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

Highly efficient Cherenkov-type terahertz generation by 2-μm wavelength ultrashort laser pulses in a prism-coupled LiNbO3 layer

Open Access Open Access

Abstract

Terahertz generation by optical rectification of femtosecond laser pulses propagating in a 40-$\mu$m thick LiNbO$_3$ layer attached to an output Si prism has been experimentally investigated for different laser wavelengths from 800 to 2100 nm. For longer wavelengths, the saturation of the optical-to-terahertz conversion efficiency has been observed at higher laser pulse energies, thus enabling higher efficiencies. In particular, record high conversion efficiency of 1.3% has been achieved with 30-$\mu$J laser pulse energy at 2100 nm.

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

1. Introduction

Optical rectification of near-infrared femtosecond laser pulses in lithium niobate (LiNbO$_3$) is an efficient method of generating intense pulsed terahertz radiation in the $\sim$0.1-2 THz frequency range on a tabletop scale. Intense electric and magnetic fields in this frequency range are in high demand for such new applications as terahertz-driven particle acceleration [13], molecular orientation and alignment [46], ultrafast spin control [7,8], and terahertz streaking [9,10]. To circumvent large velocity mismatch between terahertz waves and optical pump in LiNbO$_3$, pumping by laser pulses with tilted pulse front [11] is commonly used. The tilted-pulse-front technique requires mJ-level pumping laser pulse energy to provide both high optical intensity and a beam width much larger than the terahertz wavelength. The typical optical-to-terahertz conversion efficiency of this technique at room temperature is $\sim$0.1$\%$ [12,13]. The record efficiency of $\sim$1$\%$ has been achieved with 1.2-mJ pump pulses of optimized chirp and duration [14]. The record high terahertz pulse energy of 0.4 mJ was achieved with 60-mJ pump pulses [15].

Another way to circumvent optical-terahertz velocity mismatch in LiNbO$_3$ is Cherenkov scheme [1618]. The Cherenkov scheme requires focusing of the pump laser beam to a size smaller than the terahertz wavelength in one (focusing to a line) or two (focusing to a spot) directions, to prevent destructive interference of terahertz waves emitted by different parts of the nonlinear source. Tight focusing allows one to achieve high optical intensity, required for efficient nonlinear optical rectification, with lower energy laser pulses than in the tilted-pulse-front configuration. In recent years, the Cherenkov scheme has been substantially elaborated [1928]. At present, one of the most efficient Cherenkov-type terahertz emitters consists of a few tens of microns thick and $\sim$1$\times$1 cm$^2$ in size layer of LiNbO$_3$ attached to an output prism made of a high-resistivity Si [21,22,2426]. The pump laser beam is focused by a cylindrical lens to a line on the entrance facet of the LiNbO$_3$ layer. The laser pulse propagates in the layer as a fundamental mode of an oversized dielectric slab waveguide and emits Cherenkov half-wedge of terahertz waves to free space through the Si prism. Using a thin LiNbO$_3$ layer not only prevents diffraction broadening of the pump laser beam but also reduces attenuation of terahertz waves in strongly absorbing LiNbO$_3$. (Structures with a thinner, a few microns thick, LiNbO$_3$ layer [23] or with a micron-size LiNbO$_3$ ridge waveguide [27,28] can provide a broader bandwidth but suffer from a very low coupling efficiency of the pump laser beam to the optical waveguide.) In the first experimental demonstration of the emitter [22], the efficiency 0.1$\%$ was achieved by converting 40-$\mu$J Ti:sapphire laser pulses in a 8-mm long structure with a 50-$\mu$m thick LiNbO$_3$ layer placed between a Si prism and a BK7-glass substrate. More recently, 15-20-$\mu$J pulses from a Ti:sapphire amplified laser system were converted into terahertz radiation with an efficiency as high as 0.25$\%$ in a 1-cm long structure with a 35-$\mu$m thick LiNbO$_3$ layer placed between a Si prism and a metal substrate [25]. Thus, the Cherenkov-type emitters demonstrate practically the same conversion efficiency as the tilted-pulse-front technique but at lower (tens of $\mu$J) pump pulse energies.

The efficiency of the Cherenkov-type emitters saturates with increasing the pump pulse energy [25]. This limits the maximum efficiency. The saturation has been mainly attributed to higher-order nonlinear effects, such as self-phase modulation, self-focusing, and multi-photon absorption of laser radiation [25]. Since these effects are less pronounced for longer optical wavelengths, one can expect that increasing the pump wavelength will increase the saturation pump pulse energy. It will allow to achieve a higher efficiency by increasing the pump pulse energy. The potential of long-wavelength pumping for efficient terahertz generation has been recently recognized [2931] and experimentally demonstrated [3234] for tilted-pulse-front excitation of semiconductors, where two-photon absorption of the pump and generation of free carriers absorbing terahertz radiation are the main limiting factors. Here, we experimentally examine the effect of long-wavelength pumping on the efficiency of Cherenkov-type emitters with a thin LiNbO$_3$ layer, for which free-carrier absorption of terahertz radiation is not so crucial.

2. Experimental

The experimental scheme is shown in Fig. 1. An amplified Ti:sapphire laser with 800-nm wavelength and 50-fs pulse duration pumps an optical parametric amplifier (OPA), which has output with tunable wavelength ranging from 1300 to 2100 nm. The pulse duration of the OPA output radiation remains almost the same as that of the 800-nm exciting laser (we take it equal to 60 fs in the intensity estimations below). The beam from OPA or directly from the Ti:sapphire laser is focused by a cylindrical lens (with the focal length of 40 mm for the OPA beam or 100 mm for the laser beam) to a line on the entrance facet of a 40-$\mu$m thick, 1-cm wide, and 9-mm long LiNbO$_3$ layer attached to a rectangular Si prism with the apex angle of 41$^\circ$. (To fabricate the structure, a mm-thick slab of LiNbO$_3$ was glued to the Si prism, ground down to 40 $\mu$m, and then its facets were polished.) An iris diaphragm is used to adjust the optimal focal beam width ($\approx$20 $\mu$m) for efficiently exciting the fundamental mode of the LiNbO$_3$ slab waveguide [21]. The optical ([001]) axis of the LiNbO$_3$ crystal is in the layer plane and orthogonal to the laser propagation direction. The polarization of the pump optical beam is aligned parallel to the axis in order to maximize the optical rectification coefficient [21]. The apex angle of the prism is complimentary to the Cherenkov emission angle in Si $\theta _c\approx 49^\circ$. The emission angle is defined by formula $\cos \theta _c=n_g/n_\textrm {THz}$, where $n_g\approx 2.22$ [35] is the extraordinary optical group refractive index of LiNbO$_3$ ($n_g$ varies insignificantly over the used wavelength range) and $n_\textrm {THz}\approx 3.418$ is the terahertz phase refractive index of Si [36].

 figure: Fig. 1.

Fig. 1. Schematics of the experimental setup.

Download Full Size | PDF

The energy of the terahertz radiation emitted from the output prism was detected by a Golay cell, which was screened from stray optical radiation by a Si wafer and a low pass filter (LPF). To measure the energy spectrum of the terahertz radiation, terahertz band-pass filters were placed in front of the Golay cell. The optical beam profile and the optical power at the exit facet of the LiNbO$_3$ layer were monitored by means of a CCD camera and a power meter.

3. Results and discussion

Figure 2 shows the conversion efficiency as a function of the pump pulse energy for different pump wavelengths. Since the radius of the incident optical beam was different for 800 nm and longer wavelengths, we normalized the optical pulse energy to a unit length of the focal line, for a correct comparison of the dependences. The upper scale in Fig. 2 shows the corresponding values of the incident pump intensity, which were substantially lower than the damage threshold for LiNbO$_3$ [37,38]. In general, all curves in Fig. 2 exhibit a linear increase in the low-energy region followed by a saturation at higher energies. The slope of the initial linear segment of the curves is slightly lower for longer wavelengths. It can be explained by a decreasing wavelength dependence of the nonlinear coefficient $|d_{33}|$ of LiNbO$_3$ [3941].

 figure: Fig. 2.

Fig. 2. Conversion efficiency as a function of the pump pulse energy (per unit length of the focal line) for different pump wavelengths. The star points show the efficiency when using a metal substrate. The upper scale shows the incident pump intensity.

Download Full Size | PDF

For 800 nm, the efficiency saturates at $\sim$0.4% and $\sim$50-$\mu$J/cm pump pulse energy per unit length (or $\sim$20 $\mu$J of total pulse energy), in agreement with previous findings [25]. For longer wavelengths, the efficiency saturates at larger pump energies per unit length and reaches higher maximum values, according to our expectations. In particular, the efficiency as high as 1.2% is achieved at $\sim$300-$\mu$J/cm (or $\sim$30-$\mu$J) pulse energy of 2100-nm pump. Interestingly, placing a metal plate on the surface of the LiNbO$_3$ layer, which was opposite to the Si prism, allowed us to enhance the efficiency by about 10% (two star points in Fig. 2). For 800-nm pump, a similar metal substrate-induced enhancement of the efficiency was observed in [25].

For all the wavelengths, the efficiency saturation (and subsequent decline seen in Fig. 2) can be attributed to the nonlinear distortion and absorption of the pump pulse in LiNbO$_3$. This is confirmed by Fig. 3. Firstly, Fig. 3(a) illustrates a correlation between the saturation of the conversion efficiency and a knee in the curve of the transmitted optical energy. Both the saturation and the knee occur at $\sim$100 $\mu$J/cm. This correlation indicates that multi-photon absorption (namely, 5-photon absorption at 1500 nm in Fig. 3 [34]) and possibly other losses, such as related to the nonlinear distortion of the optical beam in the LiNbO$_3$ layer, indeed play a role in the efficiency saturation. Secondly, Figs. 3(b)–3(e) illustrate that the optical beam indeed experiences strong nonlinear distortion in the LiNbO$_3$ layer at high pump energies. In particular, the transverse profile of the optical beam changes drastically with increasing the pump energy. At low pump energies [Figs. 3(b) and 3(c)], the beam is $\sim$20 $\mu$m thick and, accordingly, fills only a part of the LiNbO$_3$ layer thickness. This indicates that the beam propagates in the central part of the layer as a fundamental mode of a dielectric slab waveguide. The light intensity distribution is uniform along the width of the optical beam [vertical direction in Figs. 3(b)–3(e)]. At high pump energies [Figs. 3(d) and 3(e)], the optical beam thickens and fills the whole output aperture of the LiNbO$_3$ layer. This indicates excitation of the higher order modes of the dielectric slab waveguide. The presence of hot spots in the beam profile is a result of nonlinear beam distortion due to high order nonlinear effects, such as self-focusing and self-phase modulation.

 figure: Fig. 3.

Fig. 3. (a) Conversion efficiency and transmitted optical pulse energy as functions of the pump pulse energy (per unit length of the focal line) for 1500 nm. (b)–(e) Images of the transmitted laser beam profile at different pump pulse energies.

Download Full Size | PDF

Figure 4(a) shows the terahertz energy spectra measured with pumping at different wavelengths. The pump energy was taken near the saturation value: 30 $\mu$J/cm for 800 nm, 90 $\mu$J/cm for 1300 nm, 220 $\mu$J/cm for 1500 nm, and 200 $\mu$J/cm for 2000 nm. For all wavelengths, the spectra extend from $\sim$0.2 to $\sim$3 THz. The two-peak shape of the spectra agrees, in general, with the spectrum measured by the standard electro-optic sampling for 800-nm pump [25]. The ratio of the spectral peaks is, however, different for 800 nm and longer (1300-2000 nm) wavelengths. For 800 nm, the main peak is at $\sim$2 THz. For longer wavelengths, the main peaks are at $\sim$0.5 THz; there are also peaks at $\sim$2 THz but they are less pronounced.

 figure: Fig. 4.

Fig. 4. (a) Terahertz energy spectra (normalized to unity) for different pump wavelengths at the saturation pump energies. (b) Terahertz energy spectra for 1500-nm, 200 $\mu$J/cm pump with and without metal substrate. (c) Terahertz energy spectra for 1300-nm pump at low and high pump energies.

Download Full Size | PDF

Figure 4(b) confirms, in general, the theoretical prediction of [24] (see also [25]) that placing a metal substrate should accentuate the lower frequencies in the generated spectrum.

Figure 4(c) shows that pumping at high energies generates lower-frequency spectrum than pumping at low energies. This can be explained by an increase in the optical beam thickness due to its nonlinear distortion at high pump energies [see Figs. 3(d) and 3(e)]. Indeed, the optical beam transverse size enters the effective optical pulse duration that defines the generated spectrum [42].

4. Conclusion

We have achieved about three-fold enhancement (up to 1.3%) in the optical-to-terahertz conversion efficiency of a LiNbO$_3$-based Cherenkov-type emitter by increasing the optical pump wavelength from 800 to 2100 nm. The enhancement has been obtained due to suppression of the efficiency saturation by long-wavelength pumping. The efficiency of 1.3% achieved with 30-$\mu$J pump pulses is record high for sub-mJ-level of optical pumping.

In comparison to optical rectification in organic crystals and tilted-pulse-front technique, the Cherenkov-type emitters provide a unique combination of high ($\sim$1%) conversion efficiency at room temperature with a wide interval of possible pump wavelengths and sub-mJ-level of optical pumping, which can be provided by high (MHz) repetition rate laser amplifiers. Such emitters hold promise for generating mW-level average terahertz power.

Funding

Russian Science Foundation (18-19-00486).

Disclosures

The authors declare no conflicts of interest.

References

1. E. A. Nanni, W. R. Huang, K.-H. Hong, K. Ravi, A. Fallahi, G. Moriena, R. J. Miller, and F. X. Kärtner, “Terahertz-driven linear electron acceleration,” Nat. Commun. 6(1), 8486 (2015). [CrossRef]  

2. D. A. Walsh, D. S. Lake, E. W. Snedden, M. J. Cliffe, D. M. Graham, and S. P. Jamison, “Demonstration of sub-luminal propagation of single-cycle terahertz pulses for particle acceleration,” Nat. Commun. 8(1), 421 (2017). [CrossRef]  

3. D.-F. Zhang, A. Fallahi, M. Hemmer, X.-J. Wu, M. Fakhari, Y. Hua, H. Cankaya, and A.-L. Calendron, “Segmented terahertz electron accelerator and manipulator (STEAM),” Nat. Photonics 12(6), 336–342 (2018). [CrossRef]  

4. S. Fleischer, Y. Zhou, R. W. Field, and K. Nelson, “Molecular orientation and alignment by intense single-cycle THz pulses,” Phys. Rev. Lett. 107(16), 163603 (2011). [CrossRef]  

5. M. Sajadi, M. Wolf, and T. Kampfrath, “Transient birefringence of liquids induced by terahertz electric-field torque on permanent molecular dipoles,” Nat. Commun. 8(1), 14963 (2017). [CrossRef]  

6. P. Zalden, L. Song, X. Wu, H. Huang, F. Ahr, O. D. Mücke, J. Reichert, M. Thorwart, P. K. Mishra, R. Welsch, R. Santra, F. X. Kärtner, and C. Bressler, “Molecular polarizability anisotropy of liquid water revealed by terahertz-induced transient orientation,” Nat. Commun. 9(1), 2142 (2018). [CrossRef]  

7. T. Kampfrath, K. Tanaka, and K. A. Nelson, “Resonant and nonresonant control over matter and light by intense terahertz transients,” Nat. Photonics 7(9), 680–690 (2013). [CrossRef]  

8. S. Baierl, M. Hohenleutner, T. Kampfrath, A. K. Zvezdin, A. V. Kimel, R. Huber, and R. V. Mikhaylovskiy, “Nonlinear spin control by terahertz-driven anisotropy fields,” Nat. Photonics 10(11), 715–718 (2016). [CrossRef]  

9. I. Grguraš, A. R. Maier, C. Behrens, T. Mazza, T. J. Kelly, P. Radcliffe, S. Düsterer, A. K. Kazansky, N. M. Kabachnik, T. Tschentscher, J. T. Costello, M. Meyer, M. C. Hoffmann, H. Schlarb, and A. L. Cavalieri, “Ultrafast X-ray pulse characterization at free-electron lasers,” Nat. Photonics 6(12), 852–857 (2012). [CrossRef]  

10. L. Zhao, Z. Wang, C. Lu, R. Wang, C. Hu, P. Wang, J. Qi, T. Jiang, S. Liu, Z. Ma, F. Qi, P. Zhu, Y. Cheng, Z. Shi, Y. Shi, W. Song, X. Zhu, J. Shi, Y. Wang, L. Yan, L. Zhu, D. Xiang, and J. Zhang, “Terahertz streaking of few-femtosecond relativistic electron beams,” Phys. Rev. X 8(2), 021061 (2018). [CrossRef]  

11. J. Hebling, G. Almási, I. Z. Kozma, and J. Kuhl, “Velocity matching by pulse front tilting for large-area THz-pulse generation,” Opt. Express 10(21), 1161–1166 (2002). [CrossRef]  

12. H. Hirori, A. Doi, F. Blanchard, and K. Tanaka, “Single-cycle terahertz pulses with amplitudes exceeding 1 MV/cm generated by optical rectification in LN,” Appl. Phys. Lett. 98(9), 091106 (2011). [CrossRef]  

13. S. B. Bodrov, A. A. Murzanev, Y. A. Sergeev, Y. A. Malkov, and A. N. Stepanov, “Terahertz generation by tilted-front laser pulses in weakly and strongly nonlinear regimes,” Appl. Phys. Lett. 103(25), 251103 (2013). [CrossRef]  

14. S.-W. Huang, E. Granados, W. R. Huang, K.-H. Hong, L. E. Zapata, and F. X. Kärtner, “High conversion efficiency, high energy terahertz pulses by optical rectification in cryogenically cooled lithium niobate,” Opt. Lett. 38(5), 796–798 (2013). [CrossRef]  

15. J. A. Fülöp, Z. Ollmann, C. Lombosi, C. Skrobol, S. Klingebiel, L. Pálfalvi, F. Krausz, S. Karsch, and J. Hebling, “Efficient generation of THz pulses with 0.4 mJ energy,” Opt. Express 22(17), 20155–20163 (2014). [CrossRef]  

16. G. A. Askar’yan, “Cerenkov radiation and transition radiation from electromagnetic waves,” Sov. Phys. JETP 15, 943–946 (1962).

17. D. A. Bagdasarian, A. O. Makarian, and P. S. Pogosian, “Cerenkov radiation from a propagating nonlinear polarization wave,” JEPT Lett. 37, 594–596 (1983).

18. D. H. Auston, K. P. Cheung, J. A. Valdmanis, and D. A. Kleinman, “Cherenkov radiation from femtosecond optical pulses in electro-optic media,” Phys. Rev. Lett. 53(16), 1555–1558 (1984). [CrossRef]  

19. A. G. Stepanov, J. Hebling, and J. Kuhl, “THz generation via optical rectification with ultrashort laser pulse focused to a line,” Appl. Phys. B: Lasers Opt. 81(1), 23–26 (2005). [CrossRef]  

20. M. Theuer, G. Torosyan, C. Rau, R. Beigang, K. Maki, C. Otani, and K. Kawase, “Efficient generation of Cherenkov-type terahertz radiation from a lithium niobate crystal with a silicon prism output coupler,” Appl. Phys. Lett. 88(7), 071122 (2006). [CrossRef]  

21. S. B. Bodrov, M. I. Bakunov, and M. Hangyo, “Efficient Cherenkov emission of broadband terahertz radiation from an ultrashort laser pulse in a sandwich structure with nonlinear core,” J. Appl. Phys. 104(9), 093105 (2008). [CrossRef]  

22. S. B. Bodrov, A. N. Stepanov, M. I. Bakunov, B. V. Shishkin, I. E. Ilyakov, and R. A. Akhmedzhanov, “Highly efficient optical-to-terahertz conversion in a sandwich structure with LN core,” Opt. Express 17(3), 1871–1879 (2009). [CrossRef]  

23. K. Suizu, K. Koketsu, T. Shibuya, T. Tsutsui, T. Akiba, and K. Kawase, “Extremely frequency-widened terahertz wave generation using Cherenkov-type radiation,” Opt. Express 17(8), 6676–6681 (2009). [CrossRef]  

24. M. I. Bakunov and S. B. Bodrov, “Si-LN-air-metal structure for concentrated terahertz emission from ultrashort laser pulses,” Appl. Phys. B: Lasers Opt. 98(1), 1–4 (2010). [CrossRef]  

25. S. B. Bodrov, I. E. Ilyakov, B. V. Shishkin, and A. N. Stepanov, “Efficient terahertz generation by optical rectification in Si-LN-air-metal sandwich structure with variable air gap,” Appl. Phys. Lett. 100(20), 201114 (2012). [CrossRef]  

26. M. I. Bakunov, E. A. Mashkovich, M. V. Tsarev, and S. D. Gorelov, “Efficient Cherenkov-type terahertz generation in Si-prism-LN-slab structure pumped by nanojoule-level ultrashort laser pulses,” Appl. Phys. Lett. 101(15), 151102 (2012). [CrossRef]  

27. S. Fan, H. Takeuchi, T. Ouchi, K. Takeya, and K. Kawase, “Broadband terahertz wave generation from a MgO:LN ridge waveguide pumped by a 1.5 $\mu$m femtosecond fiber laser,” Opt. Lett. 38(10), 1654–1656 (2013). [CrossRef]  

28. K. Takeya, T. Minami, H. Okano, S. R. Tripathi, and K. Kawase, “Enhanced Cherenkov phase matching terahertz wave generation via a magnesium oxide doped lithium niobate ridged waveguide crystal,” APL Photonics 2(1), 016102 (2017). [CrossRef]  

29. J. A. Fülöp, L. Pálfalvi, G. Almási, and J. Hebling, “Design of high-energy terahertz sources based on optical rectification,” Opt. Express 18(12), 12311–12327 (2010). [CrossRef]  

30. M. I. Bakunov, S. B. Bodrov, and E. A. Mashkovich, “Terahertz generation with tilted-front laser pulses: dynamic theory for low-absorbing crystals,” J. Opt. Soc. Am. B 28(7), 1724–1734 (2011). [CrossRef]  

31. M. I. Bakunov and S. B. Bodrov, “Terahertz generation with tilted-front laser pulses in a contact-grating scheme,” J. Opt. Soc. Am. B 31(11), 2549–2557 (2014). [CrossRef]  

32. F. Blanchard, B. E. Schmidt, X. Ropagnol, N. Thire, T. Ozaki, R. Morandotti, D. G. Cooke, and F. Legare, “Terahertz pulse generation from bulk GaAs by a tilted-pulse-front excitation at 1.8 $\mu$m,” Appl. Phys. Lett. 105(24), 241106 (2014). [CrossRef]  

33. J. A. Fülöp, G. Polónyi, B. Monoszai, G. Andriukatis, T. Balciunas, A. Pugzlys, G. Arthur, A. Baltuska, and J. Hebling, “Highly efficient scalable monolithic semiconductor terahertz pulse source,” Optica 3(10), 1075–1078 (2016). [CrossRef]  

34. G. Polónyi, B. Monoszlai, G. Gäumann, E. J. Rohwer, G. Andriukaitis, T. Balciunas, A. Pugzlys, A. Baltuska, T. Feurer, J. Hebling, and J. A. Fülöp, “High-energy terahertz pulses from semiconductors pumped beyond the three-photon absorption edge,” Opt. Express 24(21), 23872–23882 (2016). [CrossRef]  

35. O. Gayer, Z. Sacks, E. Galun, and A. Arie, “Temperature and wavelength dependent refractive index equations for MgO-doped congruent and stoichiometric LiNbO$_3$,” Appl. Phys. B: Lasers Opt. 91(2), 343–348 (2008). [CrossRef]  

36. D. Grischkowsky, S. Keiding, M. van Exter, and C. Fattinger, “Far-infrared time-domain spectroscopy with terahertz beams of dielectrics and semiconductors,” J. Opt. Soc. Am. B 7(10), 2006–2015 (1990). [CrossRef]  

37. Q. Meng, B. Zhang, S. Zhong, and L. Zhu, “Damage threshold of lithium niobate crystal under single and multiple femtosecond laser pulses: theoretical and experimental study,” Appl. Phys. A 122(6), 582 (2016). [CrossRef]  

38. Z. Su, Q. Meng, and B. Zhang, “Analysis on the damage threshold of MgO:LiNbO$_3$ crystals under multiple femtosecond laser pulses,” Opt. Mater. 60, 443–449 (2016). [CrossRef]  

39. M. Abarkan, J. P. Salvestrini, M. D. Fontana, and M. Aillerie, “Frequency and wavelength dependences of electro-optic coefficients in inorganic crystals,” Appl. Phys. B: Lasers Opt. 76(7), 765–769 (2003). [CrossRef]  

40. A. Riefer, S. Sanna, A. V. Gavrilenko, and W. G. Schmidt, “Linear and nonlinear optical response of LiNbO$_3$ calculated from first principles,” IEEE Trans. Ultrason., Ferroelect., Freq. Contr. 59(9), 1929–1933 (2012). [CrossRef]  

41. R. Schiek and T. Pertsch, “Absolute measurement of the quadratic nonlinear susceptibility of lithium niobate in waveguides,” Opt. Mater. Express 2(2), 126–139 (2012). [CrossRef]  

42. M. I. Bakunov, S. B. Bodrov, A. V. Maslov, and M. Hangyo, “Theory of terahertz generation in a slab of electro-optic material using an ultrashort laser pulse focused to a line,” Phys. Rev. B 76(8), 085346 (2007). [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 (4)

Fig. 1.
Fig. 1. Schematics of the experimental setup.
Fig. 2.
Fig. 2. Conversion efficiency as a function of the pump pulse energy (per unit length of the focal line) for different pump wavelengths. The star points show the efficiency when using a metal substrate. The upper scale shows the incident pump intensity.
Fig. 3.
Fig. 3. (a) Conversion efficiency and transmitted optical pulse energy as functions of the pump pulse energy (per unit length of the focal line) for 1500 nm. (b)–(e) Images of the transmitted laser beam profile at different pump pulse energies.
Fig. 4.
Fig. 4. (a) Terahertz energy spectra (normalized to unity) for different pump wavelengths at the saturation pump energies. (b) Terahertz energy spectra for 1500-nm, 200 $\mu$J/cm pump with and without metal substrate. (c) Terahertz energy spectra for 1300-nm pump at low and high pump energies.
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.