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

Vortex harmonic generation in indium tin oxide thin film irradiated by a two-color field

Open Access Open Access

Abstract

When a two-color Laguerre-Gaussian laser beam propagates through an indium tin oxide (ITO) material, the spatial distributions of odd- and even-order vortex harmonics carrying orbital angular momentum (OAM) are studied. The origin of vortex harmonics can be directly clarified by investigating their dependence on the incident laser field amplitude and frequency. In addition, it is shown that the spectral intensities of vortex harmonics are sensitive to the epsilon-near-zero nonlinear enhancing effects and the thickness of ITO materials. Thus the vortex harmonics can be conveniently tunable, which provides a wider potential application in optical communications based on high-order OAM coherent vortex beams.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Due to the potential applications of high-order harmonic generation in molecular tomography and production of ultrafast coherent EUV light sources, it has attracted the researchers’ wide attention [14]. The high-order harmonic generation relies on extreme nonlinear optics where the still very weak nonlinear effects requiring a longer laser-matter interaction length and/or a higher excitation laser intensity, which is usually impossible for a laser beam directly from oscillator without multi-pass amplification [14]. The occurrence of epsilon-near-zero (ENZ) materials provides an alternative scheme to make the above route become possible [5,6], because the internal field within the front surface and the third-order nonlinearity coefficients of this kind of material is significantly enhanced [7,8]. For example, a ninth-order harmonic is generated when cadmium oxide (CdO), a kind of indium tin oxide (ITO) ENZ material, driven by an around 10 GW/cm2 laser [5]. Moreover, an obvious harmonic redshift phenomenon is disclosed in ENZ materials under a few-cycle laser excitation [6]. It is pointed out that the three-step model of high-order harmonic generations in atom, the electrons should be three process: ionization, acceleration and recombination [3,9,10]. In contrast, the origin of high-order vortex harmonics in the work is results from the macroscopic coherent polarization Px(t) acting as a source of the re-emitted field [11,12].

Though the odd-order harmonics can be obtained in the above works [5,6], the investigation is limited to the single-color laser excitation case. Here we investigate the harmonic generation in the ENZ materials but under two-color laser excitation. The two color field is composed of the fundamental laser field and its double-frequency field. Via the two-color excitation, even-order harmonics are also generated in any materials without the necessity of breaking the centro-symmetry of materials based on polar molecules [11], surface effects [12,13] or nonlocal bulk effects [14,15]. The synthesized two-color field via a fundamental laser and its double-frequency field for high-order harmonic generation has been widely adopted and many numerical simulation and experimental investigation confirm its feasibility [1628]. Up to now, there exist three kinds of resources for the occurrence of even-order harmonics: a) breaking of symmetry in the field of the fundamental and its second harmonic [1620], b) modification in field coherence [2124], and c) contribution from many different nonlinear processes such as four-wave mixing and sum-frequency/difference frequency [2528]. It is meaningful to clearly explore the real generation origin of even-order harmonics driven by a two-color field.

Inspired by the above question, one can use a two-color Laguerre-Gaussian (LG) vortex beam to interact with ITO materials under the case of very weak incident laser intensity (∼1010 W/cm2). The topological charge numbers of odd- and even-order vortex harmonics with the spatial distributions can be directly distinguished by their transverse field distributions [2933] and with the help of time-frequency analysis. At the same time, the information carried by OAM in high-order vortex harmonics may be potential application in optoelectronics and optical information processing. Importantly, the spectral intensities of even-order harmonics depend sensitively on the incident field amplitude and frequency. In addition, an optimum thickness of ITO materials is determined to support the strongest harmonic signal. For example, the fourth-order vortex harmonic, the energy conversion efficiency is up to 2.3 × 10−4.

2. Theoretical model

A synthesized field via a two-color LG vortex laser, polarized along the x direction, first freely propagates in the vacuum and then is incident on the ITO material with thickness of d = 900 nm along the z direction. When the laser arrives at the front surface (zs = 60 µm), it partially penetrates the ITO material and continue to propagate until again into the vacuum. The widely adopted formula for a LG vortex laser with amplitude Elp is written as [6,2931,34]

$$\begin{aligned} {E_{lp}}(t = 0,r,\phi ,z) &= \frac{{{E_0}}}{{{{({1 + z{^{\prime}{^2}}/z_R{^2}} )}^{1/2}}}}{\left( {\frac{r}{{a({z^{\prime}} )}}} \right)^{|l |}}L_p^{|l |}\left( {\frac{{2{r^2}}}{{{a^2}({z^{\prime}} )}}} \right)\textrm{exp} \left( {\frac{{ - {r^2}}}{{{a^2}({z^{\prime}} )}}} \right)\\ &\times \textrm{exp} \left( {\frac{{ - ik{r^2}\widetilde z}}{{2({z{^{\prime}{^2}} + z_R^2} )}}} \right)\textrm{exp} ({ - il\phi } )\textrm{exp} \left( { - i({2p + |l |+ 1} ){{\tan }^{ - 1}}\frac{{z^{\prime}}}{{{z_R}}}} \right). \end{aligned}$$

ZRa02/λ is the Rayleigh length, a(z’) is the radius of beam waist at position z'=z-z0 and a(0) =a0 at z0. E0 is the electric field amplitude. z0 is the initial laser peak position and a suitable choice of z0 should ensure that the LG pulse penetrates negligibly into the media at t = 0 [11,13,29], and here z0 is set as 30 µm. Llp is the associated Laguerre polynomial with p denoting the transverse radial node number (here p = 0 just for convenient simulation). $r = \sqrt {{{({x - {x_0}} )}^2} + {{({y - {y_0}} )}^2}} $ (x0 and y0 are the axial center coordinates on the transverse x-y plane). exp(-i) describes the helical phase profile of an optical vortex, with l (0, ± 1, ± 2,…) the topological charge number and ϕ the azimuthal angle.

Because in our following numerical simulation, the Rayleigh length is ZRa02/λ∼552 µm with λ=1.28 µm and a0 = 15 µm, which is much larger than the propagation distance investigate here, i.e., ZR >> zs is satisfied, a paraxial plane-wave description of the above electric field (1) is in fact enough for our purpose. In our numerical simulation, the incident field is synthesized via a fundamental laser of LG10 (l = 1) beam and its double-frequency laser of LG20 (l = 2) beam. This incident field is assumed to have a sech-shaped envelope [6,29] and simplified as,

$$\vec{E}(t = 0,z) = {E_{10}}\textrm {sech} \left[ {\frac{{1.76z^{\prime}}}{{c{\tau_0}}}} \right]\cos \left( {{\omega_0}\frac{{z^{\prime}}}{c}} \right){\vec{e}_x} + A \cdot {E_{20}}\textrm {sech} \left[ {\frac{{1.76z^{\prime}}}{{c{\tau_0}}}} \right]\cos \left( {{\omega_2}\frac{{z^{\prime}}}{c}} \right){\vec{e}_x}$$

Here ω2 = 2ω0, and τ0 denotes the full width at half maximum (FWHM) of pulse intensity. ω0 denote the central frequency. A is the field amplitude ratio of the two components of the incident two-color field. c is the light speed at vacuum.

The interaction between this vortex laser and the ITO material can be described by combining three-dimensional Maxwell equations [6,7,35] with the paradigmatic-Kerr equation [6,35],

$$\nabla \times \overrightarrow E ={-} {\mu _0}\frac{{\partial \overrightarrow H }}{{\partial t}},\nabla \times \overrightarrow H = {\varepsilon _0}{\varepsilon _\infty }\frac{{\partial \overrightarrow E }}{{\partial t}} + \frac{{\partial \overrightarrow P }}{{\partial t}},$$
$$\frac{{{\partial ^2}\overrightarrow P }}{{\partial {t^2}}} + {\delta _e}{\omega _e}\frac{{\partial \overrightarrow P }}{{\partial t}} + \omega _e^2{\left( {1 + \frac{{{{|{\overrightarrow P } |}^2}}}{{P_s^2}}} \right)^{ - {3 / 2}}}\overrightarrow P = {\varepsilon _0}({\varepsilon _s} - {\varepsilon _\infty })\omega _e^2\overrightarrow E .$$

Ps, ωe, δe, ɛs, and ɛ are the saturation polarization, resonant frequency, loss coefficient, static dielectric permittivity, and high frequency permittivity, respectively. Here ωe = 1.327 × 1015 s-1, δe = 0.01371, ɛs = 3.846698, and ɛ=3.1178 are adopted [8,3540]. In addition, in the ENZ material investigation, the important factor is the so-called ENZ frequency. It can be obtained by setting the real part of dielectric permittivity as zero, which is corresponding to the zero-crossing-point in the dispersion relation curve [7,35],

$${\omega _0} = \frac{{{\omega _e}}}{{\sqrt {2{\varepsilon _\infty }} }}{\{{({{\varepsilon_s} + {\varepsilon_\infty }({1 - \delta_e^2} )} )+ {{[{{{({{\varepsilon_s} + {\varepsilon_\infty }({1 - \delta_e^2} )} )}^2} - 4{\varepsilon_s}{\varepsilon_\infty }} ]}^{{1 / 2}}}} \}^{{1 / 2}}}.$$

Here based on the above values of ωe, δe, ɛs, and ɛ, the ENZ frequency ω0 = 1.1108ωe, corresponding to a wavelength of λ=1.28 µm [41]. The above Maxwell equations (3) are first solved by employing Yee's finite-difference time-domain (FDTD) discretization scheme [4244] and the paradigmatic-Kerr Eq. (4) by the Runge-Kutta algorithm [45,46], respectively. Then, the transmitted electric field E(t, z=∞) after the ITO material is recorded. Finally, a Fourier transformation is conducted to the transmitted field for investigating the spectral characteristics of harmonics. The laser amplitude is E0 = 5 × 108 V/m (corresponding to laser intensity I0 = 3.33 × 1010 W/cm2), and a0 = 15 µm, A = 0.1, τ0 = 20 fs are used for the following numerical simulation.

3. Results and discussion

First, the temporal and spatial evolution of a single-color LG pulse with frequency equal to ω0 propagating through the ITO material is numerically investigated. The field waveform is recorded at a position of z = 65 µm, as shown in Fig. 1(a). the laser pulse after interaction with the ITO materials has a significantly asymmetric time-spatial domain profile, which is consistent with previous work [5,6]. It is pointed out that the laser duration (τ0 = 20 fs) is much larger than the optical cycle of incident laser pulse (∼ 4.27 fs), the redshift in high-order harmonic generation should disappear [6]. Due to the waveform change and inversion symmetry of the system, only the odd-order harmonics are found as shown in Fig. 2(a) (solid line). When another weaker single-color laser field (2ω0) whose intensity is only one percent of the former ω0 field is incident, the change in the time-spatial domain profile is trivial after interacting with the ITO material, and the corresponding harmonics cannot be found as shown in Fig. 2(a) (dashed line). But when these two fields are superposed to form a two-color beam and then as a whole interacting with the ITO material, as shown in Fig. 1(b), the rise and falling parts of the temporal waveform changes significantly if compared with the case in Fig. 1(a), where there exists obvious peak splitting in the electric field waveform. The results show that not only the odd-order harmonic generation is found, but also the even-order harmonic generation can be obtained (see dash-dotted line in Fig. 2(a)). It should be pointed out that the intensities of odd-order harmonics are nearly unchanged due to the adding of the weaker 2ω0 beam. This phenomenon can be further proved by the subsequent time frequency analyses to these vortex harmonics, as shown in Fig. 3. As shown in Fig. 3(a), the only odd-order harmonics can be found when the only single-color field (ω0) is present. In contrast, the odd- and even-order harmonics simultaneously occur under the two-color excitation as shown in Fig. 3(b).

 figure: Fig. 1.

Fig. 1. Temporal and spatial evolution of a LG pulse propagating through the ITO material under (a) single-color ω0 or (b) two-color ω0 + 2ω0 excitation with fundamental laser frequency equal to ω0.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Spectral distributions of vortex harmonics (a) at different incident LG pulses and (b) at ω0 + 0.3ω0. The 4ω0 harmonic signal intensity versus the incident laser peak amplitude of (c) ω0 laser for fixing 2ω0 field intensity and (d) 2ω0 field or fixing ω0 field intensity.

Download Full Size | PDF

 figure: Fig. 3.

Fig. 3. Time-frequency analysis to high-order vortex harmonics in Fig. 2(a) under (a) single-color ω0 and (b) two-color ω0 + 2ω0 excitation.

Download Full Size | PDF

Second, the origin of even-order harmonics maybe comes from the four-wave mixing effects [2528]. Here we take the fourth-order harmonic as an example to present a further confirmation in the case of harmonic generation from a two-color field interacting with the ITO material. First, we artificially change the original double-frequency ω2 = 2ω0 to a non-integer lower frequency ω2 = 0.3ω0 and make the same simulation procedure to investigate the harmonic distribution, as shown in Fig. 2(b). Now, beside the other peaks in high-order vortex harmonics, the fourth-order harmonic peaks at 2.3ω0 instead of the previous 4ω0. Such phenomenon keeps if we try other different frequency lasers beyond 0.3ω0 (not shown here). Second, corroborative numerical simulation and fitting is shown in Figs. 2(c) and (d), which is based on adjusting the incident laser peak amplitude of ω0 field or 2ω0 field, respectively. As shown in Fig. 2(c), fixing the 2ω0 field peak amplitude and just tuning the ω0 field peak amplitude, when the ω0 field intensity is not two high, it conforms to the law of absorbing two ω0 photons: the fourth-order harmonic signal is proportional to the square of the ω0 field intensity indicating a fourth-order harmonic photon comes from the contribution two ω0 photons because E4thE12E2. Similarly, if we fixing the ω0 field intensity and tuning the 2ω0 field peak intensity, as shown in Fig. 2(d), we find that the harmonic signal is proportional to the 2ω0 laser intensity indicating that the fourth-order harmonic photon forms by absorbing a 2ω0 photon except for two ω0 photons. Above all, the four-wave mixing mechanism for fourth-order harmonic is confirmed. Importantly, as shown in Fig. 2(c), when the incident ω0 field intensity is two high, the dependence of harmonic signal on the laser intensity deviates from the above quadratic pattern. As the laser intensity increases the signal increases to a maximum and then decreases. The reason is the nonlinear saturation absorption effect and the non-negligible absorption loss existing in the ENZ materials.

Third, in order to further assist the investigation of high-order vortex harmonics, we turn to the transverse field distributions. By utilizing the time-frequency analysis based on the wavelet-transformation technique exemplified by the widely adopted Gabor transformation [6,13,29], the whole vortex harmonic spectrum in Fig. 2(a) is decomposed by filtering each harmonic with a spectral window of 0.2ω0 width around its peak position to conveniently investigate the transverse field distributions of each harmonic. According to the criteria [2933] of vortex harmonics: the topological charge number lq of the qth-order harmonic is directly proportional to its harmonic order, i.e., lq = ql (l = 1 for a fundamental incident ω0 field of LG10), one can directly distinguish certain harmonic order from the transverse field distribution where the topological charger number is easily determined. The transverse field spatial distributions for the different order vortex harmonics are shown in Fig. 4. One can clearly distinguish the fourth-order and sixth-order harmonics from Figs. 4(c) and 4(e). It maybe in turn helpful to proving the above explanation about the generation origin of even-order harmonics.

 figure: Fig. 4.

Fig. 4. Transverse field distributions of (a) second, (b) third, (c) fourth, (d) fifth, (e) sixth and (f) seventh order harmonics under two-color excitation with fundamental laser frequency equal to ω0.

Download Full Size | PDF

Finally, to highlight the nonlinear enhancement effects of ITO materials at ENZ frequency ω0 = 1.1108ωe, we try other laser frequencies of 1.3ωe and 1.5ωe for a comparison investigation. As shown in Fig. 5(a), the intensities of vortex harmonics are obviously reduced when the incident ω0 laser frequency in the two-color field deviates a little more from the ENZ frequency of 1.1108ωe. For example, the peak of the fourth-order harmonic signal for laser frequency at 1.1108ωe is about 5 times that at 1.5ωe. In contrast, the fifth-order harmonic signal is even reduced by 30 times. Correspondingly, the intensities of the transverse field spatial distributions about the fourth-order and fifth-order harmonics will decrease simultaneously (not shown here). The mechanism for high-order vortex harmonic enhancement is attributed to the ENZ enhancement of nonlinearity coefficients [7,8]. Thus, the spectral intensities and spatial distributions of high-order vortex harmonics can be controlled effectively by the incident laser frequency.

 figure: Fig. 5.

Fig. 5. (a) Spectral distributions of high-order harmonic signals at different laser frequencies ω0. Intensities of (b) fourth-order harmonic and (c) sixth-order harmonic at different material thickness d.

Download Full Size | PDF

In addition, the ITO material thickness d has an important influence on the intensities of vortex harmonics. Let's take the fourth-order harmonic as an example. As shown in Fig. 5(b), with the thickness increase from 600 nm to 900 nm, the harmonic signal is first enhanced as the consequence of the phase-matching-free coherent enhancement effect. But with the further increase of material thickness d to 1200 nm, the intensity of harmonic signal decreases. This is due to the accumulated absorption loss becoming stronger and stronger for a long propagation distance, predicted from the Beer–Lambert law. Similarly, as for the sixth-order harmonic, with the increase of propagation distance from 300 nm to 500 nm, the signal is also enhanced and further increase the harmonic signal in turn decreases as shown in Fig. 5(c).

From the perspective of harmonic application in optical communications based on high-order OAM coherent vortex beams, the energy conversion efficiency of a high-order harmonic signal from the excitation laser is essential. The conversion efficiency can be estimated by using the ratio between the peak intensity of each harmonic and the incident laser intensity [40,41,47]. At this optimal ITO material thickness d∼900 nm (ω0 = 1.1108ωe) as shown in Fig. 5(c), the energy conversion efficiency can be numerically estimated as ∼2.3 × 10−4. It is worth noting that, such an enhancement phenomenon of high-order vortex harmonics is not limited to the ITO materials shown here. In fact, the research results obtained in this work are quite general irrespective of ENZ media, therefore the predicted harmonic behaviors must be realized in other ENZ materials.

4. Conclusion

In conclusion, we have studied the spatial distributions of high-order vortex harmonics driven by a two-color LG vortex beam in ITO materials. The results show that the generation origin of odd- and even-order harmonics can be clarified by transverse field distributions and time-frequency analyses. Interestingly, the origin of vortex harmonics can be directly clarified by investigating their dependence on the incident laser field amplitude and frequency. In addition, it is shown that the spectral intensities of vortex harmonics are sensitive to the epsilon-near-zero nonlinear enhancing effects and the thickness of ITO materials. Thus the vortex harmonics can be conveniently tunable, which provides a wider potential application in optical communications based on high-order OAM coherent vortex beams.

Funding

National Natural Science Foundation of China (12074398, 12174161).

Acknowledgments

The authors are grateful to Dr. Yi Wu for inspired discussions.

Disclosures

The authors declare that there are no conflicts of interest related to this article.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. M. Zhu, J. Tong, X. Liu, et al., “Tunnelling of electrons via the neighboring atom,” Light: Sci. Appl. 13(1), 18 (2024). [CrossRef]  

2. M. Lewenstein, P. Balcou, M. Y. Ivanov, et al., “Theory of high-harmonic generation by low-frequency laser fields,” Phys. Rev. A 49(3), 2117–2132 (1994). [CrossRef]  

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

4. C. Zhang and C. Liu, “Ideal laser waveform construction for the generation of super-bright attosecond pulses,” New J. Phys. 17(2), 023026 (2015). [CrossRef]  

5. Y. Yang, J. Lu, A. Manjavacas, et al., “High-harmonic generation from an epsilon-near-zero material,” Nat. Phys. 15(10), 1022–1026 (2019). [CrossRef]  

6. C. Zhang, Y. Zhang, H. Du, et al., “Enhancement and redshift of vortex harmonic radiation in epsilon-near-zero materials,” Opt. Express 31(5), 7725–7733 (2023). [CrossRef]  

7. A. Ciattoni, C. Rizza, A. Marini, et al., “Enhanced nonlinear effects in pulse propagation through epsilon-near-zero media,” Laser Photonics Rev. 10(3), 517–525 (2016). [CrossRef]  

8. M. Z. Alam, I. D. Leon, and R. W. Boyd, “Large optical nonlinearity of indium tin oxide in its epsilon-near-zero region,” Science 352(6287), 795–797 (2016). [CrossRef]  

9. Y. Fang, S. Lu, and Y. Liu, “Controlling photon transverse orbital angular momentum in high harmonic generation,” Phys. Rev. Lett. 127(27), 273901 (2021). [CrossRef]  

10. R. Zuo, X. Song, S. Ben, et al., “Revealing the nonadiabatic tunneling dynamics in solid-state high harmonic generation,” Phys. Rev. Research 5(2), L022040 (2023). [CrossRef]  

11. W. Yang, X. Song, S. Gong, et al., “Carrier-envelope phase dependence of few-cycle ultrashort laser pulse propagation in a polar molecule medium,” Phys. Rev. Lett. 99(13), 133602 (2007). [CrossRef]  

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

13. C. Van Vlack and S. Hughes, “Carrier-envelope-offset phase control of ultrafast optical rectification in resonantly excited semiconductors,” Phys. Rev. Lett. 98(16), 167404 (2007). [CrossRef]  

14. L. R. Suné, M. Scalora, A. S. Johnson, et al., “Study of second and third harmonic generation from an indium tin oxide nanolayer: Influence of nonlocal effects and hot electrons,” APL Photon. 5(1), 010801 (2020). [CrossRef]  

15. A. Capretti, Y. Wang, N. Engheta, et al., “Comparative study of second-harmonic generation from epsilon-near-zero indium tin oxide and titanium nitride nanolayers excited in the near-infrared spectral range,” ACS Photonics 2(11), 1584–1591 (2015). [CrossRef]  

16. T. T. Liu, T. Kanai, T. Sekikawa, et al., “Significant enhancement of high-order harmonics below 10 nm in a two-color laser field,” Phys. Rev. A 73(6), 063823 (2006). [CrossRef]  

17. C. Chen, C. H. García, Z. Tao, et al., “Influence of microscopic and macroscopic effects on attosecond pulse generation using two-color laser fields,” Opt. Express 25(23), 28684–28696 (2017). [CrossRef]  

18. K. Kondo, Y. Kobayashi, A. Sagisaka, et al., “Tunneling ionization and harmonic generation in two-color fields,” J. Opt. Soc. Am. B 13(2), 424–429 (1996). [CrossRef]  

19. J. Mauritsson, P. Johnsson, E. Gustafsson, et al., “Attosecond pulse trains generated using two color laser fields,” Phys. Rev. Lett. 97(1), 013001 (2006). [CrossRef]  

20. I. J. Kim, C. M. Kim, H. T. Kim, et al., “Highly efficient high-harmonic generation in an orthogonally polarized two-color laser field,” Phys. Rev. Lett. 94(24), 243901 (2005). [CrossRef]  

21. N. B. Tal, N. Moiseyev, and A. Beswick, “The effect of Hamiltonian symmetry on generation of odd and even harmonics,” J. Phys. B 26(18), 3017–3024 (1993). [CrossRef]  

22. T. Zuo, A. D. Bandrauk, M. Ivanov, et al., “Control of high-order harmonic generation in strong laser fields,” Phys. Rev. A 51(5), 3991–3998 (1995). [CrossRef]  

23. F. Navarrete and U. Thumm, “Two-color-driven enhanced high-order harmonic generation in solids,” Phys. Rev. A 102(6), 063123 (2020). [CrossRef]  

24. S. Mitra, S. Biswas, J. Schötz, et al., “Suppression of individual peaks in two-colour high harmonic generation,” J. Phys. B 53(13), 134004 (2020). [CrossRef]  

25. H. Eichmann, A. Egbert, S. Nolte, et al., “Polarization-dependent high-order two-color mixing,” Phys. Rev. A 51(5), R3414–R3417 (1995). [CrossRef]  

26. M. D. Perry and J. K. Crane, “High-order harmonic emission from mixed fields,” Phys. Rev. A 48(6), R4051–R4054 (1993). [CrossRef]  

27. J. B. Bertrand, H. J. Wörner, H.-C. Bandulet, et al., “Ultrahigh-order wave mixing in noncollinear high harmonic generation,” Phys. Rev. Lett. 106(2), 023001 (2011). [CrossRef]  

28. S. Gong, Z. Wang, S. Du, et al., “Coherent control of high-order harmonic generation in a two-level atom driven by intense two-colour laser fields,” J. Mod. Opt. 46(11), 1669–1676 (1999). [CrossRef]  

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

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

31. A. Picón, A. Benseny, J. Mompart, et al., “Transferring orbital and spin angular momenta of light to atoms,” New J. Phys. 12(8), 083053 (2010). [CrossRef]  

32. M. Zürch, C. Kern, P. Hansinger, et al., “Strong-field physics with singular light beams,” Nat. Phys. 8(10), 743–746 (2012). [CrossRef]  

33. G. Gariepy, J. Leach, K. T. Kim, et al., “Creating high harmonic beams with controlled orbital angular momentum,” Phys. Rev. Lett. 113(15), 153901 (2014). [CrossRef]  

34. L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw, et al., “Orbital angular momentum of light and the transformation of Laguerre-Gaussian laser modes,” Phys. Rev. A 45(11), 8185–8189 (1992). [CrossRef]  

35. C. Conti, L. Angelani, and G. Ruocco, “Light diffusion and localization in three-dimensional nonlinear disordered media,” Phys. Rev. A 75(3), 033812 (2007). [CrossRef]  

36. J. Wu, B. A. Malomed, H. Y. Fu, et al., “Self-interaction of ultrashort pulses in an epsilon-near-zero nonlinear material at the telecom wavelength,” Opt. Express 27(26), 37298–37307 (2019). [CrossRef]  

37. H. Ma, Y. Zhao, Y. Shao, et al., “Principles to tailor the saturable and reverse saturable absorption of epsilon-near-zero material,” Photonics Res. 9(5), 678–686 (2021). [CrossRef]  

38. M. Z. Alam , S. A. Schulz and J. Upham, “Large optical nonlinearity of nanoantennas coupled to an epsilon-near-zero material,” Nat. Photonics 12(2), 79–83 (2018). [CrossRef]  

39. J. Wu, Z. Xie, H. Y. Fu, et al., “Numerical investigations on the cascaded high harmonic and quasi-supercontinuum generations in epsilon-near-zero aluminum-doped zinc oxide nanolayers,” Results Phys. 24, 104086 (2021). [CrossRef]  

40. A. Ciattoni and E. Spinozzi, “Efficient second-harmonic generation in micrometer-thick slabs with indefinite permittivity,” Phys. Rev. A 85(4), 043806 (2012). [CrossRef]  

41. B. Johns, N. M. Puthoor, H. Gopalakrishnan, et al., “Epsilon-near-zero response in indium tin oxide thin films: Octave span tuning and IR plasmonics,” J. Appl. Phys. 127(4), 043102 (2020). [CrossRef]  

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

43. J. Koga, “Simulation model for the effects of nonlinear polarization on the propagation of intense pulse lasers,” Opt. Lett. 24(6), 408–410 (1999). [CrossRef]  

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

45. C. Runge, “Ueber die numerische Auflösung von Differentialgleichungen,” Math. Ann. 46(2), 167–178 (1895). [CrossRef]  

46. W. Kutta, “Beitrag zur näherungsweisen Integration totaler Differentialgleichungen,” Z. Math. Phys. 46, 435–453 (1901).

47. C. Argyropoulos, G. D. Aguanno, and A. Alú, “Giant second-harmonic generation efficiency and ideal phase matching with a double ɛ-near-zero cross-slit metamaterial,” Phys. Rev. B 89(23), 235401 (2014). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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. Temporal and spatial evolution of a LG pulse propagating through the ITO material under (a) single-color ω0 or (b) two-color ω0 + 2ω0 excitation with fundamental laser frequency equal to ω0.
Fig. 2.
Fig. 2. Spectral distributions of vortex harmonics (a) at different incident LG pulses and (b) at ω0 + 0.3ω0. The 4ω0 harmonic signal intensity versus the incident laser peak amplitude of (c) ω0 laser for fixing 2ω0 field intensity and (d) 2ω0 field or fixing ω0 field intensity.
Fig. 3.
Fig. 3. Time-frequency analysis to high-order vortex harmonics in Fig. 2(a) under (a) single-color ω0 and (b) two-color ω0 + 2ω0 excitation.
Fig. 4.
Fig. 4. Transverse field distributions of (a) second, (b) third, (c) fourth, (d) fifth, (e) sixth and (f) seventh order harmonics under two-color excitation with fundamental laser frequency equal to ω0.
Fig. 5.
Fig. 5. (a) Spectral distributions of high-order harmonic signals at different laser frequencies ω0. Intensities of (b) fourth-order harmonic and (c) sixth-order harmonic at different material thickness d.

Equations (5)

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

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 ) .
E ( t = 0 , z ) = E 10 sech [ 1.76 z c τ 0 ] cos ( ω 0 z c ) e x + A E 20 sech [ 1.76 z c τ 0 ] cos ( ω 2 z c ) e x
× E = μ 0 H t , × H = ε 0 ε E t + P t ,
2 P t 2 + δ e ω e P t + ω e 2 ( 1 + | P | 2 P s 2 ) 3 / 2 P = ε 0 ( ε s ε ) ω e 2 E .
ω 0 = ω e 2 ε { ( ε s + ε ( 1 δ e 2 ) ) + [ ( ε s + ε ( 1 δ e 2 ) ) 2 4 ε s ε ] 1 / 2 } 1 / 2 .
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.