30]. To establish the relation between electric fields of inside and outside of the solids, we need to take into account the reflection of incident pulses at the surface. The relation between the incident electric field from outside the medium, Ein, and the field in the medium, Emedium, is given by Emedium =21+nEin [21]. n represents the refractive index and is set to be 1.8992 for the laser fields with wavelength 3.25 µm adopted in the experiment. The Emedium was used in the calculations of the quasi-classical model in Fig. 1(c).

The cutoff energies in this model agree well with experimental results when the intensity is low. A small cutoff frequency discrepancy can be observed for a higher intensity. This may come from the bigger ionization rate and fast decay of laser intensity during the propagations in the crystal, which are not included in the above model. Another possible well-known reason is that accurate laser intensity is not easy to be calibrated experimentally, the standard error may be up to 10%. One can draw a conclusion that the quasi-classical model based on two bands transitions is qualitatively in accord with the experimental measurements if the error bars are added.

2.2 Comparison of quasi-classical model and TDSE simulations

The experimental data about cutoff dependence on wavelength are rarely reported. In order to further verify the above model, quantum simulations involving multiple bands are performed. We describe the light-solid interaction in one dimension, along the polarization direction of the laser fields. In the velocity-gauge treatment, the time-dependent Hamiltonian is written as,

H^(t)=H^0+emA(t)p^,
where H^0=p^22m+V(x) and V(x) is a periodic lattice potential. The specific form isV(x)=V0[1+cos(2πx/a0)], with V0 = 0.37 a.u. and lattice constant a0 = 8 a.u., which mimics the band structure of AlN [22,31]. The band gap between the conduction and valence bands is also linearly dependent on k as shown in Fig. 2. The energy band structure and time-dependent Schrödinger equation (TDSE) can be solved by using Bloch states in the k space and B-spline functions in the coordinate space, respectively. For details we refer readers to [18,19].

 figure: Fig. 2

Fig. 2 HHG spectra and field dependence of the cutoff. (a) and (b) show the whole HHG spectra calculated by Bloch states and B splines, the wavelengths of laser fields are 3.2 µm and 3.6 µm, respectively. The strength of the laser fields is fixed at 0.005 a.u. The arrows show the cutoffs calculated by the quasi-classical model. (c) and (d) show the bandgap between the conduction bands C1, C3 and the valence band, respectively. (e) and (f) show the field strength dependence of cutoff for the first and second plateaus, respectively. The wavelength of laser fields is fixed at 3.2 µm. (g) and (h) show the wavelength dependence of cutoff for the first and second plateaus, respectively. The strength of laser fields is fixed at 0.005 a.u.

Download Full Size | PDF

First, we study the harmonic spectra and field dependence of the cutoff under mid-infrared laser pulses. The electric field strength and the wavelength of the driving laser range from 0.002 a.u. to 0.006 a.u. and 2 µm to 5 µm, respectively. The harmonic spectra calculated by Bloch basis agree well with the results by B-spline basis in the coordinate space [18,19], which is shown in Fig. 2(a) and 2(b). One can find that the cutoffs of the first and second plateaus calculated by the quasi-classical model (marked by the arrows in Fig. 2(a) and 2(b)) agree with the results of TDSE well. For each electric field strength, the quasi-particle wave packet has a maximum displacement k(t)max in the k space. For the first plateau, the cutoff energy is determined by the band gap between the conduction band C1 and valence band at the k(t)max. Since the band gap has a linear dependence on k values shown in Fig. 2(c), the cutoffs depend linearly on the E0 shown in Fig. 2(e). Both of quasi-classical model and TDSE illustrate the linear dependence of the cutoff of the first plateau in Fig. 2(e), agreeing well with experimental measurements by Ghimire et al. [11]. Two or multi-plateau structure in HHG has been predicted theoretically [18,19], but not observed in experiments on crystals below the damage threshold [11,12]. However, in our previous work [32], we predicted theoretically that the second plateau may be enhanced by two-three orders in inhomogeneous fields. Multi-plateau structure in HHG from solid rare gases has been confirmed experimentally recently [20]. It is necessary to extend our proposed quasi-classical model to study higher plateau in HHG by involving multiband. In Fig. 2(e), we show the intuitive insights of the cutoff frequency of the second plateau depending linearly on the field strength. The results by the quasi-classical model and TDSE agree well. The bandgap between the conduction band C3 and valence band is linearly dependent on k approximately shown in Fig. 2(d). As a result, the cutoff frequency of the second plateau has a linear dependence on the amplitude of laser fields shown in Fig. 2(f). We also show the wavelength dependence of the cutoffs in Fig. 2(g) and 2(h). One can clearly observe the linear wavelength dependence of the cutoffs for the same reason. However, the linear dependence of cutoff on A0 is determined by the details of band structure. If the band gap between the conduction bands and valance band is not linearly dependent on k in other solid materials, the linear cutoff law in HHG will be broken.

3. HHG emission time

In order to further verify this above quasi-classical model, the underlying information in TDSE simulations must be extracted. In Fig. 3, we perform a time-frequency analysis [33,34] of HHG calculated by the above introduced TDSE solutions. In Fig. 3(b) and 3(d), the colorpanels represent the time-frequency analysis of the second and the first plateaus, respectively. The curves η(t) are the predictions of HHG emission time by the quasi-classical model in Eq. (2). One can find that the emission times of HHG predicted by the quasi-classical model and the results in TDSE agree well after taking consideration of the populations of each band presented in Fig. 3(c) into account. In the subcycle timescale, one can find the harmonic emission law. HHG photons are emitted instantaneously with the motion of quasi-particle wave packet and twice independently in a half cycle. It shows that the amplitude and waveform of laser fields control sensitively the wave packet dynamics and the harmonics emission processes. According to the emission times of the two plateaus in Fig. 3(b) and 3(d), one can find the second plateau has a time delay compared with the first plateau. The time delay indicates that the populations of the higher conduction bands are pumped from the nearest conduction band by the driving laser. In Fig. 3(c), the time delays of population on conduction bands agree with that in the time-frequency analysis well. The details of wave packet dynamics are illustrated in Fig. 4. In Fig. 4(a), it presents the wave packet dynamics of the first plateau within three steps: tunneling, Bloch electron oscillation in conduction band C1 and electron-hole pair recombination. The Bloch electron has periodic oscillations driven by laser fields and the HHG emissions follow the oscillations. In Fig. 3(d) and 3(a), one can find the Bloch electron oscillation is concurrent with harmonic emission in the time range from −1.5 o.c. to 2 o.c. It suggests that the occurrence of interband transition needs a critical strength of electric field for Zener tunneling.

 figure: Fig. 3

Fig. 3 Time-Frequency analysis of HHG. (a) is the time-dependent displacement k(t) in k space and electric field E of the laser pulses. The values have been normalized. (b) and (d) are the time-frequency analysis of the second and first plateaus in Fig. 2(a), respectively. η1, η2 and η3 represent the emission time of HHG predicted by the quasi-classical model involving transition from the conduction band C1, C2 and C3 to V, respectively. η4 represents the result involving the transition between the conduction band C3 and C1, which is similar to the transition between continuum-continuum states in the atomic systems. (c) Populations of conduction bands C1, C2 and C3. The starting times of population on conduction bands are labeled by t1, t 2, t3 values. Laser parameters are the same as those in Fig. 2(a). The envelope of the pulse is a cos2 function, and the total duration is 6 cycles. The symbol o.c. represents optical cycle.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Microscopic motion within sub-cycle timescale and electron-hole pair recombination in energy bands driven by laser pulses. (a) is the electronic dynamic processes of the first plateau including the conduction bandC1 and Valence band. (b) and (c) are the electronic dynamic processes of the second plateau including conduction bands C2 or C3 and Valence band, respectively. In (c), the recombination channel 3′ corresponds to quasi-classical trajectories η4 in Fig. 3(b). The starting times of transitions between neighbor bands are labeled by t1, 2, 3 values.

Download Full Size | PDF

In our proposed model, the population of higher conduction bands are pumped from the lower band step by step. For the second plateau, one can find the harmonic emission and wave packet dynamic processes starting from the near center of laser pulses (t2 = −0.75 o.c.) from TDSE in Fig. 3(b) and predicted from the quasi-classical model in Fig. 4(b). It can be interpreted that the Bloch electron wave packet in the conduction band C1 is driven toward the Brillouin zone boundary (BZB) at the moment t2, shown in Fig. 3(a), when the electric field strength is near zero and only a small portion of electrons can tunnel to the conduction band C2 in Fig. 4(b). This may explain that the intensity of the second plateau is 4-5 orders lower than that of the first plateau. However, the moments of population on higher bands can be tuned by chirp or changing the carrier-envelope phase (CEP) in few-cycle laser fields to increase the transition rates [19] and HHG intensity of the second plateau consequently. Furthermore, the moments of transition between conduction bands are regulated also by the change of the intensity or wavelength of the driving fields. It can also lead to the enhancement of intensity in the second plateau, as shown in Fig. 2(b) compared to Fig. 2(a). Our proposed model provides a deep understanding why there is a jump of cutoff energy when the laser intensity reaches a critical value as measured in [20].

From the moment at t3, in Fig. 3(a), electron wave packet in band C2 is driven back to the Brillouin zone center Γ within a quarter optical cycle, and the electric field strength has a maximum at the moment of the center of driving fields. So the Bloch electron can be almost fully pumped to conduction band C3 at Γ, shown in Fig. 4(c). As a result, the HHG from C2 and C3 merges together, without clear third plateau.

One can draw a conclusion that the intensity difference of the two plateaus is determined by the magnitudes of populations on C1 C2, and C3. The CEP or chirp of laser pulses can control electron dynamic processes and change the populations of high-lying conduction bands, so as to regulate the intensity of the second plateau. It can be used to interpret the experimental and theoretical results in [13,19].

4. Conclusion

In summary, this work introduces a quasi-classical model in the k space to understand intuitively the Bloch electron dynamics in HHG from solids. It can be used to interpret the phenomenon that HHG cutoffs depend linearly on both field strength and wavelength of the laser pulses, i.e., A0. It also explains why the intensity of the second plateau is very weak compared to the first plateau and how to enhance it by step-by-step excitation of higher conduction bands. It provides a promising way to realizes structural reconstruction of energy band by measuring the cutoffs. This model can deliver a prediction of harmonic emission for given solid materials and laser fields. This model has been confirmed by the TDSE calculations in the k space and coordinate space. This model can also predict the emission time of HHG. The details of dynamic processes in a subcycle timescale provide a powerful scheme to control HHG trajectories from the solids.

Funding

National Natural Science Foundation of China (NSFC) (11404376,11561121002).

Acknowledgment

We thank Tian-Jiao Shao, Mu-Zi Li and Xin-Qiang Wang for their valuable discussions.

References and links

1. T. Brabec and F. Krausz, “Intense few-cycle laser fields: Frontiers of nonlinear optics,” Rev. Mod. Phys. 72(2), 545–591 (2000). [CrossRef]  

2. F. Krausz and M. Ivanov, “Attosecond physics,” Rev. Mod. Phys. 81(1), 163–234 (2009). [CrossRef]  

3. P. B. Corkum, “Plasma perspective on strong field multiphoton ionization,” Phys. Rev. Lett. 71(13), 1994–1997 (1993). [CrossRef]   [PubMed]  

4. K. Zhao, Q. Zhang, M. Chini, Y. Wu, X. Wang, and Z. Chang, “Tailoring a 67 attosecond pulse through advantageous phase-mismatch,” Opt. Lett. 37(18), 3891–3893 (2012). [CrossRef]   [PubMed]  

5. E. Goulielmakis, M. Schultze, M. Hofstetter, V. S. Yakovlev, J. Gagnon, M. Uiberacker, A. L. Aquila, E. M. Gullikson, D. T. Attwood, R. Kienberger, F. Krausz, and U. Kleineberg, “Single-cycle nonlinear optics,” Science 320(5883), 1614–1617 (2008). [CrossRef]   [PubMed]  

6. J. Itatani, J. Levesque, D. Zeidler, H. Niikura, H. Pépin, J. C. Kieffer, P. B. Corkum, and D. M. Villeneuve, “Tomographic imaging of molecular orbitals,” Nature 432(7019), 867–871 (2004). [CrossRef]   [PubMed]  

7. H. J. Wörner, J. B. Bertrand, B. Fabre, J. Higuet, H. Ruf, A. Dubrouil, S. Patchkovskii, M. Spanner, Y. Mairesse, V. Blanchet, E. Mével, E. Constant, P. B. Corkum, and D. M. Villeneuve, “Conical intersection dynamics in NO2 probed by homodyne high-harmonic spectroscopy,” Science 334(6053), 208–212 (2011). [CrossRef]   [PubMed]  

8. M. Schultze, E. M. Bothschafter, A. Sommer, S. Holzner, W. Schweinberger, M. Fiess, M. Hofstetter, R. Kienberger, V. Apalkov, V. S. Yakovlev, M. I. Stockman, and F. Krausz, “Controlling dielectrics with the electric field of light,” Nature 493(7430), 75–78 (2012). [CrossRef]   [PubMed]  

9. A. Schiffrin, T. Paasch-Colberg, N. Karpowicz, V. Apalkov, D. Gerster, S. Mühlbrandt, M. Korbman, J. Reichert, M. Schultze, S. Holzner, J. V. Barth, R. Kienberger, R. Ernstorfer, V. S. Yakovlev, M. I. Stockman, and F. Krausz, “Optical-field-induced current in dielectrics,” Nature 493(7430), 70–74 (2013). [CrossRef]   [PubMed]  

10. M. Krüger, M. Schenk, and P. Hommelhoff, “Attosecond control of electrons emitted from a nanoscale metal tip,” Nature 475(7354), 78–81 (2011). [CrossRef]   [PubMed]  

11. 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(2), 138–141 (2011). [CrossRef]  

12. O. Schubert, M. Hohenleutner, F. Langer, B. Urbanek, C. Lange, U. Huttner, D. Golde, T. Meier, M. Kira, S. W. Koch, and R. Huber, “Sub-cycle control of terahertz high-harmonic generation by dynamical Bloch oscillations,” Nat. Photonics 8(2), 119–123 (2014). [CrossRef]  

13. T. T. Luu, M. Garg, S. Yu. Kruchinin, A. Moulet, M. Th. Hassan, and E. Goulielmakis, “Extreme ultraviolet high-harmonic spectroscopy of solids,” Nature 521(7553), 498–502 (2015). [CrossRef]   [PubMed]  

14. 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 523(7562), 572–575 (2015). [CrossRef]   [PubMed]  

15. F. Langer, M. Hohenleutner, C. P. Schmid, C. Poellmann, P. Nagler, T. Korn, C. Schüller, M. S. Sherwin, U. Huttner, J. T. Steiner, S. W. Koch, M. Kira, and R. Huber, “Lightwave-driven quasiparticle collisions on a subcycle timescale,” Nature 533(7602), 225–229 (2016). [CrossRef]   [PubMed]  

16. B. Zaks, R. B. Liu, and M. S. Sherwin, “Experimental observation of electron-hole recollisions,” Nature 483(7391), 580–583 (2012). [CrossRef]   [PubMed]  

17. 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 522(7557), 462–464 (2015). [CrossRef]   [PubMed]  

18. 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(4), 043839 (2015). [CrossRef]  

19. Z. Guan, X. X. Zhou, and X. B. Bian, “High-order-harmonic generation from periodic potentials driven by few-cycle laser pulses,” Phys. Rev. A 93(3), 033852 (2016). [CrossRef]  

20. G. Ndabashimiye, S. Ghimire, M. Wu, D. A. Browne, K. J. Schafer, M. B. Gaarde, and D. A. Reis, “Solid-state harmonics beyond the atomic limit,” Nature 534(7608), 520–523 (2016). [CrossRef]   [PubMed]  

21. 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(6), 064302 (2015). [CrossRef]  

22. P. G. Hawkins, M. Y. Ivanov, and V. S. Yakovlev, “Effect of multiple conduction bands on high-harmonic emission from dielectrics,” Phys. Rev. A 91(1), 013405 (2015). [CrossRef]  

23. P. G. Hawkins and M. Y. Ivanov, “Role of subcycle transition dynamics in high-order-harmonic generation in periodic structures,” Phys. Rev. A 87(6), 063842 (2013). [CrossRef]  

24. S. Ghimire, A. D. DiChiara, E. Sistrunk, G. Ndabashimiye, U. B. Szafruga, A. Mohammad, P. Agostini, L. F. DiMauro, and D. A. Reis, “Generation and propagation of high-order harmonics in crystals,” Phys. Rev. A 85(4), 043836 (2012). [CrossRef]  

25. C. R. McDonald, G. Vampa, P. B. Corkum, and T. Brabec, “Interband Bloch oscillation mechanism for high-harmonic generation in semiconductor crystals,” Phys. Rev. A 92(3), 033845 (2015). [CrossRef]  

26. 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(7), 073901 (2014). [CrossRef]   [PubMed]  

27. 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(7), 075330 (2008). [CrossRef]  

28. S. Glutsch, “Nonresonant and resonant Zener tunneling,” Phys. Rev. B 69(23), 235317 (2004). [CrossRef]  

29. M. S. Wismer, S. Y. Kruchinin, M. Ciappina, M. I. Stockman, and V. S. Yakovlev, “Strong-Field resonant dynamics in semiconductors,” Phys. Rev. Lett. 116(19), 197401 (2016). [CrossRef]   [PubMed]  

30. M. Goano, F. Bertazzi, M. Penna, and E. Bellotti, “Electronic structure of wurtzite ZnO: Nonlocal pseudopotential and ab initio calculations,” J. Appl. Phys. 102(8), 083709 (2007). [CrossRef]  

31. J. C. Slater, “A soluble problem in energy bands,” Phys. Rev. 87(5), 807–835 (1952). [CrossRef]  

32. T. Y. Du, Z. Guan, X. X. Zhou, and X. B. Bian, “Enhanced high-order harmonic generation from periodic potentials in inhomogeneous laser fields,” Phys. Rev. A 94(2), 023419 (2016). [CrossRef]  

33. C. Chandre, S. Wiggins, and T. Uzer, “Time-frequency analysis of chaotic systems,” Physica D 181(3-4), 171–196 (2003). [CrossRef]  

34. P. Antoine, B. Piraux, and A. Maquet, “Time profile of harmonics generated by a single atom in a strong electromagnetic field,” Phys. Rev. A 51(3), R1750–R1753 (1995). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1
Fig. 1 (a) The first conduction (red line) and the valence (blue line) bands of ZnO along the Γ-M direction of the Brillouin zone from [30]. (b) The band gap between the conduction and valence bands along the Γ-M direction as a function of wave vector k. (c) Comparison of the cutoff energies of HHG from the experimental results [11] and the proposed quasi-classical model.
Fig. 2
Fig. 2 HHG spectra and field dependence of the cutoff. (a) and (b) show the whole HHG spectra calculated by Bloch states and B splines, the wavelengths of laser fields are 3.2 µm and 3.6 µm, respectively. The strength of the laser fields is fixed at 0.005 a.u. The arrows show the cutoffs calculated by the quasi-classical model. (c) and (d) show the bandgap between the conduction bands C1, C3 and the valence band, respectively. (e) and (f) show the field strength dependence of cutoff for the first and second plateaus, respectively. The wavelength of laser fields is fixed at 3.2 µm. (g) and (h) show the wavelength dependence of cutoff for the first and second plateaus, respectively. The strength of laser fields is fixed at 0.005 a.u.
Fig. 3
Fig. 3 Time-Frequency analysis of HHG. (a) is the time-dependent displacement k(t) in k space and electric field E of the laser pulses. The values have been normalized. (b) and (d) are the time-frequency analysis of the second and first plateaus in Fig. 2(a), respectively. η1, η2 and η3 represent the emission time of HHG predicted by the quasi-classical model involving transition from the conduction band C1, C2 and C3 to V, respectively. η4 represents the result involving the transition between the conduction band C3 and C1, which is similar to the transition between continuum-continuum states in the atomic systems. (c) Populations of conduction bands C1, C2 and C3. The starting times of population on conduction bands are labeled by t1, t 2, t3 values. Laser parameters are the same as those in Fig. 2(a). The envelope of the pulse is a cos2 function, and the total duration is 6 cycles. The symbol o.c. represents optical cycle.
Fig. 4
Fig. 4 Microscopic motion within sub-cycle timescale and electron-hole pair recombination in energy bands driven by laser pulses. (a) is the electronic dynamic processes of the first plateau including the conduction bandC1 and Valence band. (b) and (c) are the electronic dynamic processes of the second plateau including conduction bands C2 or C3 and Valence band, respectively. In (c), the recombination channel 3′ corresponds to quasi-classical trajectories η4 in Fig. 3(b). The starting times of transitions between neighbor bands are labeled by t1, 2, 3 values.

Equations (4)

Equations on this page are rendered with MathJax. Learn more.

k(t)= k 0 + e A(t),
η(t)= ε c (k(t)) ε v (k(t)) ω 0
η cutoff =Δ ε k max k max A 0 E 0 λ
H ^ (t)= H ^ 0 + e m A(t) p ^ ,
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.