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

Quasi-phase matched second harmonic generation in periodically poled Rb-doped KTiOPO4 ridge waveguide

Open Access Open Access

Abstract

A 10.8 µm wide ridge waveguide was fabricated by diamond-blade dicing in an ion-exchanged periodically poled Rb-doped KTiOPO4 sample. The waveguide was used to generate blue second harmonic light at 468.8 nm in the TM00 mode through first order Type I quasi-phase matching, exploiting the large d33 coefficient of the crystal. It was evaluated using a cw Ti:Sapphire laser, and 6.7 µW of blue light was generated with 5.8 mW of fundamental radiation at 933.8 nm coupled through the waveguide. The results presented here pave the way for efficient nonlinear processes in a waveguide format.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Integrated optics [1] provide means for efficient nonlinear interactions thanks to the ability to provide high confinement over long interaction lengths, and it has been a field of intense research since the early days of nonlinear optics [1]. With quasi-phase matching (QPM), χ2 and χ3 interactions can be tailored for specific interactions [2]. Practical QPM waveguides were first realised in LiNbO3 [3,4] and KTiOPO4 (KTP) by diffusion techniques [5,6]. At the time, the field was driven by the need of a blue laser for reading of optical discs, and power levels in the blue exceeding 100 mW were obtained with KTP waveguides [7], but after the invention of the blue laser diode [8] this technology was essentially discarded. By the development of electric field poling [9] and novel waveguide fabrication methods, waveguide devices have become of interest again and exploited in several contemporary applications like spectroscopy, metrology sensing and quantum optics [10–13].

To reach the higher conversion efficiencies, which is necessary for low power applications like quantum communication [11,12], the modal cross section needs to be small, the overlap high and the losses low. Most of the work in this direction has exploited various LiNbO3 waveguides, either based on proton exchange [14] or adhered waveguides in which a ridge structure has been diced out [15], or nanophotonic waveguides [16]. KTP is a potential alternative to LiNbO3 for waveguide fabrication, as it has high nonlinear coefficients, strong resistance to optical damage and a wide transparency [17]. Ion exchange [18], ion implantation [19], diffusion bonding [20] and laser writing [21] have all been successful methods for realization of waveguides in KTP. Recently, Chen et al. [22] and Volk et al. [23] demonstrated that it was possible to obtain low loss ridge waveguides in ion-implanted and ion-exchanged KTP by precise diamond blade dicing. These waveguides were used to demonstrate Type II second harmonic generation (SHG) birefringent phase matching with a normalized efficiency of 0.17%/Wcm2 to the green [23]. To further increase the efficiency and flexibility in the nonlinear interaction, one should try to combine this technology with periodically poled structures to obtain QPM of Type I for KTP so the strong d33 coefficient could be exploited [24].

In this work, we have fabricated ridge waveguides by precise diamond dicing of a planar Rb-ion exchanged waveguide in a periodically poled Rb-doped KTP (PPRKTP) sample and used it to generate blue light at 468.8 nm through first order (m = 1) Type I QPM SHG. To the best of our knowledge, this work represents the first demonstration of a waveguide in Rb-doped KTP. We achieved a conversion efficiency of 0.12% with a cw Ti:Sapphire laser, with 5.8 mW of fundamental radiation measured at the waveguide output.

2. Waveguide fabrication and characterization

Bulk Rb-doped KTP (RKTP) was chosen as the crystal material as it can be periodically poled with high quality, and the dense gratings necessary for advanced applications [25–28]. 10×5×1 mm samples (x,y,z) of 0.3% Rb-doped KTP were first periodically poled with a period of Λ = 5.82 µm over a 8 × 4 mm area by the standard electric-field poling technique [25]. Afterwards, the sample was selectively etched to reveal the domain structure. It was then tested for first order (m = 1) QPM SHG using a cw Ti-Sapphire laser, focused in the crystal with optimum focusing parameters [29]. The sample was found to be well-poled with a normalized conversion efficiency of 1.8%/Wcm at a wavelength of 933.8 nm at room temperature. This value is in close agreement with what is expected from a first order QPM grating with 50% duty-cycle.

Waveguides were fabricated by immersing the samples in RbNO3 for 2.5 hrs at 345°C degrees, in order to produce a planar waveguide that propagates a single mode at 930 nm. The Rb+-ions diffuse into the crystal through the z-faces, replace outdiffusing K+- ions during the ion-exchange, forming a Rb-concentration and refractive index profile following an erfc(z/d) where z is the depth coordinate and d is the characteristic diffusion depth [18]. The crystal was then selectively etched again, and we confirmed no alteration in the domain structure due to the ion-exchange process. Subsequently, we characterize the planar waveguide at blue (457 nm), red (633 nm) and near-IR (853 nm) wavelengths by prism coupling. We observed two modes in the blue, while the waveguide guided one mode only in the red and the near-IR. The ion-exchange in RKTP was slower and the refractive index change considerably lower than what is obtained for regular flux-grown, undoped KTP for the same exchange conditions. It is attributed to the fact that the crystal already contains a small amount of Rb in the relatively open channels that these crystals consist of in the z-direction. These ions partly block additional diffusion and slow down the process [17,30].

Several ridge waveguides of width varying from 8 to 12 µm were then fabricated on the polar face, i.e, z-face, of the sample by diamond blade dicing using a dicing saw DFD 640, DISCO Corp., and blade P1A863 SD6000N100BR50. The dicing parameters were 30000 rpm and 0.25 mm/s. An example of the diced waveguide is seen in Fig. 1, where Fig. 1(a) shows one of the end faces and Fig. 1(b) shows the top view. Note that the domain structure can be clearly seen in Fig. 1(b) with a duty-cycle close to 50%. Compared to waveguides diced in LiNbO3 we found that RKTP is slightly more sensitive and more care must be taken to avoid chipping and/or cracking of the sample. We believe that this has to do with difference in crystal structure and the stress induced by ion exchange for the KTP family. While Rb-exchanged channel waveguides [18], as well as diamond diced Rb-waveguides [23], can have very low loss, our waveguides typically transmitted from 5% to 20% of the incident light. Indeed, when light was launched in the waveguide substantial scattering could be seen when looking from atop with a microscope. Furthermore, one can see slight chipping on the polar faces along the waveguide, which should be the main reason for scattering, see Fig. 1(b). We hence believe that further optimization of the dicing process should be possible.

 figure: Fig. 1

Fig. 1 (a) Photo of the end-face of a 10.8 µm diced waveguide and (b) top picture of ridge waveguide where the domain structure as well as the surface roughness can be seen. (c) Corresponding near field distribution of the TM00 mode measured at 937 nm. Simulated intensity distribution at (d) the fundamental and (e) SH wavelengths.

Download Full Size | PDF

The waveguides were then evaluated for SHG by launching the Ti-sapphire laser using an objective lens with 10X magnification and 0.25 numerical aperture. The measured fundamental TM00 mode can be seen in Fig. 1(c). Simulations of the waveguide mode profile were carried out using COMSOL, where we assumed an erfc-function for the refractive index profile. Figure 1(d) shows the simulated fundamental TM00 mode and Fig. 1(e) shows the SHG TM00 mode profile. From the overlap between the focused Gaussian input beam and the simulated TM00 waveguide mode, we calculate the coupling efficiency to be 84%. Three distinct SH modes were generated, TM00, TM10 and TM02, when tuning the fundamental wavelength from 930 nm to 940 nm, see Fig. 2. The near field photographs of the generated blue modes are shown together with their respective peaks. The highest blue power, 6.7 µW, was obtained for the TM00 mode with 5.8 mW of fundamental power at 937.3 nm measured at the waveguide output. This corresponds to a conversion efficiency of 0.12%. To compare it to other nonlinear waveguide experiments it is common to calculate a normalized conversion efficiency:

ηexperimental=P2ωPω2L2100%=31%/Wcm2,
where Pω and P2ω are the power at the fundamental and SH wavelengths, respectively, and L is the length of the poled structure, in cm. Since the fabricated waveguides show non negligible scattering along their length, we have considered the fundamental power at the output of the waveguide for the calculation displayed in Eq. (1). Therefore, the calculated efficiency corresponds to an upper limit. For a more accurate calculation, further investigation is needed in order to characterize the experimental coupling efficiency as well as the propagation losses.

 figure: Fig. 2

Fig. 2 Wavelength scan showing the SH phase matching for the TM00, TM10 and the TM02 modes, where the SHG power has been normalized to unity of the highest peak. Their intensity profiles are shown as insets.

Download Full Size | PDF

The overlap area between the simulated modes at the fundamental and SH wavelengths, displayed in Fig. 1(d) and 1(e), is Aovl 50 µm2. Considering a lossless waveguide, the normalized conversion efficiency would then be [31]:

ηtheoretical=8π2deff2Nω2N2ωc0ε0λω2Aovl=102%/Wcm2,
where Nω and N2ω are the modal refractive indices at the fundamental and SH wavelengths, respectively, λω is the fundamental wavelength, c0is the speed of light in vacuum, ε0is the vacuum permittivity, deffis the effective second order nonlinear coefficient, given by 2d33/π for a first order grating and Type I QPM. The phase-matching curve was approximately 4 times wider than what is expected theoretically, and what we measured for the bulk. We attribute this to a variation in modal refractive index, probably due to small width variations along the waveguide, which is known to broaden the bandwidth and reduce the efficiency [31,32]. Both the width variation along the waveguide length and the chipping refers to a still unoptimized dicing process.

To identify the phase matching peaks, we have plotted the waveguide dispersion for the fundamental and SH wavelengths together with the corresponding modes for the waveguide in Fig. 3. Here the modal refractive index for the blue has been adjusted with the QPM condition:

N2ω=Nω+λω2Λ
where Nωand N2ω are the modal refractive indices at the fundamental and SH wavelengths, respectively, and Λ is the poling period. The phase matching wavelength peaks appear in close agreement with what was calculated using the refractive index data obtained from prism coupling, see Fig. 3. In this figure it is also displayed, with dashed lines, the refractive index curve for bulk, calculated using Sellmeier equations [33]. For this calculation, we used the measured refractive indices for the planar waveguide modes and taking into account that for the ridge waveguides the indices should be smaller, their values were fitted to the measured values of phase matching wavelengths. Using Sellmeier equations [33] the poling period was designed for phase matching the fundamental wavelength of 933.5 nm, and we observe that it occurred at 933.8 nm.

 figure: Fig. 3

Fig. 3 Calculated dispersion for the fundamental (in red) and SH (in blue) modes. The dotted lines represent the bulk refractive indices and the vertical lines mark the phase matching wavelengths from right to left, the TM00 mode in IR to TM00 and TM10, as well as dotted the bulk phase matching.

Download Full Size | PDF

3. Summary

In conclusion, we have fabricated a 10.8 µm wide diamond-blade diced ridge waveguide in Rb-exchanged PPRKTP with a period of 5.82 µm and generated blue SH light at 468.8 nm in the TM00 mode. To the best of our knowledge this is the first Type I QPM diamond-diced waveguide SHG in a crystal from the KTP family. As these crystals can be poled with much shorter periods, one should be able to study several novel processes, as mirrorless optical parametric oscillators [34] and entangled sources with signal and idler photons in the opposite direction in such waveguides. Furthermore, by improving the dicing process the chipping could be eliminated and the loss thereby much reduced. With a more homogeneous waveguide the phase-matching bandwidth would be narrowed, further improving the efficiency.

Funding

Knut and Alice Wallenberg Foundation and the Swedish Research Council.

References

1. G. I. Stegeman and C. T. Seaton, “Nonlinear integrated optics,” J. Appl. Phys. 58(12), R57–R78 (1985). [CrossRef]  

2. J. A. Armstrong, N. Bloembergen, J. Ducuing, and P. S. Pershan, “Interactions between light waves in a nonlinear dielectric,” Phys. Rev. 127(6), 1918–1939 (1962). [CrossRef]  

3. J. Webjörn, F. Laurell, and G. Arvidsson, “Fabrication of periodically domain-inverted lithium niobate channel waveguides for second harmonic generation,” J. Lightwave Technol. 7(19), 1597–1600 (1989). [CrossRef]  

4. E. J. Lim, M. M. Fejer, and R. L. Byer, “Second harmonic generation of green light in a periodically-poled planar lithium niobate waveguide,” Electron. Lett. 25(3), 174–175 (1989). [CrossRef]  

5. C. J. van der Poel, J. D. Bierlein, J. B. Brown, and S. Colak, “Efficient type I blue second‐harmonic generation in periodically segmented KTiOPO4 waveguides,” Appl. Phys. Lett. 57(20), 2074–2076 (1990). [CrossRef]  

6. F. Laurell, J. B. Brown, and J. D. Bierlein, “Simultaneous generation of UV and visible light in segmented KTP waveguides,” Appl. Phys. Lett. 62(16), 1872–1874 (1993). [CrossRef]  

7. F. Laurell, “Periodically poled materials for miniature light sources,” Opt. Mater. 11(2-3), 235–244 (1999). [CrossRef]  

8. S. Nakamura, M. Senoh, S. Nagahama, N. Iwasa, T. Yamada, T. Matsushita, H. Kiyoku, Y. Sugimoto, T. Kozaki, H. Umemoto, M. Sano, and K. Chocho, “InGaN/GaN/AlGaN-Based Laser Diodes with Modulation-Doped Strained-Layer Superlattices,” Jpn. J. Appl. Phys. 36(2), L1568–L1571 (1997). [CrossRef]  

9. M. Yamada, N. Nada, M. Saitoh, and K. Watanabe, “First-order quasiphase matched LiNbO3 waveguide periodically poled by applying an external field for efficient blue second-harmonic generation,” Appl. Phys. Lett. 62(5), 435–436 (1993). [CrossRef]  

10. M. Jansson, J. Roeraade, and F. Laurell, “Laser-induced fluorescence detection in capillary electrophoresis with blue light from a frequency-doubled diode laser,” Anal. Chem. 65(20), 2766–2769 (1993). [CrossRef]  

11. M. Pelton, P. Marsden, D. Ljunggren, M. Tengner, A. Karlsson, A. Fragemann, C. Canalias, and F. Laurell, “Bright, single-spatial-mode source of frequency non-degenerate, polarization-entangled photon pairs using periodically poled KTP,” Opt. Express 12(15), 3573–3580 (2004). [CrossRef]   [PubMed]  

12. S. Tanzilli, W. Tittel, H. De Riedmatten, H. Zbinden, P. Baldi, M. De Micheli, D. B. Ostrowsky, and N. Gisin, “PPLN waveguide for quantum communication,” Eur. Phys. J. D 18(2), 155–160 (2002). [CrossRef]  

13. L. Yu, C. M. Natarajan, T. Horikiri, C. Langrock, J. S. Pelc, M. G. Tanner, E. Abe, S. Maier, C. Schneider, S. Höfling, M. Kamp, R. H. Hadfield, M. M. Fejer, and Y. Yamamoto, “Two-photon interference at telecom wavelengths for time-bin-encoded single photons from quantum-dot spin qubits,” Nat. Commun. 6(1), 8955 (2015). [CrossRef]   [PubMed]  

14. K. R. Parameswaran, R. K. Route, J. R. Kurz, R. V. Roussev, M. M. Fejer, and M. Fujimura, “Highly efficient second-harmonic generation in buried waveguides formed by annealed and reverse proton exchange in periodically poled lithium niobate,” Opt. Lett. 27(3), 179–181 (2002). [CrossRef]   [PubMed]  

15. S. Kurimura, Y. Kato, M. Maruyama, Y. Usui, and H. Nakajima, “Quasi-phase-matched adhered ridge waveguide in LiNbO3,” Appl. Phys. Lett. 89(19), 191123 (2006). [CrossRef]  

16. R. Luo, Y. He, H. Liang, M. Li, and Q. Lin, “Highly tunable efficient second-harmonic generation in a lithium niobate nanophotonic waveguide,” Optica 5(8), 1006–1011 (2018). [CrossRef]  

17. J. D. Bierlein and H. Vanherzeele, “Potassium titanyl phosphate: properties and new applications,” J. Opt. Soc. Am. B 6(4), 622–633 (1989). [CrossRef]  

18. J. D. Bierlein, A. Ferretti, L. H. Brixner, and W. Y. Hsu, “Fabrication and characterization of optical waveguides in KTiOPO4,” Appl. Phys. Lett. 50(18), 1216–1218 (1987). [CrossRef]  

19. P. Bindner, A. Boudrioua, J. C. Loulergue, and P. Moretti, “Formation of planar optical waveguides in potassium titanyl phosphate by double implantation of protons,” Appl. Phys. Lett. 79(16), 2558–2560 (2001). [CrossRef]  

20. V. Boutou, A. Vernay, C. Félix, F. Bassignot, M. Chauvet, D. Lupinski, and B. Boulanger, “Phase-matched second-harmonic generation in a flux grown KTP crystal ridge optical waveguide,” Opt. Lett. 43(15), 3770–3773 (2018). [CrossRef]   [PubMed]  

21. S. Campbell, R. R. Thomson, D. P. Hand, A. K. Kar, D. T. Reid, C. Canalias, V. Pasiskevicius, and F. Laurell, “Frequency-doubling in femtosecond laser inscribed periodically-poled potassium titanyl phosphate waveguides,” Opt. Express 15(25), 17146–17150 (2007). [CrossRef]   [PubMed]  

22. C. Chen, C. E. Rüter, M. F. Volk, C. Chen, Z. Shang, Q. Lu, S. Akhmadaliev, S. Zhou, F. Chen, and D. Kip, “Second harmonic generation of diamond-blade diced KTiOPO4 ridge waveguides,” Opt. Express 24(15), 16434–16439 (2016). [CrossRef]   [PubMed]  

23. M. F. Volk, C. E. Rüter, M. Santandrea, C. Eigner, L. Padberg, H. Herrmann, C. Silberhorn, and D. Kip, “Fabrication of low-loss Rb-exchanged ridge waveguides in z-cut KTiOPO4,” Opt. Mater. Express 8(1), 82–87 (2018). [CrossRef]  

24. H. Karlsson and F. Laurell, “Electric field poling of flux grown KTiOPO4,” Appl. Phys. Lett. 71(24), 3474–3476 (1997). [CrossRef]  

25. Q. Jiang, P. A. Thomas, K. B. Hutton, and R. C. C. Ward, “Rb-doped potassium titanyl phosphate for periodic ferroelectric domain inversion,” J. Appl. Phys. 92(5), 2717–2723 (2002). [CrossRef]  

26. S. Wang, V. Pasiskevicius, and F. Laurell, “High-efficiency frequency converters with periodically-poled Rb-doped KTiOPO4,” Opt. Mater. 30(4), 594–599 (2007). [CrossRef]  

27. A. Zukauskas, N. Thilmann, V. Pasiskevicius, F. Laurell, and C. Canalias, “5 mm thick periodically poled Rb-doped KTP for high energy optical parametric frequency conversion,” Opt. Mater. Express 1(2), 201–206 (2011). [CrossRef]  

28. A. Zukauskas, G. Strömqvist, V. Pasiskevicius, F. Laurell, M. Fokine, and C. Canalias, “Fabrication of submicrometer quasi-phase-matched devices in KTP and RKTP,” Opt. Mater. Express 1(7), 1319–1325 (2011). [CrossRef]  

29. G. D. Boyd and D. A. Kleinman, “Parametric interactions of focused Gaussian light beams,” J. Appl. Phys. 39(8), 3597–3639 (1968). [CrossRef]  

30. M. G. Roelofs, P. A. Morris, and J. D. Bierlein, “Ion exchange of Rb, Ba, and Sr in KTiOPO4,” J. Appl. Phys. 70(2), 720–728 (1991). [CrossRef]  

31. F. Laurell and G. Arvidsson, “Frequency doubling in Ti:MgO:LiNbO3 channel waveguides,” J. Opt. Soc. Am. B 5(2), 292–299 (1988). [CrossRef]  

32. M. M. Fejer, G. A. Magel, D. H. Jundt, and R. L. Byer, “Quasi-phase-matched second harmonic generation: tuning and tolerances,” IEEE J. Quantum Electron. 28(11), 2631–2654 (1992). [CrossRef]  

33. T. Y. Fan, C. E. Huang, B. Q. Hu, R. C. Eckardt, Y. X. Fan, R. L. Byer, and R. S. Feigelson, “Second harmonic generation and accurate index of refraction measurements in flux-grown KTiOPO(4),” Appl. Opt. 26(12), 2390–2394 (1987). [CrossRef]   [PubMed]  

34. C. Canalias and V. Pasiskevicius, “Mirrorless optical parametric oscillator,” Nat. Photonics 1(8), 459–462 (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 (3)

Fig. 1
Fig. 1 (a) Photo of the end-face of a 10.8 µm diced waveguide and (b) top picture of ridge waveguide where the domain structure as well as the surface roughness can be seen. (c) Corresponding near field distribution of the TM00 mode measured at 937 nm. Simulated intensity distribution at (d) the fundamental and (e) SH wavelengths.
Fig. 2
Fig. 2 Wavelength scan showing the SH phase matching for the TM00, TM10 and the TM02 modes, where the SHG power has been normalized to unity of the highest peak. Their intensity profiles are shown as insets.
Fig. 3
Fig. 3 Calculated dispersion for the fundamental (in red) and SH (in blue) modes. The dotted lines represent the bulk refractive indices and the vertical lines mark the phase matching wavelengths from right to left, the TM00 mode in IR to TM00 and TM10, as well as dotted the bulk phase matching.

Equations (3)

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

η experimental = P 2ω P ω 2 L 2 100%=31%/Wc m 2 ,
η theoretical = 8 π 2 d eff 2 N ω 2 N 2ω c 0 ε 0 λ ω 2 A ovl =102%/Wc m 2 ,
N 2ω = N ω + λ ω 2Λ
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.