• No results found

Development of low-loss tio2 waveguides

N/A
N/A
Protected

Academic year: 2021

Share "Development of low-loss tio2 waveguides"

Copied!
9
0
0

Bezig met laden.... (Bekijk nu de volledige tekst)

Hele tekst

(1)

Development of low-loss TiO

2

waveguides

I. H

EGEMAN

, M. D

IJKSTRA

, F. B. S

EGERINK

, W. L

EE

,

AND

S. M.

G

ARCIA

-B

LANCO*

Optical Sciences Group, MESA+ Institute for Nanotechnology, University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands

*s.m.garciablanco@utwente.nl

Abstract: TiO2channel waveguides were fabricated using a DC sputter deposition process,

followed by photolithography and reactive ion etching. A SiO2cladding was deposited using

evaporation. SEM, TEM and Raman measurements indicate the presence of both an amorphous and a crystalline phase. As the layer thickness increases, poly-crystalline structures start forming. Loss measurements were performed by imaging the scattered light from the top of the channel waveguides and fitting an exponential decay to the intensity profile. Propagation losses of 7.8 ± 0.52 dB/cm at a wavelength of 632.8 nm and 0.68 ± 0.46 dB/cm at a wavelength of 1010 nm were experimentally characterized.

© 2020 Optical Society of America under the terms of theOSA Open Access Publishing Agreement

1. Introduction

The demand for photonic integrated circuits is rapidly increasing. TiO2is an emerging material

for integrated photonics, which has received increasing interest in the last few years [1–5]. Its high refractive index (i.e., 2.2-2.3 for amorphous layers and even higher when in its crystalline phases anatase or rutile [1]) permits the design of compact integrated circuits. TiO2is transparent

over a broad wavelength range, spanning from visible to infrared wavelengths, due to its high bandgap of 3 eV [6]. Its high refractive index makes TiO2a promising material for waveguide

enhanced Raman spectroscopy [7], experimentally showing a 50 fold higher Raman efficiency compared to Si3N4waveguides [4].

TiO2can be doped with, for example, rare earth ions [8] to achieve optical gain in different

wavelength ranges. The low phonon energy of TiO2, with a maximum at 633 cm−1for anatase

[9], limits the amount of non-radiative decay. Examples of active photonic applications include Si:TiO2waveguides doped with Er3+ions [10] and rare-earth ion doped layers produced by the

sol-gel method [11,12].

TiO2exhibits a high non-linear refractive index, approximately 3 times larger than that of

silicon nitride [8]. Several non-linear processes have been recently demonstrated in TiO2

waveguides, such as third harmonic generation [13], supercontinuum generation [2], spectral broadening [14] and four wave mixing [5].

The thermo-optic coefficient of TiO2 is negative and ranges between −0.5×10−4 K−1and

−2.15×10−4K−1[1517]. This variance is related to the variation in the material characteristics

of the layer, such as crystallinity, stress and density, as a result of varying deposition methods. By combining TiO2with a material with a positive thermo-optic coefficient, athermal waveguides

can be designed. Athermal waveguides have been demonstrated by combining TiO2with Si3N4

[18,19] and silicon [17]. Such athermal behavior was demonstrated over a wavelength range of up to 300 nm [20].

In order for TiO2to find actual applications, low propagation losses need to be obtained to

get high quality devices, which still remains a challenge. The lowest losses reported to date for channel waveguides fabricated using conventional fabrication techniques (i.e., deposition, lithography and reactive ion etching) are 9.7 dB/cm at 632.8 nm [21] and 4 dB/cm at 1550 nm [22]. Lower losses were obtained using a dielectric lift-off process, which resulted in 7.5 dB/cm

#380793 https://doi.org/10.1364/OE.380793

(2)

at 632.8 nm and 1.2 dB/cm at 1550 nm [3]. The lift-off process, however, limits the freedom in the design of photonic integrated circuits.

In this work, we developed a DC reactive sputtering process to deposit low-loss TiO2layers.

Sputter deposition allows for fast wafer-scale deposition of thin layers providing good control over their stoichiometry and morphology. Patterning is performed using UV contact lithography followed by reactive ion etching. These techniques are widely available and allow fast processing at the wafer scale. Waveguides are cladded with evaporated SiO2to reduce the index contrast

between core and cladding, which decreases optical propagation losses. Propagation losses as low as 7.8 ± 0.52 dB/cm at 632.8 nm and 0.68 ± 0.46 dB/cm at 1010 nm were measured in the fabricated waveguides, which are comparable to the losses obtained with the lift-off process [3], but considerably lower than any previous work utilizing similar standard fabrication techniques [5,21,22].

2. Fabrication

A thin TiO2layer was deposited using DC reactive sputter deposition from a 99.999% pure 4

inch titanium target in the TCOater in the MESA+ cleanroom. The substrate consists of a 4 inch silicon wafer with an 8 µm thick layer of thermal SiO2. During deposition, the target to substrate

distance was kept at 176 mm, the reactor pressure at 6·10−3 mbar and the temperature of the

substrate was kept at room temperature (i.e., between 20-30 °C). A DC power of 500 W was applied to the titanium target. No annealing was performed after deposition, since it is known to lead to high amounts of crystallization, increased surface roughness [23,24] and a high extinction coefficient [25].

The O2flow has major impact on the quality of the deposited films. As the O2content in

the reactor increases, the bias voltage between the target and the wafer increases. The increase in the bias voltage is related to an increase in the oxidation state of the target with increased O2flow, passing from fully metallic to fully oxidized. Since the secondary electron yield of

titanium is higher than that of TiO2 [26], an increasingly oxidized target exhibits a reduced

secondary electron current and, since the applied power is kept constant, the bias voltage increases. Figure1(a) shows the variation of bias voltage as a function of O2flow into the chamber. As

the O2flow is increased, the argon flow is adjusted to keep a total flow of 40 sccm. The data

is collected for an increasing O2flow rate. At around 4.5 sccm of O2flow, a jump in the bias

is observed, known as the ‘knee point’ [27]. During deposition both the sputtered titanium and the titanium target are getting oxidized. For low O2flows the removal rate at the target is

sufficient to keep the target from forming a full oxidized layer. However, at the knee point a threshold is reached causing a self-enforcing loop, where the target gets oxidized, resulting in a lower deposition rate. The lower deposition rate causes less gettering of oxygen at the wafer, which in turn increases the O2partial pressure leading to more oxidation of the target. Layers

with the highest deposition rate and the lowest propagation losses are deposited with an O2flow

just after the knee point (i.e. at higher O2 flows). For an O2 flow below the knee point, the

deposited layers are not fully oxidized, therefore exhibiting high optical propagation losses. This point is indicated in Fig.1(a) with the triangle. As the O2flow increases passed the knee point,

optically guiding layers are obtained. However, the deposition rate drastically decreases due to the increased oxidation state of the target. Layers deposited at a too high O2flow show higher

scattering losses, due to an increased crystallization of the layer [28]. Figures1(b)-(d) show the guiding (from right to left) of the first order TE slab mode in a TiO2layer, deposited at 4.5,

5 and 5.5 sccm of O2flow. The layers deposited at higher O2flows show higher propagation

losses. Coupling into the slab mode is performed using a Metricon 2010M prism coupling setup with a rutile prism. In order to get a stable process, the O2 flow is kept at 4.6 sccm. In this

(3)

waveguides have estimated losses of 3 dB/cm at 632 nm, based on an estimated 15 dB attenuation over 5 cm.

Fig. 1. (a) Bias voltage as a function of oxygen flow during sputter deposition (left axis) and corresponding deposition rate (right axis) as a function of O2flow into the sputtering

chamber. The total gas flow rate (O2+Ar) is kept at 40 sccm. The red triangle indicates a layer deposited in the metallic regime, which did not show guiding. The three squares correspond to three guiding layers. The quality of the guiding of these three layers is shown in (b) layer deposited at 4.5 sccm O2, (c) 5 sccm O2and (d) 5.5 sccm O2. The slab modes

propagate from right to left. The coupling is performed by prism coupling, which is visible at the right side of the figures.

Fig. 2. (a) High resolution scanning electron microscope image of a cleaved end facet of a 140 nm thick and 1 µm wide TiO2waveguide, before evaporating the SiO2top cladding. (b) Simulated fundamental TE mode at a wavelength of 632 nm. (c) Simulated fundamental TE mode at a wavelength of 1010 nm.

(4)

After deposition, 1 µm wide waveguides were defined using UV contact photolithography using a 1.7 µm thick photoresist layer (Olin Oir 907-17). Etching is performed using an Oxford Plasma Pro 100 Cobra reactive ion etching machine. During etching, the pressure is kept at 10 mTorr and the temperature is set at 10 °C. An ICP power of 1500 W and a CCP power of 20 W are used. The plasma consists of SF6, O2and Ar with flowrates of 25, 6 and 5 sccm respectively.

This process results in channel waveguides as shown in Fig.2(a). The waveguides are covered by a 1 µm thick SiO2cladding, which is deposited by e-beam evaporation from a 99,999% pure

SiO2target in the BAK600 in the MESA+ cleanroom. The target to substrate distance is kept

at 15 cm. Evaporation was chosen in order to keep the sample at room temperature to prevent crystallization of the TiO2. The settings for the evaporations were a beam current of 200 mA and

a base pressure of 6·10−6mbar.

Simulations using Lumerical MODE solutions show four TE modes and three TM modes at a wavelength of 632 nm and two TE modes and one TM mode at a wavelength of 1010 nm. Normalized mode profiles of the fundamental TE modes are shown in Figs.2(b) and (c) for the wavelengths of 632 nm and 1010 nm respectively. These modes show confinement of the light in the waveguide core of 76.7% and 51.5% respectively.

3. Characterization

AFM measurements were performed on deposited TiO2layers of different thicknesses with a

Bruker dimension fastscan atomic force microscope (AFM). They are shown in Figs.3(a) and (b). The surface roughness of these layers is determined by calculating the root mean square of the height profile of these images. The rms roughness values are shown in Fig.3(c) as a function of layer thickness. The surface roughness increases linearly with increasing layer thickness. The increasing grain size indicates increasing amount of crystallization [29,30]. As a result, thick waveguides exhibit higher scattering losses.

304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404

mTorr and the temperature is set at 10 °C. An ICP power of 1500 W and a CCP power of 20 W are used. The plasma consists of SF6, O2and Ar with flowrates of 25, 6 and 5 sccm respectively.

This process results in channel waveguides as shown in Fig.2(a). The waveguides are covered by a 1 µm thick SiO2cladding, which is deposited by e-beam evaporation from a 99,999% pure

SiO2target in the BAK600 in the MESA+ cleanroom. The target to substrate distance is kept

at 15 cm. Evaporation was chosen in order to keep the sample at room temperature to prevent crystallization of the TiO2. The settings for the evaporations were a beam current of 200 mA and

a base pressure of 6·10−6mbar.

Simulations using Lumerical MODE solutions show four TE modes and three TM modes at a wavelength of 632 nm and two TE modes and one TM mode at a wavelength of 1010 nm. Normalized mode profiles of the fundamental TE modes are shown in Figs.2(b) and (c) for the wavelengths of 632 nm and 1010 nm respectively. These modes show confinement of the light in the waveguide core of 76.7% and 51.5% respectively.

3. Characterization

AFM measurements were performed on deposited TiO2layers of different thicknesses with a

Bruker dimension fastscan atomic force microscope (AFM). They are shown in Figs.3(a) and (b). The surface roughness of these layers is determined by calculating the root mean square of the height profile of these images. The rms roughness values are shown in Fig.3(c) as a function of layer thickness. The surface roughness increases linearly with increasing layer thickness. The increasing grain size indicates increasing amount of crystallization [29,30]. As a result, thick waveguides exhibit higher scattering losses.

Fig. 3. (a-b) Surface images made by AFM of TiO2layers with deposited thickness of 32

and 172 nm respectively. (c) Surface roughness (root mean square) as a function of deposited TiO2layer thickness.

The crystallinity of the layer was determined using transmission electron microscope (TEM) imaging. The TEM-sample is prepared by thinning a cross section of the TiO2by using focused

ion beam milling, to obtain a thin TiO2lamella of several tens of nanometers. To prevent damage

to the TiO2 layer during the thinning process, a protective coating of 3-4 µm of platinum is

deposited on top of the TiO2layer. This deposition might, however, cause amorphization of

the top surface of the sample. This is prevented by applying a 20 nm carbon layer before the platinum deposition. The TEM imaging is performed using a Philips CM300ST FEG TEM. Figure4(a) shows an overview of the TEM sample, consisting of the SiO2 substrate, TiO2

layer and a protective coating of carbon and platinum. In this image, it is possible to observe a column-like structure appearing after the initial 50 nm of the TiO2 layer. The same structure

can be seen in the SEM image, Fig.2(a). A zoomed-in image of the initial 50 nm of the TiO2

layer is shown in Fig.4(b). The corresponding Fourier transform (FT) is shown in Fig.4(c). The almost smooth halo indicates that the first 50 nm of the deposited layer consists of predominantly amorphous TiO2. A zoomed-in image of the top part of the TiO2layer exhibiting the column-like

Fig. 3. (a-b) Surface images made by AFM of TiO2layers with deposited thickness of 32

and 172 nm respectively. (c) Surface roughness (root mean square) as a function of deposited TiO2layer thickness.

The crystallinity of the layer was determined using transmission electron microscope (TEM) imaging. The TEM-sample is prepared by thinning a cross section of the TiO2by using focused

ion beam milling, to obtain a thin TiO2lamella of several tens of nanometers. To prevent damage

to the TiO2 layer during the thinning process, a protective coating of 3-4 µm of platinum is

deposited on top of the TiO2 layer. This deposition might, however, cause amorphization of

the top surface of the sample. This is prevented by applying a 20 nm carbon layer before the platinum deposition. The TEM imaging is performed using a Philips CM300ST FEG TEM. Figure4(a) shows an overview of the TEM sample, consisting of the SiO2 substrate, TiO2

layer and a protective coating of carbon and platinum. In this image, it is possible to observe a column-like structure appearing after the initial 50 nm of the TiO2layer. The same structure

(5)

Fig. 4. (a) TEM image of a 140 nm thick TiO2layer on a SiO2substrate, covered by carbon

and platinum protective coatings. (b) Zoomed-in TEM image of the first 50 nm of the TiO2

layer, showing the border with the SiO2under-cladding (interface on the top right corner).

(c) Fourier transform (FT) of Fig.3(b), showing amorphous structure. (d) Zoomed-in TEM image of the top TiO2layer. (e) FT of Fig.3(d) (full image), showing the appearance of

crystalline phases. (f) FT of the marked subsection of Fig.3(d), showing some features that are hidden in the full FT due to averaging effects.

layer is shown in Fig.4(b). The corresponding Fourier transform (FT) is shown in Fig.4(c). The almost smooth halo indicates that the first 50 nm of the deposited layer consists of predominantly amorphous TiO2. A zoomed-in image of the top part of the TiO2layer exhibiting the column-like

structure, is shown in Fig.4(d). As the thickness of the deposited layer increases, a crystalline phase becomes visible, as shown in Figs.4(e) and (f), which correspond to the FT of the full area and the marked area of Fig.4(d), respectively. The main features in Figs.4(e) and 4(f) show periodicities of around 2.1 and 2.8 Å, which correspond to the Ti-O and O-O bonds in anatase, respectively [31]. A possible explanation for the structure of the deposited film could be found in the structure zone model [32]. The deposited films in this work show poly-crystalline features, shown in the TEM images, and changing structure with increasing film thickness, shown in the AFM images, which corresponds to zone T in the structure zone model.

Confocal Raman measurements were performed on a 100 nm thick TiO2layer on a fused silica

substrate to further confirm the crystalline structure of the layers. A WITEC alpha300R/S/A instrument was utilized. Pump light with a wavelength of 785 nm and a power of 185 mW is used in combination with a 100x objective with an NA of 0.9. The spectrum is measured with an integration time of 1 s and 10 accumulations. It was verified that the power density in the laser focal point did not induce the crystallization of the layer. The Raman spectrum is shown in Fig.5, which clearly shows the presence of the anatase crystalline phase with peaks at 151, 409, 515 and 633 cm−1[9].

(6)

Fig. 5. Raman spectrum of a 100 nm thick TiO2layer on a fused silica substrate.

The propagation losses of a 140 nm thick and 1 µm wide cladded waveguide are measured by fiber coupling and imaging the scattered light from the top of the chip. For the losses at 632.8 nm

Fig. 6. (a) Propagation of light at a wavelength of 632.8 nm. (b) Exponential fit resulting in losses of 7.8 ± 0.52 dB/cm at 632.8 nm. (c) Propagation of light at a wavelength of 1010 nm. (d) Exponential fit resulting in losses of 0.68 ± 0.46 dB/cm at 1010 nm.

(7)

a HeNe laser is used, the measurements at 770, 859, 979 and 1010 nm are performed using a Ti-sapphire laser, the wavelength of which is measured using a spectrometer. After optimizing the alignment for maximum coupling efficiency, top view images are taken with a Point Grey monochrome camera (BFLY-U3-23S6M-C). The linearity of camera was measured to be 0.1% over 2 decades of optical power. Figure6(a) shows a top view image of the light propagation at 632.8 nm. The corresponding exponential fit of the intensity as a function of propagation length is given in Fig.6(b), showing propagation losses of 7.8 ± 0.52 dB/cm. The same is shown in Figs.6(c) and6(d) for a wavelength of 1010 nm. Losses as low as 0.68 ± 0.46 dB/cm can be observed. For the measurement at 1010 nm, the scattering point has not been taken into account for the exponential fit, since this single defect is not representative for the overall waveguide losses. The error margin of the losses is obtained by calculating the standard deviation of the fit parameters.

The wavelength dependent propagation loss is summarized in Fig.7. The significant increase in propagation losses towards the visible wavelength range, indicate that the losses are mainly caused by scattering, which is a result of the poly-crystallinity of the layers and top and sidewall roughness. This is to be expected, since TiO2only starts absorbing below a wavelength of roughly

420 nm. For higher wavelengths, where the loss starts to decay, the relative error increases. Due to the short length of the spirals utilized for these measurements, the lowest value for the loss that can be measured is limited. Loss-measurements at wavelengths of 1310 nm and 1550 nm where non-successful, due to low signal-to-noise ratio. At these wavelengths, the Si substrate becomes transparent. The rough background of the chip scatters the light and causes higher amounts of background noise.

Fig. 7. Propagation loss as a function of wavelength.

Further reduction of the propagation losses could be obtained by reducing surface roughness by using a chemical mechanical polishing step, as it has been done in the Si3N4waveguide platform

[33]. Furthermore the sidewall roughness might be reduced by optimizing the etching recipe and by using e-beam lithography for higher resolution of the resist. Further optimization of the deposition process can reduce the amount of crystallinity, which will reduce volume scattering losses and reduce the surface roughness.

4. Conclusion

TiO2waveguides were fabricated with DC reactive sputter deposition, followed by

photolithog-raphy and reactive ion etching. A SiO2cladding is applied to obtain a less fragile device with

lower propagation losses. Combination of TEM imaging and Raman measurements revealed the presence of anatase crystalline phase in the upper TiO2layer. The lower TiO2layer remained

(8)

of the layer increases, which corresponds to an increased crystallinity of the layer. Channel waveguide losses as low as 0.68 ± 46 dB/cm at a wavelength of 1010 nm were measured. The waveguide losses increase towards lower wavelengths, due to the increased amount of scattering losses. Further reduction of the propagation losses might be obtained by using E-beam lithography, chemical mechanical polishing and further optimization of the deposition process.

Funding

Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWA.ID.17.100).

Acknowledgments

We would like to thank Mike Dikkers of the MESA+ cleanroom staff for useful discussions about reactive sputtering and for support with the sputter coater utilized in this work. We would also like to thank Rico Keim to carry out the TEM measurements and Mark Smithers for high resolution SEM image, both from the MESA+ staff. René Heideman, from LioniX International, is also acknowledged for useful discussions.

Disclosures

The authors declare no conflicts of interest.

References

1. M. Lamy, C. Finot, J. Fatome, J. Arocas, J. C. Weeber, and K. Hammani, “Demonstration of high-speed optical transmission at 2 µm in titanium dioxide waveguides,”Appl. Sci.7(6), 631 (2017).

2. K. Hammani, L. Markey, M. Lamy, B. Kibler, J. Arocas, J. Fatome, A. Dereux, J. C. Weeber, and C. Finot, “Octave Spanning Supercontinuum in Titanium Dioxide Waveguides,”Appl. Sci.8(4), 543 (2018).

3. C. C. Evans, C. Liu, and J. Suntivich, “Low-loss titanium dioxide waveguides and resonators using a dielectric lift-off fabrication process,”Opt. Express23(9), 11160–1169 (2015).

4. C. C. Evans, C. Liu, and J. Suntivich, “TiO2Nanophotonic Sensors for Efficient Integrated Evanescent Raman Spectroscopy,”ACS Photonics3(9), 1662–1669 (2016).

5. X. Guan, H. Hu, L. K. Oxenløwe, and L. H. Frandsen, “Compact titanium dioxide waveguides with high nonlinearity at telecommunication wavelengths,”Opt. Express26(2), 1055–1063 (2018).

6. M. Landmann, E. Rauls, and W. G. Schmidt, “The electronic structure and optical response of rutile, anatase and brookite TiO2,”J. Phys.: Condens. Matter24(19), 195503 (2012).

7. A. Dhakal, A. Raza, F. Peyskens, A. Z. Subramanian, S. Clemmen, N. Le Thomas, and R. Baets, “Efficiency of evanescent excitation and collection of spontaneous Raman scattering near high index contrast channel waveguides,”

Opt. Express23(21), 27391–27404 (2015).

8. R. Ronchin, A. Chiasera, M. Montagna, R. Rolli, C. Tosello, R. R. Goncalves, A. Chiappini, M. Ferrari, L. Zampedri, S. Pelli, G. C. Righini, A. Monteil, V. Foglietti, A. Minotti, R. M. Almeida, A. Marques, and V. Soares, “Erbium-activated silica-titania planar and channel waveguides prepared by rf-sputtering,”Proc. SPIE4990, 38 (2001).

9. W. F. Zhang, Y. L. He, M. S. Zhang, Z. Yin, and Q. Chen, “Raman scattering study on anatase TiO2nanocrystals,”J. Phys. D: Appl. Phys.33(8), 912–916 (2000).

10. T. Kornher, K. Xia, R. Kolesov, B. Villa, S. Lasse, C. S. Sandu, E. Wagner, S. Harada, G. Benvenuti, H. W. Becker, and J. Wrachtrup, “Amorphous Silicon-Doped Titania Films for on-Chip Photonics,”ACS Photonics4(5), 1101–1107 (2017).

11. A. Bahtat, M. Bouderbala, M. Bahtat, M. Bouazaoui, J. Mugnier, and M. Druetta, “Structural characterisation of Er3+doped sol–gel TiO2planar optical waveguides,”Thin Solid Films323(1-2), 59–62 (1998).

12. A. Bahtat, M. Bouazaoui, M. Bahtat, C. Garapon, B. Jacquier, and J. Mugnier, “Up-conversion fluorescence spectroscopy in Er3+:TiO2planar waveguides prepared by a sol-gel process,”J. Non-Cryst. Solids202(1-2), 16–22 (1996).

13. C. C. Evans, K. Shtyrkova, O. Reshef, M. Moebius, J. D. B. Bradley, S. Griesse-Nascimento, E. Ippen, and E. Mazur, “Multimode phase-matched third-harmonic generation in sub-micrometer-wide anatase TiO2waveguides,”Opt. Express23(6), 7832–7841 (2015).

14. C. C. Evans, K. Shtyrkova, J. D. B. Bradley, O. Reshef, E. Ippen, and E. Mazur, “Spectral broadening in anatase titanium dioxide waveguides at telecommunication and near-visible wavelengths,”Opt. Express21(15), 18582 (2013).

15. S. S. Djordjevic, J. Basak, B. Guan, K. Shang, H. F. Liu, S. T. S. Cheung, L. Liao, and S. J. B. Yoo, “CMOS-compatible, athermal silicon ring modulators clad with titanium dioxide,”Opt. Express21(12), 13958 (2013).

(9)

16. O. Reshef, K. Shtyrkova, M. G. Moebius, S. Griesse-Nascimento, S. Spector, C. C. Evans, E. Ippen, and E. Mazur, “Polycrystalline anatase titanium dioxide microring resonators with negative thermo-optic coefficient,”J. Opt. Soc. Am. B32(11), 2288–2293 (2015).

17. B. Guha, J. Cardenas, and M. Lipson, “Athermal silicon microring resonators with titanium oxide cladding,”Opt. Express21(22), 26557–26563 (2013).

18. F. Qiu, A. M. Spring, and S. Yokoyama, “Athermal and high-Q hybrid TiO2-Si3N4ring resonator via an etching-free fabrication technique,”ACS Photonics2(3), 405–409 (2015).

19. J. Bovington, R. Wu, K. T. Cheng, and J. E. Bowers, “Thermal stress implications in athermal TiO2waveguides on a silicon substrate,”Opt. Express22(1), 661–666 (2014).

20. L. He, Y. Guo, Z. Han, K. Wada, L. C. Kimerling, J. Michel, A. M. Agarwal, G. Li, and L. Zhang, “Broadband athermal waveguides and devices for datacom and telecom applications,”Proc. SPIE10537, 64 (2018).

21. M. Furuhashi, M. Fujiwara, T. Ohshiro, M. Tsutsui, K. Matsubara, M. Taniguchi, S. Takeuchi, and T. Kawai, “Development of microfabricated TiO2channel waveguides,”AIP Adv.1(3), 032102 (2011).

22. J. D. B. Bradley, C. C. Evans, J. T. Choy, O. Reshef, P. B. Deotare, F. Parsy, K. C. Phillips, M. Lončar, and E. Mazur, “Submicrometer-wide amorphous and polycrystalline anatase TiO2waveguides for microphotonic devices,”Opt. Express20(21), 23821–23831 (2012).

23. T. Peng, X. Xiao, F. Ren, J. Xu, X. Zhou, F. Mei, and C. Jiang, “Influence of annealing temperature on the properties of TiO2films annealed by ex situ and in situ TEM,”J. Wuhan Univ. Technol., Mater. Sci. Ed.27(6), 1014–1019 (2012).

24. I. Hadjoub, T. Touam, A. Chelouche, M. Atoui, J. Solard, M. Chakaroun, A. Fischer, A. Boudrioua, and L. H. Peng, “Post-deposition annealing effect on RF-sputtered TiO2thin-film properties for photonic applications,”Appl. Phys. A

122(2), 78 (2016).

25. M. H. Suhail, G. Mohan Rao, and S. Mohan, “dc reactive magnetron sputtering of titanium-structural and optical characterization of TiO2films,”J. Appl. Phys.71(3), 1421–1427 (1992).

26. D. Depla, S. Heirwegh, S. Mahieu, J. Haemers, R. De Gryse, D. Depla, S. Heirwegh, S. Mahieu, J. Haemers, and R. De Gryse, “Understanding the discharge voltage behavior during reactive sputtering of oxides Understanding the discharge voltage behavior during reactive sputtering,”J. Appl. Phys.101(1), 013301 (2007).

27. W. D. Sproul, “Summary Abstract: Reactively sputtered TiN, ZrN, and HfN,”J. Vac. Sci. Technol., A3(3), 580–581 (1985).

28. L. Jiang, C. C. Evans, O. Reshef, and E. Mazur, “Optimizing anatase-TiO2deposition for low-loss planar waveguides,”

Proc. SPIE8626, 86261D (2013).

29. S. Boukrouh, R. Bensaha, S. Bourgeois, E. Finot, and M. C. Marco de Lucas, “Reactive direct current magnetron sputtered TiO2thin films with amorphous to crystalline structures,”Thin Solid Films516(18), 6353–6358 (2008). 30. J. M. Bennett, E. Pelletier, G. Albrand, J. P. Borgogno, B. Lazarides, C. K. Carniglia, R. A. Schmell, T. H. Allen,

T. Tuttle-Hart, K. H. Guenther, and A. Saxer, “Comparison of the properties of titanium dioxide films prepared by various techniques,”Appl. Opt.28(16), 3303–3317 (1989).

31. M. Horn and C. F. Schwerdtfeger, “Refinement of the structure of anatase at several temperatures,”Z. Kristallogr.

136(3-4), 273–281 (1972).

32. S. Mahieu, P. Ghekiere, D. Depla, and R. De Gryse, “Biaxial alignment in sputter deposited thin films,”Thin Solid Films515(4), 1229–1249 (2006).

33. X. Ji, F. A. S. Barbosa, S. P. Roberts, A. Dutt, J. Cardenas, Y. Okawachi, A. Bryant, A. L. Gaeta, and M. Lipson, “Ultra-low-loss on-chip resonators with sub-milliwatt parametric oscillation threshold,”Optica4(6), 619 (2017).

Referenties

GERELATEERDE DOCUMENTEN

For the first time in an African population, we have explored the diagnostic utility of fasting plasma glucose or random blood glucose alone and the added value of 50g

Omstreeks die tijd zijn meer geitenhouders begonnen met duurmelken, met goede resultaten maar ook met vragen over de bedrijfsvoering en de voeding.. Op verzoek van de Nevem

• Registreren: Het is zinvol om alle waarnemingen en tellingen te registreren, omdat daarmee de ontwikkeling van ziekten en plagen goed gevolgd kan worden en het resultaat van

It start with religion as sacred text and film, moving to the value of film for religious education, followed by a focus on how the act of watching film creates an opportunity

De NHG geeft aan dat allergietesten in de nieuwe richtlijn alleen voor nog niet ontdekte allergieën gedaan moeten worden en in ieder geval voor verwijzing van de patiënt naar

In light of the increasing importance of the new paradigm of brand co-creation, challenging the traditional company-centric approach to brand management and realising that multiple

Hierin schuilt een interessante les voor betrokke- nen in het Waddengebied: pu- blieke steun als directe voor- waarde, om überhaupt tot een oplossing te komen.

generated by polymerization through the construction of reversible covalent bonds including Diels–Alder linkages, imines, acylhydrazones, oximes, boronate esters, and