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

Near-stoichiometric Ti:Sc:LiNbO3 strip waveguide for integrated optics

Open Access Open Access

Abstract

We demonstrate near-stoichiometric Ti:Sc:LiNbO3 strip waveguide fabrication starting from a congruent LiNbO3 substrate with a technological process in sequence of Sc3+-diffusion-doping, Ti diffusion, and post Li-rich vapor transport equilibration. We show that the waveguide is in a near-stoichiometric composition environment, well supports single-mode propagation at 1.5 μm wavelength under both TE and TM polarizations, shows considerable polarization dependence, and has a loss ≤ 0.4/0.7 dB/cm for TE/TM mode. The Ti4+ surface profile can be fitted by a sum of two error functions, the depth profile can be fitted by a Gaussian function, and the Sc3+-profile part which has a concentration above the threshold of photorefractive damage entirely covers the waveguide, showing that the waveguide is expected to be optical-damage-resistant.

© 2016 Optical Society of America

1. Introduction

LiNbO3 (LN) is an important optical material for (nonlinear) integrated optics. Its excellent electro-optic, acousto-optic, and nonlinear optical properties, together with the possibility of using well-established technique to produce high-quality waveguide of low loss, have led to the development of high functionality devices and optical integrated circuits. However, the these devices suffer from serious photorefractive damage. Besides Mg2+ [1], some other optical-damage-resistant dopants have been reported. These include divalent Zn2+ [2], trivalent Sc3+ [3,4], Tm3+ [5] and In3+ [6], and tetravalent Hf4+ [7], Zr4+ [8,9] and Sn4+ [10]. Among them, Sc3+ doping requires a relatively low optical-damage-resistant threshold concentration Cth = 2 mol% for an LN with a Li2O content of 48.5 mol%, and the threshold is about 2.5 times lower than that of Mg2+ doping, ~5 mol%. Moreover, as the composition approaches the stoichiometry, the threshold lowers to 0.4 mol% [4]. Low threshold of optical-damage-resistant dopants is required to improve the material homogeneity and the optical quality of crystal when codoped with rare-earth ions, such as Er3+, and to increase the diffusivity and solid solubility of codoped rare earth ions. In addition, a near-stoichiometric (NS) LN displays some properties better than the congruent material, such as larger electro-optic and nonlinear coefficients, lower coercive electric field needed for periodical poling of ferroelectric microdomain. Ti4+ diffused LN (Ti:LN) waveguide is a basic unit of LN-based passive and active devices. The waveguide is of the merits of lower waveguide loss, higher thermal, electric and chemical stabilities, easily realizing rare-earth doping and retained crystalline phase. An NS Ti:LN doped with Sc3+ (Ti:Sc:LN) would be a promising component for developing various optical-damage-resistant active or passive devices. The realization of the NS Ti:Sc:LN waveguide would open up some new applications and construct a supporting platform for the development of related active and passive devices. Towards such a long-term goal, we should firstly demonstrate how to fabricate an NS Ti:Sc:LN single-mode strip waveguides. Seeing that a Sc3+-doped LN plate is attained not as easily as an undoped one, as an alternative, Sc3+ can be incorporated into the crystal by thermal diffusion method. A congruent Ti:Sc:LN strip waveguide can be fabricated at first by Ti4+ diffusion following Sc3+-diffusion-doping in an initially congruent LN substrate. The two-step process considers that Sc3+ diffuses definitely slower than Ti4+ [11, 12] and its profile must entirely cover the waveguide to satisfy the optical-damage-resistant demand. After Ti4+ diffusion, post Li-rich vapor transport equilibration (VTE) is carried out to bring the waveguide to an NS composition. In the previous work, Sc3+ diffusion properties [13] and Ti4+/Sc3+ co-diffusion characteristics [12] have been studied. Here, we demonstrate the fabrication and characterization of an NS Ti:Sc:LN strip waveguide.

2. Experimental description

The NS Ti:Sc:LN waveguide was fabricated starting from a congruent LN with a technological process in sequence of Sc3+ diffusion doping, Ti diffusion and post Li-rich VTE. Figure 1(a) depicts the fabrication procedure of the waveguide.

 figure: Fig. 1

Fig. 1 (a) Fabrication procedure, (b) image, and (c) near-field patterns of TE and TM modes at 1547 nm wavelength of 8-μm-wide NS Ti:Sc:LN strip waveguide.

Download Full Size | PDF

  • 1) Sc3+ diffusion doping: A commercial Z-cut congruent LN plate was used as the starting material. A 130 ± 2 nm thick Sc2O3 (99.999%) film was coated onto 2/3 surface part. The other 1/3 part of surface was uncoated for reference. After Sc2O3 coating, the plate was annealed in wet O2. The diffusion temperature/duration is 1060 °C/16 h.
  • 2) Waveguide fabrication: An array of Ti strips with an initial width of 8 μm and a thickness of 150 ± 2 nm were delineated on the 1/2 part of the Sc3+-diffused surface (the other 1/2 part of Sc3+-diffused surface was not coated for reference). Two adjacent strips separate by 200 μm. The Ti diffusion was carried out at 1060 °C for 6 h in wet O2.
  • 3) Post Li-rich VTE: After Ti diffusion, the sample was subjected to a Li-rich VTE at 1100 °C for 10 h. The Li-rich atmosphere was created in a closed Li-rich two-phase crucible.

The sample surface can be divided into three parts: undoped area, Sc3+-only doped part and Sc3+/Ti4+-diffused array waveguide part. After each annealing stage, the surface refractive index at the three surface parts were measured. On the surface part with the waveguide array, the measurements were carried out by aligning the light beam perpendicular to the strip waveguide axis to avoid the excitation of waveguide mode, which may affect the result. The refractive index was measured at the 1553 nm wavelength using a Metricon 2010 prism coupler. Both ordinary and extraordinary indices were measured.

The near-field mode was measured by endface butt-coupling technique. A polarized light emitted from a 1547 nm single-frequency laser was coupled into the waveguide via endface butt-coupling between a section of polarization-maintaining fiber and a channel waveguide. The magnified near-field image of guided mode was projected onto an infrared Hamamatsu digitizing video camera (Vidicon 1000) through a Nikon microscopic objective lens.

The surface and/or depth profiles of diffused Sc3+ and Ti4+ ions were analyzed by secondary ion mass spectrometry (SIMS). A time-of-flight secondary ion mass spectrometer (ToF SIMS V; ION-TOF GmbH, Münster, Germany) was used to analyze the surface Ti4+ profile and the depth profiles of all of the substrate ions 6Li, 93Nb, 16O and the diffused ions 45Sc and 48Ti. The surface profile was obtained by doing surface mapping with a raster size of 135 × 135 μm2. For the depth analysis, a Cs+-beam (~45 μm in diameter) of 32 nA at 3 keV was used to sputter a crater of 150 × 150 μm2 on the strip waveguide surface and a pulsed bismuth ion beam (pulsed current: 1 pA, pulsed energy: 25 keV) was used to analyze the yields of secondary ions 6Li, 93Nb, 16O, 45Sc and 48Ti as a function of time. Positive secondary ions were detected. Ions from a central area of 18 × 18 μm2 inside the erosion crater rastered on the strip waveguide were detected. A low energy pulsed electron gun was used to reduce the surface charge accumulation. For the same purposes, a 50 nm thick Au film was coated onto the sample surface to be analyzed before the SIMS analysis. The trace and depth of the erosion crater were measured by a Tencor Alpha Step 200 profilometer (KLA-Tencor Corp., Milpitas, CA). The depth resolution is better than 5 nm.

3. Results and discussion

It is crucial to know if Sc3+ doping affects the LN refractive index, and if so, it is unclear if its contribution is comparable to the Ti4+-induced increment Δno (Δne), which is ~0.006 (0.012) at 1.5 µm wavelength for a usual congruent Ti:LN waveguide fabricated by diffusion of ~100-nm-thick Ti-strips at 1030-1060 °C for 9 h, which gives rise to a surface Ti4+ concentration 1.2 × 1021 ions/cm3. The contribution can be determined by comparing the index values at Sc3+-only doped and undoped surface parts. In the earlier work, we have studied it and concluded that the contribution is small compared to the Δno,e of usual congruent waveguide [13]. The conclusion is further verified by present work. The measurement, which has an error of ± 1 × 10−3, shows that, after the first, second and third heat treatment stages, the no (ne) value at the Sc3+-only doped surface part is 2.2113 (2.1376), 2.2115 (2.1379) and 2.2113 (2.1262), respectively, and that at the undoped part is 2.2113 (2.1377), 2.2114 (2.1378) and 2.2112 (2.1260), respectively. We note that Sc3+ doping contribution is on the order of 10−4 for both cases of no and ne, and is minor in both cases of congruent and NS compositions. It is also crucial to know the Li2O content at waveguide surface. Here, the refractive index measurement method is used to evaluate the Li2O-contents at each part of sample surface based on the Li-composition-dependent birefringence [14]. The refractive index measurement accuracy, 1.0 × 10−3, results in a Li2O-content uncertainty of 0.1 mol%. First, evaluation was performed on the as-grown congruent plate, for which the index was measured to be no = 2.2112 and ne = 2.1375. As expected, it has a Li2O-content 48.6 ± 0.1 mol%, which is in good agreement with the nominal value. After the VTE process, the Li2O-contents at the Sc3+-only doped and undoped surface parts are evaluated as 49.6 and 49.7 ± 0.1 mol%, respectively. Both are near-stoichiometric and can be thought as same within the error. It is impossible to obtain the Li2O-content accurately on the surface of an individual strip waveguide on the basis of the refractive index measurement by prism coupler as the coupler is designed for a thin film and slab waveguide rather than the strip waveguide. The reason is detailed below. The working light spot of the prism coupler has a diameter of 1 mm. The width of strip waveguide under study is usually several micrometers only (8 μm here) and the waveguides separate by 200 μm each other. The area part of waveguide surface seen by the light spot is no more than 5%. Thus, the measured pattern of the light reflected from sample surface is mainly contributed from the substrate surface instead of the narrow strip waveguide surface. Moreover, the Ti4+ presence, which induces the index increase, makes it difficult to obtain the correct Li2O-content of the strip waveguide. We have no choice but to roughly evaluate the waveguide composition by referencing the situation at the surface part away from the waveguide region. We have done the refractive index measurement on the surface part with the waveguide array. The resultant composition can be approximated as that at the surface part away from the waveguide because the effect of strip waveguide presence on the result is small as demonstrated above. The refractive indices measured at the surface part with the waveguide array are no = 2.2114 and ne = 2.1264 at the 1553 nm wavelength. The corresponding birefringence yields a Li2O content of 49.7 ± 0.1 mol%, which is the same as the values at the Sc3+-only doped and undoped surface parts. Although the exact Li composition at the strip waveguide surface cannot be obtained, it should not have a large difference from that of the surface part where no waveguide is defined. In words, the waveguide under study is in an NS composition environment.

It is worthwhile to note that the real Li2O content in the Sc3+-doped layer is not so high because some Sc3+ ions enter into Li+ sites (ScLi) and others enter into Nb5+ sites as its concentration CSc is above the threshold Cth according to the commonly accepted defect model for LN, and this is the case for the sample studied here because the CSc at the sample surface, as shown later, is 3.7 mol% that is far above the Cth ( = 0.4 mol% for an NS LN [4]). Each ScLi requires additional two Li vacancies (V). The real composition in the Sc3+-doped layer should consider the ratio R = ([LiLi] + [ScLi] + [VLi])/([NbNb] + [ScNb]), where the sign [XY] denotes the concentration of X ion on Y site. The above-mentioned 49.6 mol% Li2O content only implies that the R value in the Sc3+-doped layer is equivalent to that of a pure LN having a Li2O content 49.6 mol%. The real Li2O content there is smaller than 49.6 mol%. One can anticipate that Li2O content in the waveguide should have an even lower value as Ti4+ enters both Li and Nb sites too [15].

Figure 1(b) shows the waveguide surface image (1000 × ). The waveguide has a smooth and flat surface. The end-fire experiment shows that the waveguide well supports single-mode (at 1.5 μm wavelength) under both transverse electric (TE) and magnetic (TM) polarizations. Figure 1(c) shows the TE and TM mode patterns. Figures 2(a) and 2(b) show the TE- and TM-mode light intensity profiles along the width direction x and depth direction y. The light intensity of the guided mode follows a Gauss function Axexp[-2(x/Wx)2] in the x direction and a Hermite-Gauss function Ayy2exp[-2(y/Wy)2] in the y direction. The black lines represent the Gauss or Hermite-Gauss fit. The mode size Wx × Wy is 5.7 × 4.5 μm2 for the TE mode and 4.9 × 4.1 μm2 for the TM mode. As expected, the confinement to the TM mode is considerably stronger than that to the TE mode because Ti-induced refractive index increment Δne >Δno [for a Z-cut LN substrate, TM (TE) mode concerns Δne (Δno)].

 figure: Fig. 2

Fig. 2 TE- and TM-mode light intensity profiles along (a) x and (b) y directions of 8-μm-wide NS Ti:Sc:LN strip waveguide.

Download Full Size | PDF

The waveguide loss was evaluated from the insertion loss measured at 1547 nm wavelength. The fiber-to-fiber insertion loss of the 8-μm-wide, 1.5-cm-long waveguide was measured to be 4.5/5.6 dB under the TE/TM polarizations. From the obtained mode sizes, the coupling loss between the waveguide and a single-mode fiber (with a mode field diameter of 10.3 μm) is 1.8/2.1 dB for the TE/TM mode. The total reflection loss is 0.3 dB. The measured insertion loss also includes the etalon effect, which may cause 0-2.0 dB loss in the resonator formed by the two endfaces of the waveguide and 0-0.5 dB loss in each air resonator formed by the endfaces of waveguide and fiber. The pitch of interference fringes is on the submicron order and requires a measurement accuracy of a few tens of nanometer for the resonator length. Such an accuracy is not easily achieved. So, the etalon effect cannot be determined accurately. With ignored etalon effect, the waveguide loss was evaluated as 0.4/0.7 dB/cm for the TE/TM mode.

Figure 3(a) shows the Ti4+ profile (magenta balls) on the waveguide surface. The experimental data can be well fitted by a sum of two error functions. The fitting expression and parameter values are indicated (W is the initial Ti-strip width). The diffusion width dx is 8.6 ± 0.2 µm. A mean lateral Ti4+ diffusivity is evaluated as 1.15 ± 0.03 μm2/h. Figure 3(b) shows the depth profiles of 6Li, 93Nb, 16O, 45Sc and 48Ti in the waveguide. One can see that both the Ti4+ and Sc3+ profiles can be well fitted by a Gaussian function, indicating that the diffusion reservoir, which is shared by Ti4+ and Sc3+, has been exhausted. The fitting expressions are indicated. The Ti4+- or Sc3+-concentration profile can be written as CTi, Sc(y) = C0Ti, Scexp[-(y/dTi, Sc)2], where C0Ti, Sc is the surface Ti4+ or Sc3+ concentration and dTi (dSc) is the 1/e Ti4+ (Sc3+) diffusion depth having a value of dTi = 5.8 ± 0.1 μm (dSc = 7.2 ± 0.1 μm). The mean bulk diffusivity is 0.47 ± 0.02 (0.41 ± 0.01) μm2/h for Ti4+ (Sc3+). In the congruent LN, Sc3+ diffuses definitely slower than Ti4+ due to their difference in ionic radius (0.8 Ǻ for Sc3+ and 0.6 Ǻ for Ti4+) [12]. For the diffusion system concerned here, however, both ions show similar diffusivities. Moreover, the Sc3+ (Ti4+) diffusivity is considerably larger (lower) than the value of their respective single diffusion case [13], [16], [17]. These characteristics are associated with the multiple annealing stages, Ti4+-assisted Sc3+ diffusion [12] and post VTE, which causes composition increase and hence Sc3+ or Ti4+ diffusivity decrease [18]. It is worthwhile to mention that the Ti4+ and/or Sc3+ out-diffusion does not occur during the VTE process. In the early stage of VTE, the Ti4+ and Sc3+ continuously diffuse deeper and their profiles become slightly broader. As the VTE is prolonged, Li+ ions diffuse fast into the bulk (Li+ diffuses 103-fold faster than Ti4+ or Sc3+). The waveguide layer quickly reaches to the NS composition. Consequently, the mobilities of Ti4+ and Sc3+ degrade largely. Both ions cannot diffuse out of crystal.

 figure: Fig. 3

Fig. 3 (a) Surface Ti4+ profile and (b) depth profiles of 6Li, 93Nb, 16O, 45Sc and 48Ti.

Download Full Size | PDF

Based on the law of mass conservation, the Ti4+ surface concentration is determined as C0Ti = (16.6 ± 0.5) × 1020 ions/cm3, equivalent to (9.0 ± 0.4) mol%. The Sc3+ surface concentration is determined as C0Sc = (6.9 ± 0.2) × 1020 ions/cm3, equivalent to (3.7 ± 0.1) mol%.

Next, we turn to discuss the photorefractive damage issue of the waveguide under study. To realize the optical-damage-resistant function, two requirements should be satisfied, i. e., the waveguide layer must be entirely covered by the Sc3+ profile as mentioned above, and the CSc at the 1/e waveguide depth should be above the threshold Cth. First, we pay attention to the former requirement. For the waveguide under study, the Sc3+ profile entirely covers the Ti4+ profile as dSc ( = 7.2 ± 0.1 μm) is considerably larger than dTi ( = 5.8 ± 0.1 μm). Speaking strictly, this requirement should be that the Sc3+ profile entirely covers the refractive index profile rather than the Ti4+ profile. This is because the 1/e depth of refractive index change, named dyo (dye) for the ordinary (extraordinary) ray, does not always equal the dTi as the Ti4+-induced refractive index increment ∆no,e and the Ti4+ concentration CTi does not always follow a linear relationship. For example, for a usual congruent Ti:LN waveguide, both follow a nonlinear relation with a power index αo = 0.53-0.55 [15], [19] for the ordinary ray, and a near or exact linearity with a power index αe = 0.83 [15] or 1.0 [19] for the extraordinary ray. For the NS Ti:Sc:LN waveguide studied here, the αo,e value is unknown at present. However, we have already known the value (αo = 0.90 and αe = 1.0 [20]) for an NS Ti:LN waveguide fabricated by the same method and condition of Ti diffusion followed by the VTE (1100 °C/10 h), which causes similar surface Li2O content 49.8 mol%. Since Sc3+ doping has little contribution to the LN refractive index, it is reasonable to consider that the αo,e value of the NS Ti:LN waveguide is also valid for the NS Ti:Sc:LN studied here. To reach an optical-damage-resistant waveguide, it requires dSc ≥ dTio,e1/2. For the studied waveguide, dSc should be at least dTio1/2 = 6.1 μm for the ordinary ray and dTie1/2 = 5.8 μm for the extraordinary ray. One can see that the practical Sc3+ profile (dSc = 7.2 μm) is still considerably larger than the required for both cases of ordinary and extraordinary rays.

Subsequently, attention is paid to the latter requirement mentioned above. The nonlinearity between ∆no,e and CTi does not affect the requirement on the surface Sc3+ concentration, i. e. C0SceCth (e = 2.71828). The C0Sc at the waveguide surface and at 1/e Ti4+ concentration depth are 3.7 mol% and C0Sc/e = 1.4 mol%, respectively, which are 9.3 and 3.5 times higher than Cth ( = 0.4 mol% for an NS LN [4]). Moreover, the Sc3+ profile shows that the depth where the Sc3+ concentration is above Cth can be extended from dSc ( = 7.2 μm) to 10.7 μm. One can see from Fig. 2(b) that the mode field depth range of 0-5 μm is indeed entirely covered by the Sc3+ profile part above the threshold (0.0-10.7 μm) for both cases of TE and TM modes. This implies that the studied waveguide is expected to be optical-damage-resistant. Future work should investigate the optical damage resistance of the NS Ti:Sc:LN waveguide using the holographic Bragg grating method developed for a strip waveguide [21].

Finally, based on the SIMS result, the refractive index profile in the NS waveguide studied can be modeled as no,e(x,y) = nso,e + Δn1o,eerfc(-y/do,e) + Δno,e{erf[(W/2 + x)/dxo,e] + erf[(W/2-x)/dxo,e]}exp[-(y/dyo,e)2/{2erf[W/(2dxo,e)]}, where -∞<x< + ∞, y≥0, the first term represents the substrate index, the second term reflects the Li composition (i. e., Li-rich VTE) effect on index [22], and the third term denotes the Ti-induced index increment profile. Δn1o,e is the composition-induced maximum index change. The do,e, dxo,e and dyo,e are the width or depths of index increment profiles.

4. Conclusion

We have demonstrated NS Ti:Sc:LN strip waveguide fabricated by Sc3+-diffusion-doping, Ti diffusion and post Li-rich VTE. We show that the waveguide is in an NS composition environment, well supports both the TE and TM modes, shows considerable polarization dependence, is single-mode at 1.5 μm, and has a loss ≤ 0.4/0.7 dB/cm for the TE/TM mode. The Ti4+ surface profile can be fitted by a sum of two error functions and the depth profile can be fitted by a Gaussian function. The mode field profile is entirely covered by the Sc3+ profile part having a Sc3+ concentration above the threshold for both cases of TE and TM modes, implying that the waveguide is expected to be optical-damage-resistant. The NS Ti:Sc:LN strip waveguide is promising for active and passive integrated optics.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (NSFC) under Project nos. 50872089, 61077039 and 61377060, by the Research Grants Council of the Hong Kong Special Administrative Region, China, under Project no 11211014, by the Tianjin Science and Technology Commission of China under Project no. 16JCZDJC37400, and by Overseas, Hong Kong & Macao Scholars Cooperative Researching Fund of NSFC.

References and links

1. D. A. Bryan, R. Gerson, and H. E. Tomaschke, “Increased optical damage resistance in lithium niobate,” Appl. Phys. Lett. 44(9), 847–849 (1984). [CrossRef]  

2. T. R. Volk, V. I. Pryalkin, and N. M. Rubinina, “Optical-damage-resistant LiNbO3:Zn crystal,” Opt. Lett. 15(18), 996–998 (1990). [CrossRef]   [PubMed]  

3. J. K. Yamamoto, K. Kitamura, N. Iyi, S. Kimura, Y. Furukawa, and M. Sato, “Increased optical damage resistance in Sc2O3-doped LiNbO3,” Appl. Phys. Lett. 61(18), 2156–2158 (1992). [CrossRef]  

4. M. Nakamura, S. Takekawa, Y. W. Liu, and K. Kitamura, “Crystal growth of Sc-doped near-stoichiometric LiNbO3 and its characteristics,” J. Cryst. Growth 281(2–4), 549–555 (2005). [CrossRef]  

5. J. P. de Sandro, J. K. Jones, D. P. Shepherd, M. Hempstead, J. Wang, and A. C. Tropper, “Non-photorefractive CW Tm-indiffused Ti:LiNbO3 waveguide laser operating at room temperature,” IEEE Photonics Technol. Lett. 8(2), 209–211 (1996). [CrossRef]  

6. K. Kasemir, K. Betzler, B. Matzas, B. Tiegel, T. Wahlbrink, M. Wöhlecke, B. Gather, N. Rubinina, and T. Volk, “Influence of Zn/In codoping on the optical properties of lithium niobate,” J. Appl. Phys. 84(9), 5191–5193 (1998). [CrossRef]  

7. L. Razzari, P. Minzioni, I. Cristiani, V. Degiorgio, and E. P. Kokanyan, “Photorefractivity of hafnium-doped congruent lithium-niobate crystals,” Appl. Phys. Lett. 86(13), 131914 (2005). [CrossRef]  

8. Y. F. Kong, S. G. Liu, Y. J. Zhao, H. D. Liu, S. L. Chen, and J. J. Xu, “Highly optical damage resistant crystal: Zirconium-oxide-doped lithium niobate,” Appl. Phys. Lett. 91(8), 081908 (2007). [CrossRef]  

9. G. Nava, P. Minzioni, W. B. Yan, J. Parravicini, D. Grando, E. Musso, I. Cristiani, N. Argiolas, M. Bazzan, M. V. Ciampolillo, A. Zaltron, C. Sada, and V. Degiorgio, “Zirconium-doped lithium niobate: photorefractive and electro-optical properties as a function of dopant concentration,” Opt. Mater. Express 1(2), 270–277 (2011). [CrossRef]  

10. L. Wang, S. Liu, Y. Kong, S. Chen, Z. Huang, L. Wu, R. Rupp, and J. Xu, “Increased optical-damage resistance in tin-doped lithium niobate,” Opt. Lett. 35(6), 883–885 (2010). [CrossRef]   [PubMed]  

11. D. L. Zhang, C. X. Qiu, W. H. Wong, D. Y. Yu, and E. Y. B. Pun, “Optical-damage-resistant Ti-diffused LiNbO3 strip waveguide doped with zirconium,” IEEE Photonics Technol. Lett. 26(17), 1770–1773 (2014). [CrossRef]  

12. D. L. Zhang, Q. Zhang, C. X. Qiu, W. H. Wong, and E. Y. B. Pun, “Control of scandium diffusion in LiNbO3 single crystal by co-diffusion of titanium,” J. Mater. Sci. 50(12), 4149–4159 (2015). [CrossRef]  

13. D. L. Zhang, C. X. Qiu, W. H. Wong, and E. Y. B. Pun, “Diffusion properties of scandium in lithium niobate crystal,” J. Am. Ceram. Soc. 97(9), 2903–2908 (2014). [CrossRef]  

14. U. Schlarb and K. Betzler, “Refractive indices of lithium niobate as a function of temperature, wavelength, and composition: A generalized fit,” Phys. Rev. B Condens. Matter 48(21), 15613–15620 (1993). [CrossRef]   [PubMed]  

15. E. Zolotoyabko, Y. Avrahami, W. Sauer, T. H. Metzger, and J. Peisl, “High-temperature phase transformation in Ti-diffused waveguide layers of LiNbO3,” Appl. Phys. Lett. 73(10), 1352–1354 (1998). [CrossRef]  

16. S. Fouchet, A. Carenco, C. Daguet, R. Guglielmi, and L. Riviere, “Wavelength dispersion of Ti induced refractive index change in LiNbO3 as a function of diffusion parameters,” J. Lightwave Technol. 5(5), 700–708 (1987). [CrossRef]  

17. M. Fukuma and J. Noda, “Optical properties of titanium-diffused LiNbO3 strip waveguides and their coupling-to-a-fiber characteristics,” Appl. Opt. 19(4), 591–597 (1980). [CrossRef]   [PubMed]  

18. R. J. Holmes and D. M. Smyth, “Titanium diffusion into LiNbO3 as a function of stoichiometry,” J. Appl. Phys. 55(10), 3531–3535 (1984). [CrossRef]  

19. E. Strake, G. P. Bava, and I. Montrosset, “Guided modes of Ti:LiNbO3 channel waveguides: a novel quasi-analytical technique in comparison with the scalar finite-element method,” J. Lightwave Technol. 6(6), 1126–1135 (1988). [CrossRef]  

20. D. L. Zhang, J. Kang, C. X. Qiu, Q. Zhang, W. H. Wong, and E. Y. B. Pun, “Refractive index in Ti:LiNbO3 fabricated by Ti diffusion and post-Li-rich VTE,” IEEE Photonics Technol. Lett. 27(10), 1132–1135 (2015). [CrossRef]  

21. J. Hukriede, D. Runde, and D. Kip, “Fabrication and application of holographic Bragg gratings in lithium niobate channel waveguides,” J. Phys. D Appl. Phys. 36(3), R1–R16 (2003). [CrossRef]  

22. D. L. Zhang, P. Zhang, H. J. Zhou, and E. Y. B. Pun, “Characterization of near-stoichiometric Ti:LiNbO3 strip waveguides with varied substrate refractive index in the guiding layer,” J. Opt. Soc. Am. A 25(10), 2558–2570 (2008). [CrossRef]   [PubMed]  

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) Fabrication procedure, (b) image, and (c) near-field patterns of TE and TM modes at 1547 nm wavelength of 8-μm-wide NS Ti:Sc:LN strip waveguide.
Fig. 2
Fig. 2 TE- and TM-mode light intensity profiles along (a) x and (b) y directions of 8-μm-wide NS Ti:Sc:LN strip waveguide.
Fig. 3
Fig. 3 (a) Surface Ti4+ profile and (b) depth profiles of 6Li, 93Nb, 16O, 45Sc and 48Ti.
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.