• Photonics Research
  • Vol. 9, Issue 3, 317 (2021)
Wendong Tian1, Fei Liang1、2、*, Dazhi Lu1, Haohai Yu1、3、*, and Huaijin Zhang1
Author Affiliations
  • 1State Key Laboratory of Crystal Materials and Institute of Crystal Materials, Shandong University, Jinan 250100, China
  • 2e-mail: liangfei@sdu.edu.cn
  • 3e-mail: haohaiyu@sdu.edu.cn
  • show less
    DOI: 10.1364/PRJ.414570 Cite this Article Set citation alerts
    Wendong Tian, Fei Liang, Dazhi Lu, Haohai Yu, Huaijin Zhang. Highly efficient ultraviolet high-harmonic generation from epsilon-near-zero indium tin oxide films[J]. Photonics Research, 2021, 9(3): 317 Copy Citation Text show less

    Abstract

    High-harmonic generation in the ultraviolet region is promising for wireless technology used for communications and sensing. However, small high-order nonlinear coefficients prevent us from obtaining high conversion efficiency and functional photonic devices. Here, we show highly efficient ultraviolet harmonic generation extending to the fifth order directly from an epsilon-near-zero indium tin oxide (ITO) film. The real part of the annealed ITO films was designed to reach zero around 1050 nm, matching with the central wavelength of an Yb-based fiber laser, and the internal driving electric field was extremely enhanced. A high energy conversion efficiency of 10-4 and 10-6 for 257.5 nm (fourth-order) and 206 nm (fifth-order) ultraviolet harmonic generation was obtained, which is at least 2 orders of magnitude higher than early reports. Our results demonstrate a new route for overcoming the inefficiency problem and open up the possibilities of compact solid-state high-harmonic generation sources at nanoscale.

    1. INTRODUCTION

    Nonlinear optical harmonic generation is a typical strong-field physical process, which requires strong electric field intensity to trigger light–matter interaction [13]. For example, second-order nonlinearity, including second-harmonic generation (SHG), optical parametric oscillation (OPO), and sum frequency generation (SFG), extends the laser spectral range from visible to mid-infrared, and even the terahertz (THz) region [46]. Furthermore, high-order harmonic generation (HHG), producing coherent ultraviolet (UV) light, extreme-ultraviolet (EUV) light, and soft X-ray sources [7], is gaining increasing attention for space-to-space communication, laser manufacturing, and attosecond physics [810]. The realization of compact and reliable UV coherent sources at nanoscale is now opening the door to industrial and scientific applications with improved reliability, better material compatibility, and greater resolution owing to their shorter wavelengths, such as wafer scribing, photolithography, and flow cytometry [11,12].

    Since the 1980s, HHG in the UV and EUV region has been studied in atomic gases and bulk crystals for decades [13]. However, the conversion efficiency is very low (<106 for the fifth and higher orders) [14]. According to nonlinear optical principles [3], the lowest-order induced polarization P(2) would be comparable to the linear polarization P(1) when the amplitude of the applied field E is of the order of the atomic electric field strength Eat (5.1×1011  V/m). For condensed matter, first-order susceptibility χ(1) is of the order of unity; hence, second-order susceptibility χ(2) would be expected on the order of 1/Eat, that is, 1012  m/V. For higher-order nonlinearity, the χ(n) susceptibility would be reduced by scaling of (1/Eat)n, and the response is too inefficient to be detectable. Therefore, it remains a great challenge to obtain direct UV HHG with high conversion efficiency, especially pumped by the commercial light sources, e.g.,  the Ti:sapphire laser (λ800  nm) and Yb-based fiber laser (λ1030  nm).

    Clearly, a strong driving field (external and internal) is the sufficient condition for improving the conversion efficiency. By applying an external mid-infrared femtosecond laser with ultrahigh peak power, direct HHG has been realized in ZnO crystal (pump source, 3.2–3.7 μm, up to 27th order) [15] and MoS2 monolayer (pump source, 4.1 μm, up to 13th order) [16]. Besides, another strategy to avoid the inefficiency problem is the enhancement of the internal electric field. This could be realized in some special nonlinear media with a refractive index (permittivity epsilon) for the interacting wavelengths near zero [17]. Based on Maxwell’s equations and boundary conditions of the electric displacement vector, the epsilon-near-zero (ENZ) effect can greatly boost the internal laser field, which is appealing for highly efficient nonlinear harmonic generation. As presented in artificial metamaterials with zero refractive index at 1330 nm [18], the conversion efficiency of the third-order four-wave mixing (FWM) was increased to 105. Moreover, the ENZ effect is also available in some semiconductor films with tunable ENZ wavelength λENZ by doping or annealing [19,20], such as Sn:In2O3 (ITO), Al:ZnO (AZO), and Y:CdO. To our best knowledge, only Yang et al. reported a UV seventh-order harmonic wave (λ297  nm) from In-doped CdO film pumped by a complex Ti:sapphire optical parametric amplifier (OPA) system at 2080 nm [21], in which the energy conversion efficiency is on the order of 1010.

    Herein, we reported UV nonlinear harmonic generation from indium tin oxide (ITO) film epitaxially grown on a silicon substrate by magnetron sputtering. ITO films were selected because their carrier concentration could be adjusted by annealing treatment, and the real part of its permittivity reaches zero around 1050 nm, thus matching with the central wavelength of the Yb-based fiber laser without additional nonlinear conversion [22]. Using a 1030 nm femtosecond laser, odd- and even-order harmonic waves up to fifth order (second, third, fourth, and fifth) were observed, corresponding to the shortest UV wavelength at 206 nm, with an unprecedented high conversion efficiency of 106 for the fifth order. This work opens new functionalities and applications of ENZ media in nonlinear optics [23,24], such as high-order frequency conversion and entangled photon sources directly pumped by a Ti:sapphire laser and Yb-fiber laser.

    2. FABRICATION AND CHARACTERIZATION OF THE ITO SAMPLE

    (a) X-ray diffraction (XRD) pattern of ITO film; the standard diffraction of In2O3 is plotted as comparison. The inset graph is the enlargement from 29° to 32°. (b) SEM image of the ITO film. (c) AFM characterization of ITO film. (d) Raman spectrum of ITO film.

    Figure 1.(a) X-ray diffraction (XRD) pattern of ITO film; the standard diffraction of In2O3 is plotted as comparison. The inset graph is the enlargement from 29° to 32°. (b) SEM image of the ITO film. (c) AFM characterization of ITO film. (d) Raman spectrum of ITO film.

    The morphology and thickness of ITO film were characterized by scanning electron microscopy (SEM) and atomic force microscopy (AFM). In Fig. 1(b), the SEM image of oblique ITO film is rough and granular. This imperfectly flat surface is favorable for following nonlinear optical experiments as to avoid zero electric field component along the z direction and reduced interaction strength between the incident light electric field and the ITO surface [25]. Figure 1(c) demonstrates that the thickness of ITO film is around 103 nm. It is worth emphasizing that this is shorter than coherent length (tens of micrometers) of harmonic generation in nonlinear process. Therefore, phase mismatch along the forward direction is nearly negligible. In addition, Raman spectroscopy was measured with a 532 nm excitation laser in Fig. 1(d). There are five strong Raman active modes with Raman shifts at 305, 364, 405, 492, and 627  cm1, respectively, which is consistent with previous reports [26].

    3. PERMITTIVITY AND HIGH-HARMONIC GENERATION IN THE ITO FILM

    ITO film is a common transparent conducting oxide film with adjustable carrier concentration and exhibits near-zero permittivity around the near-infrared region. According to Maxwell’s equations and boundary conditions of the electric displacement vector, the reduced permittivity will bring an enhanced electric field, which is favorable to strengthen the light–matter interaction and is appealing for obtaining phase mismatch-free second-order harmonic generation and even HHG.

    (a) Calculated real and imaginary parts of the permittivity of ITO film as a function of wavelength. (b) Calculated field enhancements of the ITO film as a function of incident laser wavelength. (c) Calculated field enhancements of the structure as a function of location at 1.03 μm. (d) Schematic of nonlinear harmonic generation on the ENZ sample composed of silicon substrate and ITO film. The pump laser is vertically incident from the ITO side. Notably, the light beam direction of harmonic waves is only a diagrammatic sketch, not the real propagating direction. (e) The harmonic spectrum under different incident pump power densities. The intensity of the third-, fourth-, and fifth-harmonic generation was enlarged for clarity.

    Figure 2.(a) Calculated real and imaginary parts of the permittivity of ITO film as a function of wavelength. (b) Calculated field enhancements of the ITO film as a function of incident laser wavelength. (c) Calculated field enhancements of the structure as a function of location at 1.03 μm. (d) Schematic of nonlinear harmonic generation on the ENZ sample composed of silicon substrate and ITO film. The pump laser is vertically incident from the ITO side. Notably, the light beam direction of harmonic waves is only a diagrammatic sketch, not the real propagating direction. (e) The harmonic spectrum under different incident pump power densities. The intensity of the third-, fourth-, and fifth-harmonic generation was enlarged for clarity.

    In the nonlinear optical process, the nth-order electric polarization intensity P(n) depends on the incident light electric field intensity E as follows [3]:P(n)=ε0χ(n)En,where ε0 is the vacuum epsilon and χ(n) is the nth-order nonlinear susceptibility. With the extreme enhancement of the light electric field intensity at the ITO film interface [Fig. 2(c)], the induced nonlinear electric polarization intensity will be strongly increased, thereby leading to possible nonlinear HHG at nanoscale films.

    The schematic illustration and experimental setup for nonlinear high-harmonic generation are illuminated in Fig. 2(d). A mode-locked femtosecond laser system (central wavelength 1030  nm, pulse width 250  fs, and repetition rate 200  kHz) was focused by a 100× objective to a spot radius of 10 μm. The high-harmonic signals were collected from reflected signals via the same objective using charge-coupled device (CCD) camera, and the respective powers of nth-order harmonics were measured by a powermeter. It is worthy to point out that the light beam direction in Fig. 2(d) is only a diagrammatic sketch, not the real propagating direction of high-harmonic waves.

    Figure 2(e) depicts the high-harmonic generation in ITO film pumped by 1030 nm femtosecond laser. Clearly, the second-order harmonic generation was first measured with incident power intensity of 35.6  GW·cm2, corresponding to a vacuum field strength of 0.52  V·nm1. With gradually improved pumped intensity, the third-, fourth-, and fifth-harmonic waves start to emerge with excitation intensity at 45.2, 371.8, 891.2  GW·cm2 (vacuum field strength of 0.58, 1.67, and 2.59  V·nm1), respectively. The improved excitation threshold for fourth- and fifth-harmonic generation may be attributed to additional absorption of ITO film at 257.5 and 206 nm. Notably, this excitation energy is comparable to the typical intensity used for HHG from bulk or monolayer solids [13]. For instance, 0.5  TW·cm2 (3.25 μm, ninth order) in zinc oxide [15], and 0.7  TW·cm2 (4.1 μm, seventh order) in MoS2 monolayer [16]. Moreover, this excitation intensity is slightly higher than that for ENZ In:CdO film [21] (2.08 μm, 0.3  GW·cm2 for the third order and 0.5  GW·cm2 for the fifth order), in which the real part of its permittivity crosses zero at a wavelength of 2.09 μm. The excitation intensity for HHG in ITO film would be further reduced when the central wavelength is nearer to 1051 nm due to the large resonant field enhancement.

    (a) Spectrum of the fundamental wave. (b) The second-harmonic spectrum of ITO film and GaAs (111) crystal with pump intensity of 42.2 GW⋅ cm−2. (c) The third-harmonic spectrum of ITO film and Si (100) with pump intensity of 51.4 GW⋅ cm−2. (d) The fourth-harmonic and (e) fifth-harmonic spectra of ITO film with pump intensity of 438 and 944 GW⋅ cm−2. (f) The generated harmonic power as a function of incident pump power. Solid lines show the power scaling law In and are plotted to guide the eye.

    Figure 3.(a) Spectrum of the fundamental wave. (b) The second-harmonic spectrum of ITO film and GaAs (111) crystal with pump intensity of 42.2  GWcm2. (c) The third-harmonic spectrum of ITO film and Si (100) with pump intensity of 51.4  GWcm2. (d) The fourth-harmonic and (e) fifth-harmonic spectra of ITO film with pump intensity of 438 and 944  GWcm2. (f) The generated harmonic power as a function of incident pump power. Solid lines show the power scaling law In and are plotted to guide the eye.

    Same as the experimental facility above, the third-harmonic response of ITO film was studied with the wavelength of pump light at 1030 nm and the incident power of 51.4  GWcm2. In order to estimate the third-order susceptibility of ITO film, Si (100) wafer was selected as a reference material with a third-order susceptibility of 1.05×1016  m2V2 [33]. In Fig. 3(c),the third-harmonic generation intensity of ITO film is 9.42 times stronger than that of Si (100) wafer. According to third-order nonlinear optical theory, the third-harmonic intensity (I(3ω)) can be determined byI(3ω)(3ω2cn(3ω))2|χ(3)E3|2,where χ(3) is the third-order susceptibility. Similar to SHG discussed above, the electric field in Si medium decreases to 1/3.95, due to the epsilon of Si (100) being 3.95 at 1030 nm [34]. Therefore, the internal electric field of ITO film is nearly 23.5 times stronger than that on Si wafer. Taking this enhancement factor into account, the third-order susceptibility χ(3) of ITO film is determined to be 2.48×1020  m2V2, which is in good agreement with previous studies (2.37×1020  m2V2 at 1.9 μm) [35] and 4 orders of magnitude smaller than that of Si wafer. This also confirms that the resonantly boosted electric field brings enhanced third-order nonlinear response. The third-harmonic intensity keeps unchanged with the variation of the polarization angle of incident laser, which is also consistent with in-plane isotropy.

    More impressively, due to a strongly enhanced electric field, we also measured direct fourth- and fifth-harmonic generation on ENZ ITO film, which is very rare in common nonlinear optical materials. As shown in Figs. 3(d) and 3(e), the fourth- and fifth-harmonic generation signals are measured by a CCD camera with the incident power of 438  GWcm2 and 944  GWcm2, respectively. When we continue to improve the incident pump intensity, a supercontinuum spectrum from 303 to 945 nm was collected under incident power of 9.56  TWcm2. Generally, supercontinuum generation on a surface is a challenging task because the surficial nonlinear optical efficiency is usually limited by finite light–matter interaction length [25]. Benefitting from a greatly enhanced electric field, a high conversion efficiency up to 1.2% is obtained in supercontinuum generation under 29.3  TWcm2 incident power. It is credible to measure the sixth-, seventh-, and even higher-order harmonic waves, however, beyond the detectable spectral range of our experimental setup.

    The output powers of high-harmonic generation are also measured quantitatively. Figure 3(f) shows the harmonic intensity as a function of the excitation intensity, with perfect scaling law I2, I3, and I4 for second-, third-, and fourth-harmonic waves. The fifth-harmonic energy was estimated from the measured fourth-harmonic energy by referring to Fig. 2(e) [21]. Clearly, the fifth-harmonic intensity slightly deviates from fifth-order power law. This could be attributed to two reasons. (i) The fifth-harmonic wave power is too low to be detected by our powermeter. The fluctuation of relative intensity of fourth- and fifth-harmonic waves could bring larger error and power deviation. (ii) Under a strong expiation field, the non-perturbative effect starts to become significant for high-harmonic generation. This case has been reported in many early HHG experiments [11,12,16,21].

    The energy conversion efficiency for the second-, third-, fourth-, and fifth-harmonic generation is determined to be around 3.20×103, 7.05×104, 1.59×104, and 7.08×106, respectively, under a pump intensity of 5.01  TWcm2. This efficiency is a strongly improved contrast to typical efficiency (1010) of second-order harmonic generation on the surfaces [36]. Meanwhile, it is nearly 2 orders of magnitude higher than that in In:CdO film (105 for the third, 108 for the fifth, 1010 for the seventh harmonics) [21]. Taking the short interaction length in the ITO film (hundreds of nanometers) into account, this unprecedented efficiency suggests the giant potential of ENZ materials in highly integrated nonlinear optoelectronics.

    (a) Experimental setup for 515 nm pumped second-harmonic generation in ITO film. HR and HT represent high reflectivity (>99%) and high transmittance (>95%). (b) Experimental setup for 1030 nm pumped fourth-harmonic generation in ITO film. (c) The second-harmonic power as a function of incident 515 nm pump intensity and the fourth-harmonic power as a function of incident 1030 nm pump intensity.

    Figure 4.(a) Experimental setup for 515 nm pumped second-harmonic generation in ITO film. HR and HT represent high reflectivity (>99%) and high transmittance (>95%). (b) Experimental setup for 1030 nm pumped fourth-harmonic generation in ITO film. (c) The second-harmonic power as a function of incident 515 nm pump intensity and the fourth-harmonic power as a function of incident 1030 nm pump intensity.

    Besides ITO film, the ENZ effect is also available in other materials, including ZrN (λENZ485.4  nm) [37], TiN (λENZ520  nm) [31], AZO (λENZ1550  nm) [25], high concentration Y-doped CdO (λENZ5.3  μm) [38], CdO (λENZ11.3  μm) [38], and various artificial metamaterials. Especially, the ENZ wavelengths of Au film (λENZ796  nm) [39] and Ag-SiO2 multilayer films (λENZ820890  nm) [40] are matched to the Ti:sapphire laser. Therefore, this enhanced high-harmonic generation could be obtained in wide spectral range by applying suitable ENZ media. This λENZ modulation strategy by annealing treatment is also significant for other nonlinear processes pumped by the Ti:sapphire laser and fiber laser, such as difference frequency generation (DFG), optical rectification (OR), FWM, and THz generation.

    4. CONCLUSION

    In summary, we reported the UV high-harmonic generation in annealed ITO film for the first time. Pumped by an Yb-based 1030 nm femtosecond laser, odd- and even-order harmonic waves up to fifth order at 206 nm were observed with an unprecedented high conversion efficiency of 106. Assisted by the extremely enhanced electric field in ENZ region and tunable ENZ wavelength, ITO film provides a newfangled platform to study nonlinear light–matter interaction within hundreds of nanometers thickness, including strong-field physics and attosecond science. Meanwhile, ITO film is compatible with silicon photonics, which could have further applications in highly integrated nonlinear photonic devices based on nanolayers.

    References

    [1] T. H. Maiman. Stimulated optical radiation in ruby. Nature, 187, 493-494(1960).

    [2] P. A. Franken, G. Weinreich, C. W. Peters, A. E. Hill. Generation of optical harmonics. Phys. Rev. Lett., 7, 118-120(1961).

    [3] R. W. Boyd. Nonlinear Optics(2008).

    [4] C. T. Chen, G. L. Wang, X. Y. Wang, Z. Y. Xu. Deep-UV nonlinear optical crystal KBe2BO3F2—discovery, growth, optical properties and applications. Appl. Phys. B, 97, 9-25(2009).

    [5] V. Petrov. Frequency down-conversion of solid-state laser sources to the mid-infrared spectral range using non-oxide nonlinear crystals. Prog. Quantum Electron., 42, 1-106(2015).

    [6] D. Yan, Y. Wang, D. Xu, P. Liu, C. Yan, J. Shi, H. Liu, Y. He, L. Tang, J. Feng, J. Guo, W. Shi, K. Zhong, Y. H. Tsang, J. Yao. High-average-power, high-repetition-rate tunable terahertz difference frequency generation with GaSe crystal pumped by 2  μm dual-wavelength intracavity KTP optical parametric oscillator. Photon. Res., 5, 82-87(2017).

    [7] F. Krausz, M. Ivanov. Attosecond physics. Rev. Mod. Phys., 81, 163-234(2009).

    [8] Z. Xu, B. M. Sadler. Ultraviolet communications: potential and state-of-the-art. IEEE Commun. Mag., 46, 67-73(2008).

    [9] T. Popmintchev, M. C. Chen, D. Popmintchev, P. Arpin, S. Brown, S. Ališauskas, G. Andriukaitis, T. Balčiunas, O. D. Mücke, A. Pugzlys, A. Baltuška, B. Shim, S. E. Schrauth, A. Gaeta, C. Hernández-García, L. Plaja, A. Becker, A. Jaron-Becker, M. M. Murnane, H. C. Kapteyn. Bright coherent ultrahigh harmonics in the keV X-ray regime from mid-infrared femtosecond lasers. Science, 336, 1287-1291(2012).

    [10] A. J. Uzan, G. Orenstein, A. Jimenez-Galan, C. McDonald, R. E. F. Silva, B. D. Bruner, N. D. Klimkin, V. Blanchet, T. Arusi-Parpar, M. Kruger, A. N. Rubtsov, O. Smirnova, M. Ivanov, B. H. Yan, T. Brabec, N. Dudovich. Attosecond spectral singularities in solid-state high-harmonic generation. Nat. Photonics, 14, 183-188(2020).

    [11] N. Yoshikawa, T. Tamaya, K. Tanaka. High-harmonic generation in graphene enhanced by elliptically polarized light excitation. Science, 356, 736-738(2017).

    [12] M. Sivis, M. Taucer, G. Vampa, K. Johnston, A. Staudte, A. Yu. Naumov, D. M. Villeneuve, C. Ropers, P. B. Corkum. Tailored semiconductors for high-harmonic optoelectronics. Science, 357, 303-306(2017).

    [13] S. Ghimire, D. A. Reis. High-harmonic generation from solids. Nat. Phys., 15, 10-16(2018).

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

    [15] S. Ghimire, A. D. DiChiara, E. Sistrunk, P. Agostini, L. F. DiMauro, D. A. Reis. Observation of high-order harmonic generation in a bulk crystal. Nat. Phys., 7, 138-141(2010).

    [16] H. Liu, Y. Li, Y. S. You, S. Ghimire, T. F. Heinz, D. A. Reis. High-harmonic generation from an atomically thin semiconductor. Nat. Phys., 13, 262-265(2016).

    [17] M. Z. Alam, I. De Leon, R. W. Boyd. Large optical nonlinearity of indium tin oxide in its epsilon-near-zero region. Science, 352, 795-797(2016).

    [18] H. Suchowski, K. O’Brien, Z. J. Wong, A. Salandrino, X. B. Yin, X. Zhang. Phase mismatch–free nonlinear propagation in optical zero-index materials. Science, 342, 1223-1227(2013).

    [19] Y. Wang, A. Capretti, L. Dal Negro. Wide tuning of the optical and structural properties of alternative plasmonic materials. Opt. Mater. Express, 5, 2415-2430(2015).

    [20] J. Zhao, H. Zhang, X. Zhang, D. Li, H. Lu, M. Xu. Abnormal behaviors of Goos–Hänchen shift in hyperbolic metamaterials made of aluminum zinc oxide materials. Photon. Res., 1, 160-163(2013).

    [21] Y. Yang, J. Lu, A. Manjavacas, T. S. Luk, H. Liu, K. Kelley, J.-P. Maria, E. L. Runnerstrom, M. B. Sinclair, S. Ghimire, I. Brener. High-harmonic generation from an epsilon-near-zero material. Nat. Phys., 15, 1022-1026(2019).

    [22] Y. Wang, A. C. Overvig, S. Shrestha, R. Zhang, R. Wang, N. Yu, L. Dal Negro. Tunability of indium tin oxide materials for mid-infrared plasmonics applications. Opt. Mater. Express, 7, 2727-2739(2017).

    [23] I. Liberal, N. Engheta. Near-zero refractive index photonics. Nat. Photonics, 11, 149-158(2017).

    [24] Q. Guo, Y. Cui, Y. Yao, Y. Ye, Y. Yang, X. Liu, S. Zhang, X. Liu, J. Qiu, H. Hosono. A solution-processed ultrafast optical switch based on a nanostructured epsilon-near-zero medium. Adv. Mater., 29, 1700754(2017).

    [25] W. Tian, F. Liang, S. Chi, C. Li, H. Yu, H. Zhang, H. Zhang. Highly efficient super-continuum generation on an epsilon-near-zero surface. ACS Omega, 5, 2458-2464(2020).

    [26] H. Kim, C. M. Gilmore, A. Pique, J. S. Horwitz, H. Mattoussi, H. Murata, Z. H. Kafafi, D. B. Chrisey. Electrical, optical and structural properties of indium-tin-oxide thin films for organic light-emitting devices. J. Appl. Phys., 86, 6451-6461(1999).

    [27] V. M. Shalaev. Optical negative-index metamaterials. Nat. Photonics, 1, 41-48(2007).

    [28] G. V. Naik, J. L. Schroeder, X. Ni, A. V. Kildishev, T. D. Sands, A. Boltasseva. Titanium nitride as a plasmonic material for visible and near-infrared wavelengths. Opt. Mater. Express, 2, 478-489(2012).

    [29] S. Bergfeld, W. Daum. Second-harmonic generation in GaAs: experiment versus theoretical predictions of χxyz(2). Phys. Rev. Lett., 90, 036801(2003).

    [30] J. S. Blakemore. Semiconducting and other major properties of gallium-arsenide. J. Appl. Phys., 53, R123-R181(1982).

    [31] A. Capretti, Y. Wang, N. Engheta, L. Dal Negro. 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 Photon., 2, 1584-1591(2015).

    [32] W. J. Wang, J. H. Xu, X. Liu, Y. Q. Jiang, G. M. Wang, X. Z. Lu. Second harmonic generation investigation of indium tin oxide thin films. Thin Solid Films, 365, 116-118(2000).

    [33] Y. J. Chen, G. M. Carter. Measurement of third order nonlinear susceptibilities by surface plasmons. Appl. Phys. Lett., 41, 307-309(1982).

    [34] G. Lin, Y. Chang, E. Liu, H. Kuo, H. Lin. Low refractive index Si nanopillars on Si substrate. Appl. Phys. Lett., 90, 181923(2007).

    [35] N. Ueda, H. Kawazoe, Y. Watanabe, M. Takata, M. Yamane, K. Kubodera. Third-order nonlinear optical susceptibilities of electroconductive oxide thin films. Appl. Phys. Lett., 59, 502-503(1991).

    [36] C. S. Tian, Y. R. Shen. Recent progress on sum-frequency spectroscopy. Surf. Sci. Rep., 69, 105-131(2014).

    [37] O. Reshef, I. De Leon, M. Z. Alam, R. W. Boyd. Nonlinear optical effects in epsilon-near-zero media. Nat. Rev. Mater., 4, 535-551(2019).

    [38] S. Saha, B. T. Diroll, J. Shank, Z. Kudyshev, A. Dutta, S. N. Chowdhury, T. S. Luk, S. Campione, R. D. Schaller, V. M. Shalaev, A. Boltasseva, M. G. Wood. Broadband, high‐speed, and large‐amplitude dynamic optical switching with yttrium‐doped cadmium oxide. Adv. Funct. Mater., 29, 1908377(2019).

    [39] I. De Leon, Z. Shi, A. C. Liapis, R. W. Boyd. Measurement of the complex nonlinear optical response of a surface plasmon-polariton. Opt. Lett., 39, 2274-2277(2014).

    [40] R. M. Kaipurath, M. Pietrzyk, L. Caspani, T. Roger, M. Clerici, C. Rizza, A. Ciattoni, A. Di Falco, D. Faccio. Optically induced metal-to-dielectric transition in epsilon-near-zero metamaterials. Sci. Rep., 6, 27700(2016).

    Wendong Tian, Fei Liang, Dazhi Lu, Haohai Yu, Huaijin Zhang. Highly efficient ultraviolet high-harmonic generation from epsilon-near-zero indium tin oxide films[J]. Photonics Research, 2021, 9(3): 317
    Download Citation