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

Origin of unusual even-order harmonic generation by a vortex laser

Open Access Open Access

Abstract

When a spatially-inhomogeneous few-cycle vortex laser interacting with quantum wells, besides the general odd-order harmonics occur, unusual “even-order” ones can be found, which are clarified to possess even orders, however their topological charge numbers are not consistent with those predicted from the vortex transformation criterion (i.e., topological charge number should be directly proportional to its harmonic order for any a harmonic). The origin is the broader spectral width of odd-order harmonics tails at the positions of even-order harmonics due to the short-duration pulse excitation, whose contribution overwhelms the spatial-inhomogeneity degree.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

With the development of few-cycle laser technology, novel investigations have been focused on the so-called extreme nonlinear optics [1,2], in which carrier-wave Rabi flopping [3,4], third-harmonic generation in disguise of second-harmonic generation [5] and on the other novel phenomena are disclosed. Recently, if the above few-cycle excitation laser is simultaneously spatially-inhomogeneity, even-order harmonics occur besides the general odd-order ones due to the broken inversion symmetry [69]. Moreover, the existing results show that the inhomogeneity of the excitation field can obviously extend the cutoff energy in the high-order harmonic generation process [6,7] and introduce an appreciable modification to the energy-resolved photoelectron spectra [8].

Beyond this, if the excitation laser possesses a phase singularity, that is a vortex laser (for example the simplest Laguerre-Gaussian (LG) beam), this phase singularity can transfer to high-order vortex harmonics [1016], THz beams [17,18] or nonlinear precursors [19] via vortex transformation. Due the phase wind imprints an orbital angular momentum (OAM) to the beam [11], the high-order optical vortex beam has more potential applications than low-order ones such as in quantum cryptography [13] and photo-excitation [14]. In addition, based on the criterion that the topological charge number lq of a certain q-order harmonic is directly proportional to its harmonic order q, lq=ql (l being the topological charge number of the fundamental incident LG beam) [1120], the vortex laser excitation finds the application in the origin clarification of unusual spectral components [9,16]. In the following, we consider the excitation laser both with the spatial inhomogeneity and with the phase singularity interacting with quantum wells, a nice target with convenient energy-gap tunability [9]. It is found that there exist unusual ‘even-order’ vortex harmonics, not real even-order ones. The origin of unusual ‘even-order’ harmonics is that the spectral width of odd-order harmonics is so broad that it tails and enhances at the spectral positions of the so-called even-order harmonics, whose descriptions about this origin is different with that in previous works [610].

2. Theory

This incident spatially inhomogeneous LG pulse polarized along x direction and propagated through the square quantum well along z direction is written as [69,19],

$$\vec{E}(t = 0,\;r,\;\phi ,\;z) = {E_{lp}}\,{\mathop{\textrm {sech}}\nolimits} \left[ {\frac{{1.76({z - {z_0}} )}}{{c{\tau_0}}}} \right]\cos \left[ {\frac{{{\omega_0}({z - {z_0}} )}}{c}} \right][{1 + \varepsilon \cdot h(x )} ]{\vec{e}_x}.$$
Here ω0 and τ0 are a central frequency and pulse duration of full width at half maximum (FWHM) of intensity, respectively. h(x) represents the functional form of the inhomogeneous field, which usually approximated as a power series $h(x )= \sum\nolimits_{i = 1}^N {{b_i}{x^i}} ,$ with the coefficients bi obtained by fitting the results from a finite-element simulation. Generally speaking, sometimes only the linear order term of h(x) retained, i.e., h(x) =x is applicable [610]. ɛ is called the inhomogeneous factor characterizing the spatial-inhomogeneity degree of the incident field. The field amplitude Elp is defined as,
$$\begin{aligned} {E_{lp}}(t = 0,\;r,\;\phi ,\;z) & = \frac{{{E_0}}}{{{{({1 + {{\widetilde z}^2}/z_R^2} )}^{1/2}}}}{\left( {\frac{r}{{a({\widetilde z} )}}} \right)^{|l |}}L_p^{|l |}\left( {\frac{{2{r^2}}}{{{a^2}({\widetilde z} )}}} \right)\exp \left( {\frac{{ - {r^2}}}{{{a^2}({\widetilde z} )}}} \right)\\ & \times \exp \left( {\frac{{ - ik{r^2}\widetilde z}}{{2({{{\widetilde z}^2} + z_R^2} )}}} \right)\exp ({ - il\phi } )\exp \left( { - i({2p + |l |+ 1} ){{\tan }^{ - 1}}\frac{{\widetilde z}}{{{z_R}}}} \right), \end{aligned}$$
with ${\widetilde z}$=z-z0, ZR the Rayleigh range, $a({\widetilde z} )$ the beam radius and a0=a(0) the beam waist at z0, $L_p^{|l |}$ the associated Laguerre polynomial. Here the most important factor charactering the laser vortex property is exp(-ilϕ) with l (l = 0, ±1, ±2,…) being the topological charge (TC) number, ϕ being the azimuthal angle, and indicating the helical phase of vortex laser. In addition, the p denotes the transverse radial node number [1316]. E0 and k0/c are the amplitude of the electric field and carrier wave number, respectively.

The Maxwell-Bloch (M-B) model describing the whole laser-matter interaction takes the form [10,16,19]

$$\frac{{\partial \vec{H}}}{{\partial t}} ={-} \frac{1}{{{\mu _0}}}\nabla \times \vec{E},\quad \frac{{\partial \vec{E}}}{{\partial t}} = \frac{1}{{{\varsigma _0}}}\nabla \times \vec{H} - \frac{1}{{{\varsigma _0}}}\frac{{\partial \vec{P}}}{{\partial t}},$$
$$\frac{{\partial {\rho _{12}}}}{{\partial t}} ={-} i\left( {{\omega_{12}}{\rho_{12}} + \frac{{d{E_x}}}{\hbar }n} \right) - \frac{1}{{{\tau _1}}}{\rho _{12}},\quad \frac{{\partial n}}{{\partial t}} = i\frac{{2d}}{\hbar }{E_x}({\rho _{12}} - \rho _{12}^\ast ) - \frac{1}{{{\tau _2}}}n.$$
Equation (3) is Maxwell equations used for describing laser propagation in media with $\vec{E}$, $\vec{H}$ and $\vec{P}$ being are the electric, magnetic vectors and macroscopic polarization, respectively. ς0 and μ0 are vacuum permittivity and permeability, respectively. In contrast, Eq. (4) is Bloch equations used for describing the medium response to the induced laser, with ρ12 being the complex microscopic polarization and n=ρ22-ρ11 indicating the particle population difference at two intersubband energy levels with an energy difference of ħω12 [21,22], which is related to the quantum well width, and the corresponding dipole transition matrix moment d [9]. The transverse dephasing time τ1 and the longitudinal excited-state lifetime τ2 are generally in the picosecond time scale [4,19] and can be safely ignored in the simulation due to the short-time laser excitation. The choice of z0 ensures the excitation laser e penetrate negligibly into the medium at t = 0 [21]. The M-B equations are solved by employing Yee's finite-difference time-domain (FDTD) discretization [23] scheme combined with the predictor-corrector or Runge-Kutta algorithm [2427]. In the following, the laser-matter interaction parameters with those are adopted [9,10,22]: ω0=6.0×1014 s−1 (∼3.14 µm), τ0=18 fs (∼1.72 laser period), τ1=1.0 ps, τ2=0.5 ps, z0=20 µm, a0=12 µm, N = 6.0×1016 cm−3 and E0=3.57×108 V/m.

3. Results and discussion

The interaction between a few-cycle LG10 (i.e., l = 1 and p = 0) beam and symmetric quantum wells is numerically simulated, in which the well-width is 62 Å (ω12=1.11ω0) [9,10]. The laser spectra under different spatial-inhomogeneity factors ɛ are first investigated. Without inhomogeneity ɛ=0 (solid line in Fig.  1(a)), there exist only the first and the third order harmonics due to the weak and short pulse duration. These two adjacent harmonics with an exponentially decaying intensity change and little increasing peak width, clearly indicating the characteristics of the low-order perturbative theory. When the spatial inhomogeneity of the excitation field is introduced (ɛ=0.006 for example), the additional spectral components appear between the adjacent odd-order harmonics (dashed line in Fig.  1(a)). When the inhomogeneity is further increased (ɛ=0.012), these spectral components grows into clear peaks positioning exactly at even multiple of the fundamental frequency (dash-dotted line in Fig.  1(a)). They can thus call as even-order harmonics. However, if we turn to check their corresponding transverse field distributions, the TC number for each harmonics is not consistent with the prediction from the vortex transformation criterion [1016] that the topological charge number of each harmonic is directly proportional to its harmonic order, as shown in Fig.  2.

 figure: Fig. 1.

Fig. 1. Laser spectra under different (a) inhomogeneity degree, (b) propagation distances, and (c) pulse durations with τ0=18 fs, ɛ=0.012, z = 50 µm and well width 62 Å (ω12=1.11ω0). (d) Same with (a), but for well width of 60 Å (ω12=1.18ω0).

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Transverse field distributions of the (a) first, (b) “second”, (c) third, (d) “fourth”, (e) fifth, and (f) “sixth” order harmonics in the x-y plane, which are obtained by using a spectral filter with a width of 0.2ω0. ɛ=0.012 and the other parameters are same with those in Fig.  1(a).

Download Full Size | PDF

From Fig.  2, one can find that for the first, third, and fifth -order harmonics, the transverse field distributions in the x-y plane perpendicular to the propagation direction indicate the TC number is the value as expected. For example, Fig.  2(a) clearly shows one positive and one negative beads, whose total shape looks like a Chinese Tai Chi diagram, indicating its TC number is one, which is consistent with its harmonic order 1. As for that for the third-order harmonic (see Fig.  2(c)), there exist three positive and three negative beads, indicating the TC number is 3, same with its corresponding harmonic order 3. The same point can be as well as for that in Fig.  2(e), where TC number is 5, equal to the harmonic order 5. Thus the TC number of any odd-order harmonic is directly proportional to its harmonic order, i.e., lq=ql (here l = 1). However, if turning to the even-order ones (see Figs.  2(b), 2(d), and 2(f)), the TC numbers indicated by them are not equal to the corresponding harmonic order values. That shown in Fig.  2(b) exhibits 1 positive and 1 negative peaks and the TC number is 1, which is same with the first-order harmonic, but the harmonic number is 2. It seems that the first-order harmonic shows some contribution to the so-called second-order harmonic. The similar things happened to the other even-order harmonics (Figs.  2(d) and (2f)). This phenomenon seems propose a controversy, and why it occurs and how to solve it?

Natural focus will be on the spatial-inhomogeneity degree of the excitation field and propagation effects and maybe one wonder whether they make the above phenomenon occur. The transverse field distributions (see Fig.  3) for a smaller inhomogeneity factor show the similar pattern with those in Fig.  2, except weaker intensity, as well as those after a new propagation distance (see Fig.  4). Both inhomogeneity degree and propagation show certain influence on the spectral characteristics, but the above controversy is still retained. In addition, when the well-width changes from 62 Å to 60 Å (ω12=1.11ω0 to 1.18ω0), the above phenomenon of transverse field distributions for a certain order harmonic can also be found as shown in Fig.  1(d) and Fig.  5.

 figure: Fig. 3.

Fig. 3. Same as in Fig.  2, but for ɛ=0.006.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Same as in Fig.  2, but for z = 32 µm.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. Same as in Fig.  2, but for well-width 60 Å.

Download Full Size | PDF

However, thinking that the first-order harmonic shows the same TC number with the second-order harmonic (see Figs.  2(a) and 2(b)), the former seems to give some contribution to the latter. If we turn to changing the pulse duration of the excitation field, for example doubling from 18 fs from 38 fs, as seen from the spectrum in Fig.  1(c), the above unusual even-order harmonics significantly weaken, indicating the contribution from the large spectral width of a short excitation laser. For an even longer laser pulse, the corresponding calculation (not shown here) confirms that the spatial-inhomogeneity introduced even-order harmonics are normal with the TC numbers directly proportional to its harmonic order.

4. Conclusion

In summary, the spatially inhomogeneous few-cycle vortex beam propagating in quantum wells has been investigated with the introduce of the spatial-inhomogeneity of the excitation field, the even-order harmonics should occur beyond the conventional odd-order ones. However, the transverse field distributions of these even-order vortex harmonics presents that the corresponding TC numbers are not those predicted from the vortex transformation criterion that the topological charge number of each harmonic is directly proportional to its harmonic order. This controversy has been solved step by step via the investigations of the contributions from inhomogeneity and pulse bandwidth. If the contribution from large pulse bandwidth is dominant, the controversy is emerging. In contrast, if that from spatial-inhomogeneity is dominant, the transverse field distribution for each even-order vortex harmonics would show normal TC number. In addition, propagation effects and spatial-inhomogeneity both show certain influence on these unusual harmonics.

Funding

National Natural Science Foundation of China (Grants Nos. 61775087,11674312).

Acknowledgments

C.J.Z. gratefully acknowledges the support of open fund of the state key laboratory of high field laser physics of SIOM and the Key Program in the Youth Elite Support Plan in Universities of Anhui Province (gxyqZD2018039).

Disclosures

The authors declare no conflicts of interest.

References

1. R. Zuo, X. Song, X. Liu, S. Yang, and W. Yang, “The influence of intraband motion on the interband excitation and high harmonic generation,” Chin. Phys. B 28(9), 094208 (2019). [CrossRef]  

2. M. Wegener, Extreme Nonlinear Optics (Springer-Verlag, Heidelberg, 2005).

3. S. Hughes, “Breakdown of the area theorem: carrier-wave Rabi flopping of femtosecond optical pulses,” Phys. Rev. Lett. 81(16), 3363–3366 (1998). [CrossRef]  

4. O. D. Mücke, T. Tritschler, M. Wegener, U. Morgner, and F. X. Kärtner, “Signatures of carrier-wave Rabi flopping in GaAs,” Phys. Rev. Lett. 87(5), 057401 (2001). [CrossRef]  

5. T. Tritschler, O. D. Mücke, M. Wegener, U. Morgner, and F. X. Kärtner, “Evidence for Third-Harmonic Generation in Disguise of Second-Harmonic Generation in Extreme Nonlinear Optics,” Phys. Rev. Lett. 90(21), 217404 (2003). [CrossRef]  

6. M. F. Ciappina, J. Biegert, R. Quidant, and M. Lewenstein, “High-order-harmonic generation from inhomogeneous fields,” Phys. Rev. A 85(3), 033828 (2012). [CrossRef]  

7. I. Yavuz, E. A. Bleda, Z. Altun, and T. Topcu, “Generation of a broadband xuv continuum in high-order-harmonic generation by spatially inhomogeneous fields,” Phys. Rev. A 85(1), 013416 (2012). [CrossRef]  

8. M. F. Ciappina, J. A. Pérez-Hernández, T. Shaaran, J. Biegert, R. Quidant, and M. Lewenstein, “Above-threshold ionization by few-cycle spatially inhomogeneous fields,” Phys. Rev. A 86(2), 023413 (2012). [CrossRef]  

9. C. Zhang, C. Liu, and Z. Xu, “Control of higher spectral components by spatially inhomogeneous fields in quantum wells,” Phys. Rev. A 88(3), 035805 (2013). [CrossRef]  

10. C. Zhang and C. Liu, “Carrier-envelope-phase dependence of harmonics induced by a few-cycle vortex laser,” Laser Phys. Lett. 17(12), 125401 (2019). [CrossRef]  

11. X. Zhang, B. Shen, Y. Shi, X. Wang, L. Zhang, W. Wang, J. Xu, L. Yi, and Z. Xu, “Generation of intense high-order vortex harmonics,” Phys. Rev. Lett. 114(17), 173901 (2015). [CrossRef]  

12. A. Mair, A. Vaziri, G. Weihs, and A. Zeilinger, “Entanglement of the orbital angular momentum states of photons,” Nature 412(6844), 313–316 (2001). [CrossRef]  

13. M. Zürch, C. Kern, P. Hansinger, A. Dreischuh, and C. Spielmann, “Strong-field physics with singular light beams,” Nat. Phys. 8(10), 743–746 (2012). [CrossRef]  

14. A. Picon, A. Benseny, J. Mompart, J. R. V. de Aldana, L. Plaja, G. F. Calvo, and L. Roso, “Transferring orbital and spin angular momenta of light to atoms,” New J. Phys. 12(8), 083053 (2010). [CrossRef]  

15. L. Rego, J. San Román, A. Picón, L. Plaja, and C. Hernández-García, “Nonperturbative twist in the generation of extreme-ultraviolet vortex beams,” Phys. Rev. Lett. 117(16), 163202 (2016). [CrossRef]  

16. C. Zhang, E. Wu, M. Gu, Z. Hu, and C. Liu, “Characterization method of unusual second-order-harmonic generation based on vortex transformation,” Phys. Rev. A 96(3), 033854 (2017). [CrossRef]  

17. H. Wang, Y. Bai, E. Wu, Z. Wang, P. Liu, and C. Liu, “THz necklace beam generated from two-color vortex laser-induced air-plasma,” Phys. Rev. A 98(1), 013857 (2018). [CrossRef]  

18. H. Wang, Q. Song, S. Zheng, Q. Lin, E. Wu, Y. Ai, C. Liu, and S. Xu, “Terahertz-mid-infrared anisotropic vortex beams generation via few-cycle vortex-laser-induced air plasma,” J. Opt. 21(9), 095501 (2019). [CrossRef]  

19. Y. Chen, X. Feng, and C. Liu, “Generation of nonlinear vortex precursors,” Phys. Rev. Lett. 117(2), 023901 (2016). [CrossRef]  

20. G. Gariepy, J. Leach, K. Taec Kim, T. J. Hammond, E. Frumker, R. W. Boyd, and P. B. Corkum, “Creating high-harmonic beams with controlled orbital angular momentum,” Phys. Rev. Lett. 113(15), 153901 (2014). [CrossRef]  

21. V. P. Kalosha and J. Herrmann, “Formation of optical subcycle pulses and full Maxwell-Bloch solitary waves by coherent propagation effects,” Phys. Rev. Lett. 83(3), 544–547 (1999). [CrossRef]  

22. C. Zhang, E. Wu, M. Gu, and C. Liu, “Propagation effects in the generation process of high-order vortex harmonics,” Opt. Express 25(18), 21241–21246 (2017). [CrossRef]  

23. K. Yee, “Numerical solution of initial boundary value problems involving Maxwell’s equations in isotropic media,” IEEE Trans. Antennas Propag. 14(3), 302–307 (1966). [CrossRef]  

24. X. Song, M. Wu, Z. Sheng, D. Dong, H. Wu, and W. Yang, “Carrier-envelope phase dependence of a half-cycle soliton generation in asymmetric media,” Laser Phys. Lett. 11(5), 056002 (2014). [CrossRef]  

25. R. W. Ziolkowski, J. M. Arnold, and D. M. Gogny, “Ultrafast pulse interactions with two-level atoms,” Phys. Rev. A 52(4), 3082–3094 (1995). [CrossRef]  

26. C. Zhang, W. Lu, E. Wu, and C. Liu, “Influence of frequency detuning on carrier-wave Rabi flop-related phenomena in the extreme nonlinear optics regime,” Eur. Phys. J. D 72(11), 194 (2018). [CrossRef]  

27. X. Song, G. Shi, G. Zhang, J. Xu, C. Lin, J. Chen, and W. Yang, “Attosecond time delay of retrapped resonant ionization,” Phys. Rev. Lett. 121(10), 103201 (2018). [CrossRef]  

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 (5)

Fig. 1.
Fig. 1. Laser spectra under different (a) inhomogeneity degree, (b) propagation distances, and (c) pulse durations with τ0=18 fs, ɛ=0.012, z = 50 µm and well width 62 Å (ω12=1.11ω0). (d) Same with (a), but for well width of 60 Å (ω12=1.18ω0).
Fig. 2.
Fig. 2. Transverse field distributions of the (a) first, (b) “second”, (c) third, (d) “fourth”, (e) fifth, and (f) “sixth” order harmonics in the x-y plane, which are obtained by using a spectral filter with a width of 0.2ω0. ɛ=0.012 and the other parameters are same with those in Fig.  1(a).
Fig. 3.
Fig. 3. Same as in Fig.  2, but for ɛ=0.006.
Fig. 4.
Fig. 4. Same as in Fig.  2, but for z = 32 µm.
Fig. 5.
Fig. 5. Same as in Fig.  2, but for well-width 60 Å.

Equations (4)

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

E ( t = 0 , r , ϕ , z ) = E l p sech [ 1.76 ( z z 0 ) c τ 0 ] cos [ ω 0 ( z z 0 ) c ] [ 1 + ε h ( x ) ] e x .
E l p ( t = 0 , r , ϕ , z ) = E 0 ( 1 + z ~ 2 / z R 2 ) 1 / 2 ( r a ( z ~ ) ) | l | L p | l | ( 2 r 2 a 2 ( z ~ ) ) exp ( r 2 a 2 ( z ~ ) ) × exp ( i k r 2 z ~ 2 ( z ~ 2 + z R 2 ) ) exp ( i l ϕ ) exp ( i ( 2 p + | l | + 1 ) tan 1 z ~ z R ) ,
H t = 1 μ 0 × E , E t = 1 ς 0 × H 1 ς 0 P t ,
ρ 12 t = i ( ω 12 ρ 12 + d E x n ) 1 τ 1 ρ 12 , n t = i 2 d E x ( ρ 12 ρ 12 ) 1 τ 2 n .
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.