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

Ion implanted optical channel waveguides in Er3+/MgO co-doped near stoichiometric LiNbO3: a new candidate for active integrated photonic devices operating at 1.5 µm

Open Access Open Access

Abstract

We report on the fabrication of active optical channel waveguides in Er3+/MgO co-doped near stoichiometric lithium niobate crystals by means of selective low-dose oxygen ion implantation through a specially designed photoresist stripe mask. After post-implantation treatment at 260°C for 1 h, the channel waveguides possess a propagation loss of ~1.7 dB/cm. The micro-luminescence investigation reveals that fluorescence emissions at ~1.5 µm in the waveguides are well preserved with respect to the bulk, exhibiting possible applications for integrated active photonic devices.

©2008 Optical Society of America

1. Introduction

Erbium (Er) doped optical materials are of great interest to both researchers and industries for the all-optical information systems, due to the intra-4f emissions of Er3+ ions at first telecommunication wavelength of ~1.5 µm which allow to implement active photonic device [13]. To this purpose, recent results have shown that Er doped lithium niobate (LiNbO3) crystals are promising candidates for their proper combination of laser generation and excellent electrooptic and nonlinear properties of the substrate crystals [4-6]. The mostly used commercial LiNbO3 are congruent crystals (usually known as CLN), i.e., the ratio of Li and Nb ions ([Li]/[Li+Nb]≈48.5%) in the crystals is less than the stoichiometric composition. This Li-deficient structure causes a large number of intrinsic defects, which degrades, to a certain extent, the optical performances of the crystals [7]. More recently, it has been proved that, by increasing the ratio of [Li]/[Li+Nb] to a nearly stoichiometric value (>49.5%), both the electrooptic and the nonlinear properties could be considerably improved in near stoichiometric LiNbO3 (hereafter abbreviated to SLN), compared with their congruent partners [68]. Particularly, it has also been confirmed that the Er3+ doped SLN (or co-doped with MgO, namely Er:MgO:SLN) is a promising active gain medium with a superior laser capability in both visible (VIS) and near infrared (NIR) band range over the Er3+:CLN due to the reduction of the density of non-stoichiometric defects inside the crystal [911].

Besides the choice of the optical material, a second key-factor is the adoption of a guided-wave configuration. Actually, as one of the most basic components of modern photonic networks, optical channel waveguides can confine the light propagation in the two transverse dimensions to order of micrometers, resulting in many enhanced performances for various optical applications [12]. Channel waveguides in CLN wafers have been produced by several methods for active integrated devices, such as titanium [13] or zinc [14] ion indiffusion, proton exchange [1518], ion implantation [1922], laser writing [23,24], etc. Among these techniques, the implantation of both light (e.g. H or He) [20,23,2628] and heavy ions (O or C) [20,21,28,29] (mainly depending on a physical mechanism) has been used to fabricate waveguides in large number of optical materials. The technique exhibits a wide capability of fabricating in versatile substrates, with good reproducibility and controllability, as well as with the possibility of easy combination with other techniques for more complex waveguide designs. For example, reconfigurable optical channel waveguides in LN have been fabricated by combination of O ion implantation and selective white light illumination [30]. It should be pointed out that the conventional chemical methods for CLN waveguide formation are proved to be much less effective for SLN crystals: the indiffusion coefficients of metal ions are much low in SLN because of the reduction of intrinsic defects [31] and the incorporation of protons into SLN replacing original Li ions results in a reduction of [Li]/[Li+Nb] ratio in the waveguides [6,1518] which causes a degradation in the SLN original properties.

In this work, we report, to our knowledge for the first time, the optical channel waveguides formation in Er:MgO:SLN crystal by using selective low-dose oxygen ion implantation. The guiding properties at wavelength of 632.8 nm as well as the luminescence properties at 1.5 µm of the Er ions inside the waveguide are investigated.

2. Experiments and results

The z-cut Er:MgO:SLN sample (doped with 0.2 mol% Er3+ and 1.5 mol% MgO) we use is obtained from Crystalblue Co. Ltd., China, with [Li]/[Li+Nb] ratio of ~49.7%. It is optically polished and cut with size of 7(x)×10(y)×1.2(z) mm3 (z-direction pointing at the crystal c axis). By applying standard lithography technique, one xy face of the sample is covered with specially designed photoresist stripe mask, which consists of a series of open stripes (with width of 10 µm, uncovered) with separation of 40 µm (covered by 5µm-thick photoresist stripe masks) between the adjacent channels. With this masking, the implantation of O+ ions at energy of 3 MeV and dose of 6×1014 cm-2 is performed by a 2×1.7 MV tandem accelerator at Peking University. The implantation will induce refractive index changes in the un-protected surface regions (open stripes), whilst does not affect the photoresist-mask-covered regions. In this way, the channel waveguides can be formed in the implantation-exposed stripe regions: see Fig. 1 for the schematic plot of the channel waveguide formation process. Usually for low-dose O implanted optical waveguides, moderate annealing at 200-300°C for 30 min to several hours is suitable for optimization of the waveguide performance: it eliminates the points defects/color centers induced by the implantation but does not destroy the waveguide structures. After the implantation, the sample is annealed at 260°C for 1h in air to reduce the irradiation-induced color centers for an improvement of the waveguide quality. Moreover, the cross sections of the waveguide sample are imaged by a microscope with reflected polarized light (Olympus BX51M, Japan), by which the channel waveguide configuration is clearly observed (see inset of Fig. 1). The trapezoidal shape of the waveguide cross section is due to the wedged geometry of the photoresist masks at their edges. In addition, for comparison, about 1/4 of the sample surface is not covered by any masks, allowing planar waveguide formation with same conditions to the channel one.

 figure: Fig. 1.

Fig. 1. Schematic plot of the waveguide fabrication process. The inset shows the microscope image of the transverse cross section of the channel waveguide. The dashed lines denote the location of the channel waveguide.

Download Full Size | PDF

 figure: Fig. 2.

Fig. 2. Profile of extraordinary refractive index changes Δn e induced by the ion implantation as function of the penetration depth of the incident ions inside the crystal (pointing z axis).

Download Full Size | PDF

The dark-mode spectroscopy of the planar waveguide in Er:MgO:SLN is measured by the well-known m-line technique at wavelength of 632.8 nm through a prism coupler (Model 2010, Metricon, USA). According to the dark-line spectrum, we reconstruct the profile of the extraordinary refractive index changes (Δn e) of the planar waveguide [Fig. 2, solid line] by applying the reflectivity calculation method (RCM) of the planar waveguide [32]. As one can see, the O+ ion implantation creates a positive index well (Δn w≈+0.016) in the near surface regions, by which the waveguide could confine the light propagation in a non-leaky way. Meanwhile, an optical barrier with negative index change (Δn b≈-0.004) is built at the depth of ~2µm from the surface down to the substrate. Such profile is a typical “well+barrier” distribution, which is similar to those of many ion implanted waveguides [19-22,33,34]. In addition, the peak position of the barrier is in good agreement with the mean project range (ΔR p≈2µm) of incident 3 MeV O+ ions inside Er:MgO:SLN, which means the waveguide dimension could be well determined theoretically. We also estimate the Δn e profile of the planar waveguide at 1.5µm [Fig. 2, dashed line], via numerical fit based on the Sellmeier equation of the crystal [35].

 figure: Fig. 3.

Fig. 3. (a). The 2D n e distribution of the channel waveguide at the cross section, (b). 3D plot of the calculated modal profile of the quasi-TM00 mode by FD-BPM and (c). measured near-field intensity distribution of quasi-TM00 mode at wavelength of 632.8nm.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. (a). The 2D n e distribution of the channel waveguide at the cross section and (b) 3D plot of the calculated modal profile of the quasi-TM00 mode by FD-BPM at wavelength of 1.5µm.

Download Full Size | PDF

Figure 3(a) depicts the estimated the two dimensional (2D) refractive index distribution (at 632.8 nm) of the waveguide cross section, by considering the planar waveguide index map as well as the micro-topography of the channels. With this 2D index profile, we perform a numerical simulation of the light propagation in the channel waveguides through the commercial software BeamPROP [36], which is based on the well-known finite difference beam propagation method (FD-BPM). Figure 3(b) depicts the calculated modal intensity profile of the quasi-transverse magnetic (TM00) mode of the channel waveguides at 632.8 nm. Experimentally, we use an end-face coupling arrangement (similar to Ref. 20, with a polarized He-Ne laser at 632.8 nm) to measure near field intensity distribution of the quasi-TM00 mode [Fig. 3(c)]. By comparing Figs. 3(b) and 3(c), one can conclude that there is a fairly good agreement between the numerical results and the experimental data, which suggests that the estimate of 2D index distribution is good enough for channel waveguide design purpose. We also estimate the 2D n e profile at 1.5 µm [Fig. 4(a)] in the similar way, and obtain the modal distribution of quasi-TM00 mode [Fig. 4(b)] by FD-BPM. As it can be seen, the numerical simulation suggests that the channel waveguide is still guiding at 1.5 µm.

We use the Fabry-Perot resonance method [37] to measure the propagation loss of the channel waveguides by investigating the power-oscillation effects at the output facet when we gradually heat the sample (up to 5°C). The transmission attenuation is determined to be ~1.7 dB/cm at the wavelength of 632.8 nm.

 figure: Fig. 4.

Fig. 4. Confocal room-temperature luminescence emission spectra from (a) the bulk and (b) the channel waveguide volume, corresponding to the 4 I 13/24 I 15/2 transitions

Download Full Size | PDF

Finally, the 1.5-µm wavelength luminescence properties of Er3+ ions inside the channel waveguide are investigated by using a fiber-coupled confocal microscope in back-scattering configuration. An 800-nm wavelength fiber-coupled diode at power of 40 mW is used as the excitation source. The input beam is focused into the sample by a long working-distance achromatic infrared microscope objective (×100) with N.A. of 0.9, which promotes the excitation of Er3+ ions from their fundamental state (4 I 15/2) to the excited state (4 I 9/2). The subsequent emitted luminescence at ~1.5 µm (4 I 13/24 I 15/2) is collected by using the same microscope objective and analysed by an InGaAs detector connected to a lock-in amplifier and attached to a fiber-coupled monochromator (SPEX 500M). Figures 5(a) and (b) show the room-temperature micro-luminescence spectra of 4 I 13/24 I 15/2 transitions of Er3+ ions from the bulk and the channel waveguide, respectively. As one can see, the emission spectra of the bulk and the channel waveguide, which are both centered at ~1.52 µm, exhibit similar characteristics (in terms of peak positions, widths and relative intensity between sub-stark transitions), and therefore it appears to be reasonable that the main emission characteristics of Er3+ ions are well preserved in the waveguides.

3. Conclusions

Optical channel waveguides in Er:MgO:SLN crystal have been successful produced by using selective low-dose oxygen ion implantation with a photoresist stripe mask. The guiding properties are characterized at wavelength of 632.8 nm. Numerical simulation suggests the fabricated waveguides can be useful at telecommunication wavelength of ~1.5 µm. In addition, the luminescence emissions at 1.52 µm of Er3+ ions in the waveguides are not affected by the implantation process, which shows potential applications for the Er:MgO:SLN channel waveguides as active integrated photonic devices, e.g., Er doped waveguide amplifiers (EDWA), and waveguide lasers operating at ~1.5 µm.

Acknowledgments

The authors thank D. Jaque for helpful discussions. This work is supported by National Natural Science Foundation of China (Grant No. 10735070), Spanish Ministerio de Educación y Ciencia (MAT2007-6468), and Universidad Autónoma de Madrid and Comunidad Autonoma de Madrid (Project No. CCG07-UAM/MAT-1861).

References and links

1. A. Polman, “Erbium-implanted thin film photonic materials,” J. Appl. Phys. 82, 1–39 (1997). [CrossRef]  

2. A. J. Kenyon, “Recent developments in rare-earth doped materials for optoelectronics,” Prog. Quantum Electron. 26, 225–284 (2002). [CrossRef]  

3. I. Baumann, S. Bosso, R. Brinkmann, R. Corsini, M. Dinand, A. Greiner, K. Schäfer, J. Söchtig, W. Sohler, H. Suche, and R. Wessel, “Er-doped integrated optical devices in LiNbO3,” IEEE J. Sel. Top. Quantum Electron. 2, 355–366 (1996). [CrossRef]  

4. W.-S. Yang, H.-Y. Lee, and D.-H. Yoon, “Segregation and laser properties of Er/Mg co-doped LiNbO3 single crystal,” J. Cryst. Growth 244, 49–52 (2002). [CrossRef]  

5. B. K. Das, R. Ricken, and W. Sohler, “Integrated optical distributed feedback laser with Ti:Fe:Er:LiNbO3 waveguide,” Appl. Phys. Lett. 82, 1515–1517 (2003). [CrossRef]  

6. L. Arizmendi, “Photonic application of lithium niobate crystal,” Phys. Stat. Solidi. 2, 253–283 (2004).

7. H. Hatano, K. Kitamura, and Y. Liu, “Growth and Photorefractive Properties of Stoichiometric LiNbO3 and LiTaO3,” in Photorefractive Materials and Their Applications 2: Materials, edited by P. G ünter and J.-P. Huignard (Springer, Berlin, 2007) pp. 127–164.

8. T. Fujiwara, M. Takahashi, M. Ohama, A. J. Ikushima, Y. Furukawa, and K. Kitamura, “Comparison of electro-optic effect between stoichiometric and congruent LiNbO3,” Electron. Lett. 35, 499–501 (1999). [CrossRef]  

9. S. M. Lee, T.I. Shin, Y.T. Kim, M. Habu, T. Ito, M. Natori, and D.H. Yoon, “Effect of optical properties of Er2O3 doped stoichiometric LiNbO3 single crystals and co-doped with MgO,” Mater. Sci. Eng. B 105, 34–36 (2003). [CrossRef]  

10. J. Kang, M. Lee, S. Lee, K. Lim, K. Somu, S. Takekawa, and K. Kitamura, “1.5 µm emission characteristics of Er3+-doped stoichiometric LiNbO3,” Appl. Phys. Lett. 85, 4367–4369 (2004). [CrossRef]  

11. D. L. Zhang and E. Y. B. Pun, “Emission characteristics of near-stoichiometric Er/Yb-codoped LiNbO3 crystals,” J. Appl. Phys. 99, 023101 (2006). [CrossRef]  

12. D. Kip, “Photorefractive waveguides in oxide crystals: fabrication, properties, and applications,” Appl. Phys. B 67, 131–150 (1998). [CrossRef]  

13. D. Zhang, G. Ding, and C. Chen, “Parameters Optimization of Ti-diffused Nd:MgO:LiNbO3 Channel Waveguide Lasers,” J. Mod. Opt. 47, 1623 (2000).

14. G. Lifante, E. Cantelar, J. A. Munoz, R. Nevado, J. A. Sanz-Garcia, and F. Cusso, “Zn-diffused LiNbO3:Er3+/Yb3+ as a waveguide laser material,” Opt. Mater. 13, 181–186 (1999). [CrossRef]  

15. E. Lallier, J. P. Pocholle, M. Papuchon, M. de Micheli, M. J. Li, Q. He, D. B. Ostrowsky, C. Grezes-Besset, and E. Pelletier, “Efficient Nd:MgO:LiNbO3 waveguide laser,” Electron. Lett. 26, 927–928, (1990). [CrossRef]  

16. E. Lallier, “Rare-earth-doped glass and LiNbO3 waveguide lasers and optical amplifiers,” Appl. Opt. 31, 5276–5282 (1992). [CrossRef]   [PubMed]  

17. E. Cantelar, G. Lifante, F. Cusso, M. Domenech, A. Busacca, A. Cino, and S. Riva Sanseverino, “Dualpolarization-pump CW laser operation in Nd3+:LiNbO3 channel waveguides fabricated by reverse proton exchange,” Opt. Mater. 30, 1039–1043 (2008). [CrossRef]  

18. E. M. Rodríguez, D. Jaque, E. Cantelar, F. Cussó, G. Lifante, A. C. Busacca, A. Cino, and S. R. Sanseverino, “Time resolved confocal luminescence investigations on reverse proton exchange Nd:LiNbO3 channel waveguides,” Opt. Express 15, 8805–8811 (2007). [CrossRef]   [PubMed]  

19. S. J. Field, D. C. Hanna, D. P. Shepherd, A. C. Tropper, P. J. Chandler, P. D. Townsend, and L. Zhang, “Ion-implanted Nd:MgO:LiNbO3 planar waveguide laser,” Opt. Lett. 16, 481–483 (1991). [CrossRef]   [PubMed]  

20. F. Chen, Y. Tan, L. Wang, Q. M. Lu, and H. J. Ma, “Oxygen ion implanted optical channel waveguides in Nd:MgO:LiNbO3: Fabrication, characterization and simulation,” J. Phys. D 40, 5824–5827 (2007). [CrossRef]  

21. F. Chen, Y. Tan, D. Jaque, L. Wang, X. L. Wang, and K. M. Wang, “Active waveguide in Nd3+:MgO:LiNbO3 crystal produced by low-dose carbon ion implantation,” Appl. Phys. Lett. 92, 021110 (2008). [CrossRef]  

22. P. D. Townsend, P. J. Chandler, and L. Zhang, Optical Effects of Ion Implantation (Cambridge Univ. Press, Cambridge, 1994). [CrossRef]  

23. J. Thomas, M. Heinrich, J. Burghoff, S. Nolte, A. Ancona, and A. Tünnermann, “Femtosecond laser-written quasi-phase-matched waveguides in lithium niobate,” Appl. Phys. Lett. 91, 151108 (2007). [CrossRef]  

24. A. H. Nejadmalayeri and P. R. Herman, “Rapid thermal annealing in high repetition rate ultrafast laser waveguide writing in lithium niobate,” Opt. Express 15, 10842–10854 (2007). [CrossRef]   [PubMed]  

25. H. Ilan, A. Gumennik, G. Perepelitsa, A. Israel, and A. J. Agranat, “Construction of an optical wire imprinted in potassium lithium tantalate niobate by He+ implantation,” Appl. Phys. Lett. 92, 191101 (2008). [CrossRef]  

26. C. Grivas, D. P. Shepherd, R. W. Eason, L. Laversenne, P. Moretti, C. N. Borca, and M. Pollnau, “Roomtemperature continuous-wave operation of Ti:sapphire buried channel-waveguide lasers fabricated via proton implantation,” Opt. Lett. 31, 3450–3452 (2006). [CrossRef]   [PubMed]  

27. M. Szachowicz, P. Moretti, M.-F. Joubert, M. Couchaud, and B. Ferrand, “Fabrication of H+ implanted channel waveguides in Y3Al5O12:Nd, Tm single crystal buried epitaxial layers for infrared to blue upconversion laser systems,” Appl. Phys. Lett. 90, 031113 (2007). [CrossRef]  

28. G. G. Bentini, M. Bianconi, M. Chiarini, L. Correra, C. Sada, P. Mazzoldi, N. Argiolas, M. Bazzan, and R. Guzzi, “Effect of low dose high energy O3+ implantation on refractive index and linear electro-optic properties in X-cut LiNbO3: Planar optical waveguide formation and characterization,” J. Appl. Phys. 92, 6477–6483 (2002). [CrossRef]  

29. J. Olivares, A. García-Navarro, G. García, A. Méndez, F. Agulló-López, A. García-Cabañes, M. Carrascosa, and O. Caballero, “Nonlinear optical waveguides generated in lithium niobate by swift-ion irradiation at ultralow fluences,” Opt. Lett. 32, 2587–2589 (2007). [CrossRef]   [PubMed]  

30. Y. Tan, F. Chen, M. Stepic, V. Shandarov, and D. Kip, “Reconfigurable optical channel waveguides in lithium niobate crystals produced by combination of low-dose O3+ ion implantation and selective white light illumination,” Opt. Express 16, 10465–10470 (2008). [CrossRef]   [PubMed]  

31. D. L. Zhang, W. H. Wong, and E. Y. B. Pun, “Near-stoichiometric LiNbO3 optical waveguides fabricated using vapor transport equilibration and Ti co-diffusion,” Appl. Phys. Lett. 85, 3002–3004 (2004). [CrossRef]  

32. P.J. Chandler and F. L. Lama, “A new approach to the determination of planar waveguide profiles by means of a non-stationary mode index calculation,” Optica Acta 33, 127–142 (1986). [CrossRef]  

33. E. Flores-Romero, G. V. Vázquez, H. Márquez, R. Rangel-Rojo, J. Rickards, and R. Trejo-Luna, “Optical channel waveguides by proton and carbon implantation in Nd:YAG crystals,” Opt. Express 15, 8513–8520 (2007). [CrossRef]   [PubMed]  

34. R. Degl’Innocenti, A. Guarino, G. Poberaj, and P. Günter, “Second harmonic generation of continuous wave ultraviolet light and production of β-BaB2O4 optical waveguides,” Appl. Phys. Lett. 89, 041103 (2006). [CrossRef]  

35. U. Schlarb and K. Betzler, “Refractive indices of lithium niobate as a function of temperature, wavelength and composition: A generalized fit,” Phys. Rev. B 48, 15613–15620 (1993). [CrossRef]  

36. Rsoft Design Group, Computer software BeamPROP version 8.0, http://www.rsoftdesign.com.

37. R. Regener and W. Sohler, “Loss in low-finesse Ti:LiNbO3 optical waveguide resonators,” Appl. Phys. B 36, 143–147 (1985). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1.
Fig. 1. Schematic plot of the waveguide fabrication process. The inset shows the microscope image of the transverse cross section of the channel waveguide. The dashed lines denote the location of the channel waveguide.
Fig. 2.
Fig. 2. Profile of extraordinary refractive index changes Δn e induced by the ion implantation as function of the penetration depth of the incident ions inside the crystal (pointing z axis).
Fig. 3.
Fig. 3. (a). The 2D n e distribution of the channel waveguide at the cross section, (b). 3D plot of the calculated modal profile of the quasi-TM00 mode by FD-BPM and (c). measured near-field intensity distribution of quasi-TM00 mode at wavelength of 632.8nm.
Fig. 4.
Fig. 4. (a). The 2D n e distribution of the channel waveguide at the cross section and (b) 3D plot of the calculated modal profile of the quasi-TM00 mode by FD-BPM at wavelength of 1.5µm.
Fig. 4.
Fig. 4. Confocal room-temperature luminescence emission spectra from (a) the bulk and (b) the channel waveguide volume, corresponding to the 4 I 13/2 4 I 15/2 transitions
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.