Next Article in Journal
One-Parameter Hyperbolic Dual Spherical Movements and Timelike Ruled Surfaces
Previous Article in Journal
Modified Generalized Feistel Network Block Cipher for the Internet of Things
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Waveform Modulation of High-Order Harmonics Generated from an Atom Irradiated by a Laser Pulse and a Weak Orthogonal Electrostatic Field

1
Institute of Atomic and Molecular Physics, Jilin University, Changchun 130012, China
2
Jilin Provincial Key Laboratory of Applied Atomic and Molecular Spectroscopy, Jilin University, Changchun 130012, China
3
Zhejiang Provincial Key Laboratory for Cutting Tools, Taizhou University, Taizhou 318000, China
*
Author to whom correspondence should be addressed.
Symmetry 2023, 15(4), 901; https://doi.org/10.3390/sym15040901
Submission received: 3 March 2023 / Revised: 6 April 2023 / Accepted: 6 April 2023 / Published: 12 April 2023
(This article belongs to the Section Physics)

Abstract

:
The detailed characteristics of the harmonics emission of atoms driven via a linearly polarized laser field combined with an orthogonal, weaker electrostatic field were investigated by numerically solving the time-dependent Schrödinger equation. It was found that the direction of the laser polarization and the polarization of the attosecond light, which is synthesized from the harmonic, can be controlled by the amplitude of the electrostatic field. With the analysis of the spatial distribution of the time-dependent dipole moment and the time-dependent evolution of the electronic wave packet, the control mechanism for the harmonic characters was investigated. The generation of harmonics in the vertical direction of the laser electric field is caused by the breaking of the symmetry of the time wave packet distribution. With this mechanism, we obtained circularly polarized attosecond light.

1. Introduction

When atoms, molecules, and solids are irradiated by a strong laser pulse, one can observe the high-order harmonics generation of the driving pulse [1,2,3]. Ultrashort pulses generated by harmonics can be used to detect the dynamic evolution of electrons in matter [4,5,6,7,8,9,10,11,12,13,14]. In addition, molecular electron orbitals can be imaged through tomography using high-order harmonic radiation [10,15,16,17].
The mechanism for high-order harmonics generated from atoms and molecules driven by a linearly polarized laser can be explained by the three-step model. In the model, the atom is ionized firstly by the intense laser pulse through the tunnel progress, then the motion of the ionized electron is dominated by the laser electric field, and part of the ionized electron has an opportunity to recombine with the parent ion. The harmonics is generated in the recombining process of the ionized electron [3,18,19,20,21,22,23,24,25,26,27,28,29,30,31]. Due to the linear polarization features of the driving laser pulse, the polarization of the high-order harmonics generated from the atom is usually the same as the driving laser.
However, there are many applications that require a specific polarization for the harmonic light source [32,33,34,35,36,37,38]. For example, in the ultrafast research of magnetic materials, ultrashort pulses with a circular polarization are needed. Due to the large spectral widths of the ultrashort pulses, it is very difficult to directly change the polarization of the harmonic light source. Therefore, schemes for the generation of the required polarized harmonic radiation from the variation on the driving light field or gas target have been proposed.
Mairesse et al. observed the elliptically polarized harmonic radiation from the N 2 molecules irradiated by a linearly polarized laser pulse [39,40]. Using the combination of a Ti sapphire laser and its second harmonic, Becker et al. demonstrated the circular polarization characteristics of the harmonics generated from atoms [41]. Recently, the circularly polarized harmonic is generated from the atom irradiated by a counter–rotating bichromatic circularly polarized laser pulse. The circularly polarized harmonic is used as a probe for the chiral matter [1,42,43,44,45,46,47,48]. In this work, the electrostatic fields were used to control the polarization of the harmonics.
The combined pulse schemes, which include the electrostatic or the THz fields, had been proposed to control the behaviors of the harmonics [41,49,50,51,52,53,54]. Bao et al. studied the harmonics generation from the combined pulse including the electrostatic field and the laser electric field. They found that the even–order harmonic radiation appeared in the plateau part of the harmonic spectra [53,55,56]. Yuan et al. investigated the harmonics generated from H 2 + irradiated by a few–cycle elliptically polarized laser pulse combined with a terahertz pulse. They obtained the circularly polarized harmonic and the circularly polarized attosecond pulses [54,57,58,59,60,61,62].
However, the proposed schemes cannot control the spectral characteristics of the high–order harmonics due to the lack of understanding of the generation mechanism for the harmonic polarization. In this work, the modulation of the immediate behavior of the high–order harmonics can be controlled by the addition of an orthogonal electrostatic field. More importantly, the physical mechanism of the variation in the harmonic behavior can be explained by the symmetry breaking of the electronic wave packet motion. According to the understanding of the mechanism, we modulated the field amplitude of the electrostatic field to obtain circularly polarized light. The organization of this paper is as follows. In Section 2, we give the theoretical approaches to calculate high–order harmonic spectra. In Section 3, the harmonic spectra are presented and their generation mechanism is analyzed through the evolution of the wave packet. Our main conclusions are summarized in Section 4. Atomic units are used throughout unless otherwise specified.

2. Model and Methods

In the electric dipole approximation and the length gauge, the time–dependent Schrödinger equation for the electrons of atoms irradiated by a laser pulse is [63,64,65,66,67,68]
i t ψ ( x , y , t ) = H ( x , y , t ) ψ ( x , y , t )
The Hamiltonian in Equation (1) is
H ( x , y , t ) = p 2 2 + V ( x , y ) + x E x ( t ) + y E y ( t ) = p x 2 + p y 2 2 Q ( x 2 + y 2 + a 2 ) + x E x ( t ) + y E y ( t )
V ( x , y ) is the soft–core Coulomb potential. Nuclear charge Q is equal to 1. The soft–core parameter a = 0.07 was selected, and the ground condition energy of this system is −24.587 eV (the ground condition energy of helium atoms). E x ( t ) and E y ( t ) are the electronic fields in the x–orientation and the driving domain in the y–orientation, respectively; E 1 is the linearly polarized laser pulse field, and E is a combination of ultrashort linearly polarized laser pulses and weak electrostatic fields, orthogonal to each other. The forms are as follows:
E 1 = E y ( t ) = E 0 y f ( t ) s i n ( ω t )
E = E x ( t ) + E y ( t ) = E 0 x + E 0 y f ( t ) s i n ( ω t )
where E 0 x is the electrostatic field amplitude and E 0 y and ω are the peak amplitude, as well as the angular frequency of the laser pulse electric domain, respectively. In this paper, we define E 2 for E 0 x = 0.003, E 3 for E 0 x = 0.005 and E 4 for E 0 x = 0.0375. The equation f ( t ) is as follows: f ( t ) = s i n 2 ( ω t 2 N ) , where N is the cycle number of laser pulses. Two optical periods were adopted in this paper. The wave function at any time is acquired by solving Equation (1) using the numerical scheme of the split operator [69]. On this basis, the dipole acceleration a x ( t ) and a y ( t ) in the two directions can be calculated:
a x ( t ) = d 2 d t 2 ψ ( x , y , t ) | x | ψ ( x , y , t ) = ψ ( x , y , t ) | V ( x , y ) x E x ( t ) | ψ ( x , y , t ) a y ( t ) = d 2 d t 2 ψ ( x , y , t ) | y | ψ ( x , y , t ) = ψ ( x , y , t ) | V ( x , y ) y E y ( t ) | ψ ( x , y , t )
Then, the harmonic radiation power of the harmonics emission spectra P x ( ω ) 2 and P y ( ω ) 2 is counted via the Fourier transform of this time–dependent dipole acceleration:
P x ( ω ) = 1 ω 2 ( t n t 0 ) t i t f a x ( t ) exp ( i ω t ) d t P y ( ω ) = 1 ω 2 ( t n t 0 ) t i t f a y ( t ) exp ( i ω t ) d t
Ultrashort attosecond pulses can be generated by the coherent superposition of the selected harmonics in a certain energy range. The time–dependent intensity of the attosecond pulse can be expressed as
I ( t ) = | q a q exp ( i q ω t ) | 2
q is the order of the harmonics; a q is
a q = a ( t ) exp ( i q ω t ) d t
where a ( t ) represents the dipole acceleration.

3. Results and Discussion

We first calculated the higher–order harmonics generated by atoms driven by a combined ultrashort laser pulse and a pulse with an orthogonal weak electrostatic field. The corresponding schematic diagram is shown in Figure 1a. The x–direction harmonic spectra are presented in Figure 1a. It was found that there was no harmonics emission generated from the atom irradiated by the laser field E 1 . For the harmonics generated from the combined pulse, the cutoff order was the 62nd harmonic. Since the driving laser is the ultrashort pulse, there were no clear odd harmonics that can be found in the harmonic spectrum. In the figure, one can observe the interference structure produced by the two emission trajectories generated in half an optical cycle. The y–direction harmonic spectra are presented in Figure 1b. It was found that there is a high–intensity harmonic emitted from the atom driven by E 1 and E 3 . Compared with the harmonics generated from E 3 , the harmonics emission cutoff energy and efficiency remained almost unchanged in the emission spectra generated from the atom driven by E 1 . In the high–energy part of the harmonics emission plateau, more oscillatory structures with small amplitudes can be observed. In the low–energy part of the harmonics plateau, the peak position of the intra–period intervening shifts to the high energy, as shown in the inset of Figure 1b.
In addition to the change in harmonic intensity, the differences in the harmonic waveform can be also observed. In order to study the change of the harmonic waveform more visually, we present a synthetic attosecond pulse generated from the harmonic spectra whose energy was changed from the 45st to the 60th harmonic in Figure 2. In Figure 2a, the attosecond pulse was synthesized from the harmonics generated from the atom driven by E 1 . It was found that the polarization of the attosecond pulse was linearly polarized. With the employment of the weak electric field whose amplitude was 0.003, the attosecond pulse exhibited an elliptical polarization characteristic with little ellipticity. More importantly, the ellipticity of the attosecond waveform can be controlled by the electrostatic field amplitude, as is shown in Figure 2c. From the above analysis, one knows that the addition of an orthogonal electrostatic field provides an opportunity to control the harmonic waveform.
In order to better control the harmonic waveform, one needs to analyze deeply the harmonics emission behavior of atoms irradiated by the driving field. The harmonic spectra of the atom were calculated from the time–dependent dipole moment. The dipole moment of the system is the expected value of the dipole transition operator. In the spatial region, the calculation of the time–dependent dipole moment was divided into two parts: a x + ( t ) and a x ( t ) . The dipole moments were calculated from the x > 0 regions and the x < 0 regions, respectively.
a x + ( t ) = + 0 + ψ ( x , y , t ) ψ * ( x , y , t ) ( V ( x , y ) x E x ( t ) ) d x d y a x ( t ) = + 0 ψ ( x , y , t ) ψ * ( x , y , t ) ( V ( x , y ) x E x ( t ) ) d x d y
The corresponding harmonic spectra are | P x + ( ω ) | 2 , | P x ( ω ) | 2 , respectively. The total harmonic power can be expressed as | P ( ω ) | 2 = | P x + ( ω ) + P x ( ω ) | 2 . Here,
P x + ( ω ) = 1 ω 2 ( t n t 0 ) t i t f a x + ( t ) exp ( i ω t ) d t P x ( ω ) = 1 ω 2 ( t n t 0 ) t i t f a x ( t ) exp ( i ω t ) d t
The corresponding harmonic phase ϕ x and ϕ x + can be written as:
ϕ x = arctan ( P x i m ( ω ) / P x r e ( ω ) ) , ϕ x + = arctan ( P x + i m ( ω ) / P x + r e ( ω ) ) ,
where P x i m ( ω ) ( P x + i m ( ω ) and P x r e ( ω ) ( P x + r e ( ω ) represent the imaginary part and the real part of the harmonics emission in the x < 0 regions ( x > 0 regions), respectively. The phase difference between the two spatial regions is
Δ ϕ x = ϕ x ϕ x +
One can quantitatively analyze the mechanism of the harmonic waveform control using harmonic analysis in different spatial regions. Figure 3a presents the x–directional harmonic spectra generated from the atom in field E 1 in different spatial regions. It was found that the intensities of the harmonic spectra in different spatial regions are the same. However, the intensity of the harmonic spectra in the whole spatial range is zero. The disappearance of the harmonics emission in the x–direction is caused by the interference of the harmonics generated in the two parts of space. It can be proven by the phase difference between these two parts of the harmonics emission, as shown in Figure 3b. Furthermore, the harmonics emission analysis of the atom in the E 2 field was performed, as shown in Figure 3c. It was found that the harmonic intensity of the two spatial regions was also close to being the same. The intensity difference of the harmonic spectra is presented in Figure 3b, which is marked as a pink solid line. The intensity difference is larger than the intensity calculated from the whole spatial range. This intensity weakening can also be explained by the interference between the harmonics generated in the two spatial regions. The corresponding phase difference is present in Figure 3c,d. Next, we quantitatively studied the interference effects of the harmonics in different spatial regions.
For the quantitative analysis, we defined Δ P = | P x ( ω ) | | P x + ( ω ) | , P = | P ( ω ) | 2 and Δ ϕ x = ϕ x ϕ x + . The total harmonic power of the system can be written as
| P ( ω ) | 2 = | P x + ( ω ) + P x ( ω ) | 2 = | P x + ( ω ) | 2 + | P x ( ω ) | 2 + 2 | P x + ( ω ) | | P x ( ω ) | cos Δ ϕ x = 2 | P x + ( ω ) | 2 ( 1 + cos Δ ϕ x ) + 2 Δ P | P x + ( ω ) | ( 1 + cos Δ ϕ x ) + Δ P 2 = P 1 + P 2 + P 3
Here, P 1 = 2 | P x + ( ω ) | 2 ( 1 + cos Δ ϕ x ) , P 2 = 2 Δ P | P x + ( ω ) | ( 1 + cos Δ ϕ x ) , P 3 = Δ P 2 .
The terms P 1 and P 2 are associated with the phase difference of the harmonics generated from different spatial regions, and P 3 is only associated with the intensity difference of the amplitude of harmonics emission. Figure 4a presents P, P 1 , and the harmonic spectra calculated from the TDSE. It was found that the P 1 term had little effect on the harmonics emission. P 3 played a dominant role in the harmonic spectra, as is shown in Figure 4b. Through the above analysis, it was found that the symmetry of the dipole moment had an important impact on the harmonic yield. If this symmetry is broken, clear harmonics emission will be observed in the vertical direction of the driven laser field. Next, we analyzed the reason for the symmetry breaking by using the time–dependent evolution of the wave packet.
The electron density distributions at different instants are present in Figure 5. The instants are (a,e) t 1 = 1.29 o.c., (b,f) t 2 = 1.36 o.c., (c,g) t 3 = 1.43 o.c., and (d,g) t 4 = 1.50 o.c. The red solid line is the laser electric field, and the corresponding moments are marked with yellow dots. The distributions are symmetrical in the x–direction when the driven field is E 1 , as is shown in Figure 5a–d. It is clearly observed that the wave packet re–collides with the parent ion. Interference structures generated by ionized and re–collision wave packets can be observed. For the driven field E 2 , the electronic density distributions are not symmetrical in the x–direction. The motion of electrons is affected by the electrostatic field. Therefore, the magnitudes of the harmonic spectra in the two spatial regions are no longer symmetrical. This leads to the generation of the harmonics in the vertical direction of the laser electric field.
By the above analysis, when the electrostatic field was not added in the x–direction, the electron wave packet density was symmetric about the y–axis, and there were no harmonics in the x–direction, so the synthetic line polarization laser was in the y–direction. However, when an electrostatic field was added in the x–direction, the symmetry of the original wave function was destroyed, so the x–polarization direction produced the harmonics. Furthermore, by adjusting the amplitude of the electrostatic field, the adjustment of the harmonic intensity in the x–direction can be realized. Therefore, under a certain electrostatic intensity amplitude condition, harmonics emissions with close intensity in two directions can be produced. After the electrostatic field and the driving light field act together at the moment of the harmonics emission, the direction of electron movement returning to the parent ion was deflected, and the corresponding phase difference was 0.5 π , and a circularly polarized ultrashort pulse emission could be observed. In this process, by adjusting the electrostatic field, it was possible to achieve a phase difference of 0.5 π between the phases of certain order harmonics, which led to the production of circularly polarized ultrashort pulses. Other orders of the harmonics emission spectrum would generate ultrashort pulses close to linear polarization. When the amplitude of electrostatic field was 0.0375, the harmonic spectrum was as presented in Figure 6a. By using the harmonics emission ranging from the 40th to the 47th harmonic, one can obtain an ultrashort pulse whose polarization is close to circular polarization. The amplitude evolution of the ultrashort pulse with time is presented in Figure 6b.

4. Conclusions

In summary, we theoretically studied the high–order harmonics generated from an atom irradiated by an electric field, which included a laser field and an orthogonal weak electrostatic field. It was found that the electrostatic field broke the symmetry of the electron wave packet density distribution and generated harmonics perpendicular to the laser polarization. Further, modulating the amplitude of the electrostatic field can well control the waveform of the harmonics. The variety of the harmonic waveform was determined by the motion of the ionized electrons. Applying this mechanism, a circularly polarized attosecond light pulse was generated. The finding of this work provides a simple and reliable method to achieve the optimization of the harmonic waveform.

Author Contributions

T.F.: data curation, writing—original draft, editing, and software. F.G.: supervision. J.W.: supervision. J.C.: supervision. Y.Y.: conceptualization and supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The project was supported by the National Key Research and Development Program of China (2019YFA0307700), the National Natural Science Foundation of China (NSFC) (Grant No. 12074145 and No. 11975012) and Jilin Provincial Research Foundation for Basic Research, China (Grant No. 20220101003JC).

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

We acknowledge the High Performance Computing Center of Jilin University for the supercomputer time and the high-performance computing cluster Tiger@ IAMP.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. L’Huillier, A.; Schafer, K.J.; Kulander, K.C. Theoretical aspects of intense field harmonics generation. J. Phys. B At. Mol. Opt. Phys. 1991, 24, 3315. [Google Scholar] [CrossRef]
  2. McPherson, A.; Gibson, G.; Jara, H.; Johann, U.; Luk, T.S.; McIntyre, I.; Boyer, K.; Rhodes, C.K. Studies of multiphoton production of vacuum-ultraviolet radiation in the rare gases. JOSA B 1987, 4, 595–601. [Google Scholar] [CrossRef]
  3. Krause, J.L.; Schafer, K.J.; Kulander, K.C. High-order harmonics generation from atoms and ions in the high intensity regime. Phys. Rev. Lett. 1992, 68, 3535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Wang, J.; Chen, G.; Li, S.Y.; Ding, D.J.; Chen, J.G.; Guo, F.M.; Yang, Y.J. Ultrashort-attosecond-pulse generation by reducing harmonic chirp with a spatially inhomogeneous electric field. Phys. Rev. A 2015, 92, 033848. [Google Scholar] [CrossRef]
  5. Chen, J.G.; Zeng, S.L.; Yang, Y.J. Generation of isolated sub-50-as pulses by quantum path control in the multicycle regime. Phys. Rev. A 2010, 82, 043401. [Google Scholar] [CrossRef]
  6. Krausz, F.; Ivanov, M. Attosecond physics. Rev. Mod. Phys. 2009, 81, 163. [Google Scholar] [CrossRef] [Green Version]
  7. Li, P.C.; Zhou, X.X.; Wang, G.L.; Zhao, Z.X. Isolated sub-30-as pulse generation of an He+ ion by an intense few-cycle chirped laser and its high-order harmonic pulses. Phys. Rev. A 2009, 80, 053825. [Google Scholar] [CrossRef]
  8. Vozzi, C.; Negro, M.; Calegari, F.; Sansone, G.; Nisoli, M.; De Silvestri, S.; Stagira, S. Generalized molecular orbital tomography. Nat. Phys. 2011, 7, 822–826. [Google Scholar] [CrossRef]
  9. Haessler, S.; Caillat, J.; Boutu, W.; Giovanetti-Teixeira, C.; Ruchon, T.; Auguste, T.; Diveki, Z.; Breger, P.; Maquet, A.; Carré, B.; et al. Attosecond imaging of molecular electronic wavepackets. Nat. Phys. 2010, 6, 200–206. [Google Scholar] [CrossRef] [Green Version]
  10. Itatani, J.; Levesque, J.; Zeidler, D.; Niikura, H.; Pépin, H.; Kieffer, J.C.; Corkum, P.B.; Villeneuve, D.M. Tomographic imaging of molecular orbitals. Nature 2004, 432, 867–871. [Google Scholar] [CrossRef]
  11. Salamin, Y.I.; Hu, S.; Hatsagortsyan, K.Z.; Keitel, C.H. Relativistic high-power laser–matter interactions. Phys. Rep. 2006, 427, 41–155. [Google Scholar] [CrossRef]
  12. van der Zwan, E.V.; Chirilă, C.C.; Lein, M. Molecular orbital tomography using short laser pulses. Phys. Rev. A 2008, 78, 033410. [Google Scholar] [CrossRef]
  13. Wong, M.; Le, A.T.; Alharbi, A.; Boguslavskiy, A.; Lucchese, R.; Brichta, J.P.; Lin, C.; Bhardwaj, V. High harmonic spectroscopy of the cooper minimum in molecules. Phys. Rev. Lett. 2013, 110, 033006. [Google Scholar] [CrossRef] [Green Version]
  14. Schafer, K.J.; Gaarde, M.B.; Heinrich, A.; Biegert, J.; Keller, U. Strong field quantum path control using attosecond pulse trains. Phys. Rev. Lett. 2004, 92, 023003. [Google Scholar] [CrossRef] [Green Version]
  15. Chen, Y.; Fu, L.; Liu, J. Asymmetric molecular imaging through decoding odd-even high-order harmonics. Phys. Rev. Lett. 2013, 111, 073902. [Google Scholar]
  16. Milošević, D.B.; Starace, A.F. High-order harmonics generation in magnetic and parallel magnetic and electric fields. Phys. Rev. A 1999, 60, 3160. [Google Scholar] [CrossRef] [Green Version]
  17. Bertrand, J.B.; Wörner, H.J.; Hockett, P.; Villeneuve, D.M.; Corkum, P.B. Revealing the Cooper minimum of N2 by Molecular Frame High-Harmonic Spectroscopy. Phys. Rev. Lett. 2012, 109, 143001. [Google Scholar] [CrossRef] [Green Version]
  18. Corkum, P.B. Plasma perspective on strong field multiphoton ionization. Phys. Rev. Lett. 1993, 71, 1994. [Google Scholar]
  19. Lewenstein, M.; Balcou, P.; Ivanov, M.Y.; L’huillier, A.; Corkum, P.B. Theory of high-harmonics generation by low-frequency laser fields. Phys. Rev. A 1994, 49, 2117. [Google Scholar]
  20. Yang, Y.J.; Chen, J.G.; Chi, F.P.; Zhu, Q.R.; Zhang, H.X.; Sun, J.Z. Ultrahigh harmonics generation from an atom with superposition of ground state and highly excited states. Chin. Phys. Lett. 2007, 24, 1537–1540. [Google Scholar]
  21. Yu-Jun, Y.; Gao, C.; Ji-Gen, C.; Qi-Ren, Z. A time-distinguished analysis of the harmonic structure from a model molecular ion. Chin. Phys. Lett. 2004, 21, 652. [Google Scholar] [CrossRef]
  22. Tian, Y.Y.; Li, S.Y.; Wei, S.S.; Guo, F.M.; Zeng, S.L.; Chen, J.G.; Yang, Y.J. Investigation on the influence of atomic potentials on the above threshold ionization. Chin. Phys. B 2014, 23, 053202. [Google Scholar]
  23. Wang, J.; Chen, G.; Guo, F.M.; Li, S.Y.; Chen, J.G.; Yang, Y.J. High-intensity molecular harmonics generation without ionization. Chin. Phys. B 2013, 22, 033203. [Google Scholar] [CrossRef]
  24. Yuan, H.; Yang, Y.; Guo, F.; Wang, J.; Cui, Z. Influence of multiphoton resonance excitation on the above-threshold ionization of a hydrogen atom. Opt. Express 2022, 30, 19745–19756. [Google Scholar] [CrossRef] [PubMed]
  25. Yan, J.Z.; Zhao, S.S.; Lan, W.D.; Li, S.Y.; Zhou, S.S.; Chen, J.G.; Zhang, J.Y.; Yang, Y.J. Calculation of high-order harmonics generation of atoms and molecules by combining time series prediction and neural networks. Opt. Express 2022, 30, 35444–35456. [Google Scholar] [CrossRef]
  26. Lan, W.; Wang, X.; Qiao, Y.; Zhou, S.; Chen, J.; Wang, J.; Guo, F.; Yang, Y. High-Intensity Harmonic Generation with Energy Tunability Produced by Robust Two-Color Linearly Polarized Laser Fields. Symmetry 2023, 15, 580. [Google Scholar] [CrossRef]
  27. Zhou, X.; Lock, R.; Li, W.; Wagner, N.; Murnane, M.M.; Kapteyn, H.C. Molecular recollision interferometry in high harmonics generation. Phys. Rev. Lett. 2008, 100, 073902. [Google Scholar] [CrossRef]
  28. Qiao, Y.; Huo, Y.Q.; Jiang, S.C.; Yang, Y.J.; Chen, J.G. All-optical reconstruction of three-band transition dipole moments by the crystal harmonic spectrum from a two-color laser pulse. Opt. Express 2022, 30, 9971–9982. [Google Scholar] [CrossRef]
  29. Qiao, Y.; Huo, Y.; Liang, H.; Chen, J.; Liu, W.; Yang, Y.; Jiang, S. Robust retrieval method of crystal transition dipole moments by high-order harmonic spectrum. Phys. Rev. B 2023, 107, 075201. [Google Scholar] [CrossRef]
  30. Qiao, Y.; Chen, J.; Huo, Y.; Liang, H.; Yu, R.; Chen, J.; Liu, W.; Jiang, S.; Yang, Y. Effect of the interference between interband currents on the crystal harmonic spectra. Phys. Rev. A 2023, 107, 023523. [Google Scholar] [CrossRef]
  31. Wei, S.; Pan, Y.; Qiao, Y.; Zhou, S.; Yuan, H.; Wang, J.; Guo, F.; Yang, Y. Theoretical Simulation of the High—Order Harmonic Generated from Neon Atom Irradiated by the Intense Laser Pulse. Symmetry 2023, 15, 636. [Google Scholar] [CrossRef]
  32. Porat, G.; Alon, G.; Rozen, S.; Pedatzur, O.; Krüger, M.; Azoury, D.; Natan, A.; Orenstein, G.; Bruner, B.; Vrakking, M.; et al. Attosecond time-resolved photoelectron holography. Nat. Commun. 2018, 9, 2805. [Google Scholar] [CrossRef] [Green Version]
  33. Ge, X.L.; Wang, T.; Guo, J.; Liu, X.S. Quantum-path control and isolated-attosecond-pulse generation using H 2+ molecules with moving nuclei in few-cycle laser pulses. Phys. Rev. A 2014, 89, 023424. [Google Scholar] [CrossRef]
  34. Hernández-García, C.; Picón, A.; San Román, J.; Plaja, L. Attosecond extreme ultraviolet vortices from high-order harmonic generation. Phys. Rev. Lett. 2013, 111, 083602. [Google Scholar] [CrossRef] [Green Version]
  35. Li, J.; Lu, J.; Chew, A.; Han, S.; Li, J.; Wu, Y.; Wang, H.; Ghimire, S.; Chang, Z. Attosecond science based on high harmonics generation from gases and solids. Nat. Commun. 2020, 11, 2748. [Google Scholar] [CrossRef]
  36. Kim, K.T.; Villeneuve, D.; Corkum, P. Manipulating quantum paths for novel attosecond measurement methods. Nat. Photonics 2014, 8, 187–194. [Google Scholar] [CrossRef]
  37. Lépine, F.; Ivanov, M.Y.; Vrakking, M.J. Attosecond molecular dynamics: Fact or fiction? Nat. Photonics 2014, 8, 195–204. [Google Scholar] [CrossRef]
  38. Calegari, F.; Ayuso, D.; Trabattoni, A.; Belshaw, L.; De Camillis, S.; Anumula, S.; Frassetto, F.; Poletto, L.; Palacios, A.; Decleva, P.; et al. Ultrafast electron dynamics in phenylalanine initiated by attosecond pulses. Science 2014, 346, 336–339. [Google Scholar] [CrossRef] [Green Version]
  39. Mairesse, Y.; Higuet, J.; Dudovich, N.; Shafir, D.; Fabre, B.; Mével, E.; Constant, E.; Patchkovskii, S.; Walters, Z.; Ivanov, M.Y.; et al. High harmonic spectroscopy of multichannel dynamics in strong-field ionization. Phys. Rev. Lett. 2010, 104, 213601. [Google Scholar] [CrossRef]
  40. Zhou, S.S.; Yang, Y.J.; Guo, F.M.; Chen, J.G.; Wang, J. Multielectron effect for high-order harmonics generation from molecule irradiated by bichromatic counter-rotating circularly polarized laser pulses. IEEE J. Quantum Electron. 2020, 56, 1–7. [Google Scholar]
  41. Eichmann, H.; Egbert, A.; Nolte, S.; Momma, C.; Wellegehausen, B.; Becker, W.; Long, S.; McIver, J. Polarization-dependent high-order two-color mixing. Phys. Rev. A 1995, 51, R3414. [Google Scholar] [CrossRef] [PubMed]
  42. Heslar, J.; Telnov, D.A.; Chu, S.I. Generation of circularly polarized XUV and soft-x-ray high-order harmonics by homonuclear and heteronuclear diatomic molecules subject to bichromatic counter-rotating circularly polarized intense laser fields. Phys. Rev. A 2017, 96, 063404. [Google Scholar] [CrossRef]
  43. Wang, B.; Li, X.; Fu, P. Polarization effects in high-harmonics generation in the presence of static-electric field. Phys. Rev. A 1999, 59, 2894. [Google Scholar] [CrossRef]
  44. Milošević, D.B.; Becker, W.; Kopold, R. Generation of circularly polarized high-order harmonics by two-color coplanar field mixing. Phys. Rev. A 2000, 61, 063403. [Google Scholar]
  45. Kfir, O.; Grychtol, P.; Turgut, E.; Knut, R.; Zusin, D.; Popmintchev, D.; Popmintchev, T.; Nembach, H.; Shaw, J.M.; Fleischer, A.; et al. Generation of bright phase-matched circularly-polarized extreme ultraviolet high harmonics. Nat. Photonics 2015, 9, 99–105. [Google Scholar]
  46. Qiao, Y.; Wu, D.; Chen, J.G.; Wang, J.; Guo, F.M.; Yang, Y.J. High-order harmonics generation from H 2+ irradiated by a co-rotating two-color circularly polarized laser field. Phys. Rev. A 2019, 100, 063428. [Google Scholar] [CrossRef]
  47. Huang, C.; Zhong, M.; Wu, Z. Nonsequential double ionization by co-rotating two-color circularly polarized laser fields. Opt. Express 2019, 27, 7616–7626. [Google Scholar] [CrossRef]
  48. Sharma, P.; Agrawal, E.; Jha, P. Generation of circularly polarized harmonic radiation by two color laser beams propagating in plasma. Phys. Plasmas 2018, 25, 093102. [Google Scholar] [CrossRef]
  49. Bandrauk, A.D.; Yuan, K.J. Circularly polarised attosecond pulses: Generation and applications. Mol. Phys. 2016, 114, 344–355. [Google Scholar] [CrossRef]
  50. Ge, X.L. The comparison of high-order harmonics generation of H in a circularly polarized laser pulse combined with a terahertz field and with a static electric field. Laser Phys. Lett. 2020, 17, 055301. [Google Scholar] [CrossRef]
  51. Taranukhin, V.D.; Shubin, N.Y. High-order harmonics generation by the atoms in strong bichromatic fields. Quantum Electron. 1999, 29, 638. [Google Scholar]
  52. Taranukhin, V.D.; Shubin, N.Y. High-order harmonics generation by atoms with strong high-frequency and low-frequency pumping. JOSA B 2000, 17, 1509–1516. [Google Scholar] [CrossRef]
  53. Bao, M.Q.; Starace, A.F. Static-electric-field effects on high harmonics generation. Phys. Rev. A 1996, 53, R3723. [Google Scholar] [CrossRef] [Green Version]
  54. Yuan, K.J.; Bandrauk, A.D. Single circularly polarized attosecond pulse generation by intense few cycle elliptically polarized laser pulses and terahertz fields from molecular media. Phys. Rev. Lett. 2013, 110, 023003. [Google Scholar]
  55. Lein, M.; Hay, N.; Velotta, R.; Marangos, J.P.; Knight, P. Role of the intramolecular phase in high-harmonics generation. Phys. Rev. Lett. 2002, 88, 183903. [Google Scholar] [CrossRef] [Green Version]
  56. Song, S.; Wu, L.; Qiao, Y.; Zhou, S.; Wang, J.; Guo, F.; Yang, Y. Investigation of the Spatio-Temporal Characteristics of High-Order Harmonic Generation Using a Bohmian Trajectory Scheme. Symmetry 2023, 15, 581. [Google Scholar] [CrossRef]
  57. Camper, A.; Ferré, A.; Blanchet, V.; Descamps, D.; Lin, N.; Petit, S.; Lucchese, R.; Salières, P.; Ruchon, T.; Mairesse, Y. Quantum-Path Resolved Attosecond High-Harmonic Spectroscopy. Phys. Rev. Lett. 2023, 130, 083201. [Google Scholar]
  58. Li, W.K.; Lei, Y.; Li, X.; Yang, T.; Du, M.; Jiang, Y.; Li, J.L.; Luo, S.Z.; Liu, A.H.; He, L.H.; et al. Multiphoton ionization of potassium atoms in femtosecond laser fields. Chinese Phys. Lett. 2021, 38, 053202. [Google Scholar]
  59. Qiao, Y.; Chen, J.Q.; Chen, J.G. Review on the Reconstruction of Transition Dipole Moments by Solid Harmonic Spectrum. Symmetry 2022, 14, 2646. [Google Scholar] [CrossRef]
  60. Peng, Q.F.; Peng, Z.Y.; Lang, Y.; Zhu, Y.L.; Zhang, D.W.; Lü, Z.H.; Zhao, Z.X. Decoherence Effects of Terahertz Generation in Solids under Two-Color Femtosecond Laser Fields. Chinese Phys. Lett. 2022, 39, 053301. [Google Scholar]
  61. Zhu, Y.P.; Liu, Y.R.; Zhao, X.; Kimberg, V.; Zhang, S.B. Core-excited molecules by resonant intense x-ray pulses involving electron-rotation coupling. Chinese Phys. Lett. 2021, 38, 053201. [Google Scholar] [CrossRef]
  62. Li, X.H.; Ding, B.C.; Feng, Y.F.; Wu, R.C.; Tian, L.F.; Huang, J.Y.; Liu, X.J. High Energy Inner Shell Photoelectron Diffraction in CO2. Chinese Phys. Lett. 2022, 39, 033401. [Google Scholar] [CrossRef]
  63. Guo, F.-M.; Yang, Y.-J.; Jin, M.-X.; Ding, D.-J.; Zhu, Q.-R. A theoretical strategy to generate an isolated 80-attosecond pulse. Chin. Phys. Lett. 2009, 26, 053201. [Google Scholar] [CrossRef]
  64. Jiang, T.F.; Chu, S.I. High-order harmonics generation in atomic hydrogen at 248 nm: Dipole-moment versus acceleration spectrum. Phys. Rev. A 1992, 46, 7322. [Google Scholar] [CrossRef] [Green Version]
  65. Hermann, M.R.; Fleck, J., Jr. Split-operator spectral method for solving the time-dependent Schrödinger equation in spherical coordinates. Phys. Rev. A 1988, 38, 6000. [Google Scholar] [CrossRef]
  66. Han, J.X.; Wang, J.; Qiao, Y.; Liu, A.H.; Guo, F.M.; Yang, Y.J. Significantly enhanced conversion efficiency of high-order harmonic generation by introducing chirped laser pulses into scheme of spatially inhomogeneous field. Opt. Express 2019, 27, 8768–8776. [Google Scholar] [CrossRef]
  67. Han, Y.C.; Madsen, L.B. Comparison between length and velocity gauges in quantum simulations of high-order harmonics generation. Phys. Rev. A 2010, 81, 063430. [Google Scholar] [CrossRef]
  68. Kjeldsen, T.K.; Madsen, L.B. Strong-field ionization of N2: Length and velocity gauge strong-field approximation and tunnelling theory. J. Phys. B At. Mol. Opt. Phys. 2004, 37, 2033. [Google Scholar] [CrossRef]
  69. Tong, X.M.; Chu, S.I. Probing the spectral and temporal structures of high-order harmonic generation in intense laser pulses. Phys. Rev. A 2000, 61, 021802. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) Sketch of the atom interacting with the combined laser field. (b) The black solid line of the spectrum represents the x–direction ( E 1 x ); the red solid line indicates the HHG spectrum from E 1 y when the harmonics emission is obtained by the laser pulse. The black dashed line represents the x–direction ( E 3 x ), and the red dashed line indicates the HHG spectrum from E 3 y when the harmonics emission is obtained by the orthogonal field. The inset in (b) exhibits the 30st–40th order harmonic spectrum. The laser field peak amplitude is 0.13, and the amplitude of the electrostatic amplitude of E 3 y is 0.005.
Figure 1. (a) Sketch of the atom interacting with the combined laser field. (b) The black solid line of the spectrum represents the x–direction ( E 1 x ); the red solid line indicates the HHG spectrum from E 1 y when the harmonics emission is obtained by the laser pulse. The black dashed line represents the x–direction ( E 3 x ), and the red dashed line indicates the HHG spectrum from E 3 y when the harmonics emission is obtained by the orthogonal field. The inset in (b) exhibits the 30st–40th order harmonic spectrum. The laser field peak amplitude is 0.13, and the amplitude of the electrostatic amplitude of E 3 y is 0.005.
Symmetry 15 00901 g001
Figure 2. The evolution of time in the range of 1.14 o.c. to 1.44 o.c. at the time of different combined field effects, the attosecond waveform plot for the synthesis of harmonic electric fields in the x– and y–polarization directions. In Figures (ac), E x ( t ) and E y ( t ) are the harmonic electric fields generated in the polarization direction of the weak electrostatic field (x) and the harmonic electric fields generated in the polarization direction of the laser (y), respectively. (a) Only when driving laser action ( E 1 ), (b) exhibits the combined field ( E 2 ) in which the driving laser polarization direction is orthogonal to the x–polarization direction as the electrostatic field with a field amplitude of 0.003 acting together. (c) illustrates the combined field ( E 3 ) in which the driving laser polarization direction is orthogonal to the x–polarization direction as the electrostatic field acting together with a field amplitude of 0.005. In the figure, the green points (red points) represent the time–dependent evolution of the harmonic electric field in the y (x) polarization direction, and the blue points represent the harmonic electric field diagram in the x–y plane. The solid red line is the time–dependent evolution of the harmonic electric field synthesized by the two polarization directions.
Figure 2. The evolution of time in the range of 1.14 o.c. to 1.44 o.c. at the time of different combined field effects, the attosecond waveform plot for the synthesis of harmonic electric fields in the x– and y–polarization directions. In Figures (ac), E x ( t ) and E y ( t ) are the harmonic electric fields generated in the polarization direction of the weak electrostatic field (x) and the harmonic electric fields generated in the polarization direction of the laser (y), respectively. (a) Only when driving laser action ( E 1 ), (b) exhibits the combined field ( E 2 ) in which the driving laser polarization direction is orthogonal to the x–polarization direction as the electrostatic field with a field amplitude of 0.003 acting together. (c) illustrates the combined field ( E 3 ) in which the driving laser polarization direction is orthogonal to the x–polarization direction as the electrostatic field acting together with a field amplitude of 0.005. In the figure, the green points (red points) represent the time–dependent evolution of the harmonic electric field in the y (x) polarization direction, and the blue points represent the harmonic electric field diagram in the x–y plane. The solid red line is the time–dependent evolution of the harmonic electric field synthesized by the two polarization directions.
Symmetry 15 00901 g002
Figure 3. (a) The harmonic spectra of the linearly polarized laser distribution in the x < 0 areas (black solid line), the x > 0 areas (red dotted line), as well as the full space (blue solid line). (b) The harmonic spectra from E 2 in the x < 0 areas (black solid line), the x > 0 areas (red dotted line), and the full space (blue solid line). The solid pink line is the distinction between the harmonic spectra from the left, as well as the right space parts. (c,d) The harmonic phase distinction between the left, as well as the right parts of E 1 (c) and the harmonic phase difference between the left and right partitions by E 2 (d).
Figure 3. (a) The harmonic spectra of the linearly polarized laser distribution in the x < 0 areas (black solid line), the x > 0 areas (red dotted line), as well as the full space (blue solid line). (b) The harmonic spectra from E 2 in the x < 0 areas (black solid line), the x > 0 areas (red dotted line), and the full space (blue solid line). The solid pink line is the distinction between the harmonic spectra from the left, as well as the right space parts. (c,d) The harmonic phase distinction between the left, as well as the right parts of E 1 (c) and the harmonic phase difference between the left and right partitions by E 2 (d).
Symmetry 15 00901 g003
Figure 4. Harmonic power spectrum and decomposition spectrum in the x–polarization direction by E 2 . (a) shows, given by Equation (11), the corresponding total harmonic spectrum P (red solid line), the spectrum P 1 (consisting of the dark green solid line in (a)) related to the harmonic phase difference, and the harmonic total spectrum xhhg (solid black line) calculated from the TDSE. (b) shows the spectrum P 2 related to the harmonic phase difference (corresponding to the light green solid line in (b), the decomposed spectrum P 3 related to the amplitude power of the left, as well as the right partitions (consisting of the blue solid line in (b)), and the harmonic total spectrum xhhg (solid black line) calculated from the TDSE.
Figure 4. Harmonic power spectrum and decomposition spectrum in the x–polarization direction by E 2 . (a) shows, given by Equation (11), the corresponding total harmonic spectrum P (red solid line), the spectrum P 1 (consisting of the dark green solid line in (a)) related to the harmonic phase difference, and the harmonic total spectrum xhhg (solid black line) calculated from the TDSE. (b) shows the spectrum P 2 related to the harmonic phase difference (corresponding to the light green solid line in (b), the decomposed spectrum P 3 related to the amplitude power of the left, as well as the right partitions (consisting of the blue solid line in (b)), and the harmonic total spectrum xhhg (solid black line) calculated from the TDSE.
Symmetry 15 00901 g004
Figure 5. Comparison of the wave packet evolution process of He atoms by E 1 (ad) and by E 2 (eg). The white solid lines in (a) and (e) simulate the laser fields in the y–direction corresponding to (ad) and (eh), respectively. The orange in (e) indicates the electrostatic field. The four selected moments are marked by white dots.
Figure 5. Comparison of the wave packet evolution process of He atoms by E 1 (ad) and by E 2 (eg). The white solid lines in (a) and (e) simulate the laser fields in the y–direction corresponding to (ad) and (eh), respectively. The orange in (e) indicates the electrostatic field. The four selected moments are marked by white dots.
Symmetry 15 00901 g005
Figure 6. Harmonic spectrum and waveform of the harmonic electric field synthesized a second light in the x and y directions in the presence of E 4 . (a) presents the harmonic spectrum E 4 x (black solid line) in the x–polarization direction and E 4 y (red solid line) in the y–polarization direction, respectively. (b) presents the attosecond light between 0.5 o.c. and 1 o.c. synthesized with harmonic electric fields in the x and y polarization directions.
Figure 6. Harmonic spectrum and waveform of the harmonic electric field synthesized a second light in the x and y directions in the presence of E 4 . (a) presents the harmonic spectrum E 4 x (black solid line) in the x–polarization direction and E 4 y (red solid line) in the y–polarization direction, respectively. (b) presents the attosecond light between 0.5 o.c. and 1 o.c. synthesized with harmonic electric fields in the x and y polarization directions.
Symmetry 15 00901 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Fu, T.; Guo, F.; Wang, J.; Chen, J.; Yang, Y. Waveform Modulation of High-Order Harmonics Generated from an Atom Irradiated by a Laser Pulse and a Weak Orthogonal Electrostatic Field. Symmetry 2023, 15, 901. https://doi.org/10.3390/sym15040901

AMA Style

Fu T, Guo F, Wang J, Chen J, Yang Y. Waveform Modulation of High-Order Harmonics Generated from an Atom Irradiated by a Laser Pulse and a Weak Orthogonal Electrostatic Field. Symmetry. 2023; 15(4):901. https://doi.org/10.3390/sym15040901

Chicago/Turabian Style

Fu, Tingting, Fuming Guo, Jun Wang, Jigen Chen, and Yujun Yang. 2023. "Waveform Modulation of High-Order Harmonics Generated from an Atom Irradiated by a Laser Pulse and a Weak Orthogonal Electrostatic Field" Symmetry 15, no. 4: 901. https://doi.org/10.3390/sym15040901

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop