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

Dispersion engineering of thick high-Q silicon nitride ring-resonators via atomic layer deposition

Open Access Open Access

Abstract

We demonstrate dispersion engineering of integrated silicon nitride based ring resonators through conformal coating with hafnium dioxide deposited on top of the structures via atomic layer deposition. Both, magnitude and bandwidth of anomalous dispersion can be significantly increased. The results are confirmed by high resolution frequency-comb-assisted-diode-laser spectroscopy and are in very good agreement with the simulated modification of the mode spectrum.

© 2012 Optical Society of America

1. Introduction

Silicon nitride (Si3N4) integrated planar waveguide and ring resonator structures [1] are attractive platforms for resonant nonlinear frequency conversion [2]. Moreover Si3N4 has been used for ultra-low loss integrated waveguides [3], particularly in the optical telecom band, where well established silicon on insulator waveguides [4, 5] suffer from two-photon and free carrier absorption. Besides low absorption, the waveguide dispersion plays a central role for parametric frequency conversion. In particular this applies to microresonator based optical frequency comb generation via the χ(3) non-linearity (”Kerr-combs”) [6], which has recently also been demonstrated in Si3N4 [7]. In this scheme a set of equidistant optical frequencies is generated with a spacing corresponding to the free spectral range (FSR) of the resonator. Such integrated and CMOS-compatible microresonator frequency combs potentially offer a level of compactness and integration that is presently not attainable in mode-locked laser based combs and can moreover access GHz repetition rates [8]. Microresonator based frequency combs are promising for applications like the calibration of astronomical spectrographs [9], on chip optical interconnects [2], long range data communication [10] and optical arbitrary waveform generation [11]. Moreover recently Kerr-comb generation has recently been demonstrated in the visible [12] and the mid-infrared [13]. It is well known that the attainable spectral bandwidth of Kerr-combs is limited by the resonator dispersion, that leads to a wavelength dependent FSR of the resonator [14, 15]. The resonator dispersion is described as

D2=ω12ω+ω+1,
where ω is the frequency of the fundamental resonance with azimuthal mode number . The dispersion D2 in microresonators is related to the group velocity dispersion of the structures via [16]
β2=D22πRD13,
with D1 = ω+1ω being the FSR of the resonator and R the ring radius. It has been shown that the phase noise characteristics of Kerr-comb generators are related to the ratio of cavity-decay-rate κ and dispersion [17] and a sufficiently large anomalous dispersion is advantageous for low phase noise operation of Kerr-combs (intrinsically low phase noise combs are generated for κ/D2~1). Dispersion in integrated ring resonators is composed of material dispersion (which is normal for Si3N4 at all visible and near-infrared wavelengths), geometric dispersion due to the waveguide cross-section [18], as well as an additional contribution from the finite resonator radius as observed in whispering-gallery-mode microcavities [19]. In previous numerical studies it has been shown that silicon (n = 3.5) waveguide dispersion can be decreased and flattened through a conformal coating with Si3N4 (n = 1.99) [20]. Here we demonstrate both experimentally and numerically dispersion engineering, that is a targeted increase of magnitude and bandwidth of anomalous dispersion in integrated Si3N4 ring resonators through conformal coating of hafnium dioxide (HfO2 n = 2.03). This material combination has the advantage of very small scattering losses at the HfO2/Si3N4 interface. Furthermore, the total scattering loss of the devices could potentially be reduced because the surface roughness of Si3N4 is masked by the HfO2 coating [21].

2. Fabrication

The resonators and coupling waveguides are fabricated as follows. First, a bottom cladding layer of 4 μm thick silicon dioxide (SiO2) is grown onto a 100 mm prime silicon wafer via wet oxidation. The thickness of the bottom cladding is chosen such that the compressive stress of the SiO2 cancels out the tensile stress of the Si3N4. A 750 nm thick stoichiometric Si3N4 layer is grown with low-pressure-chemical-vapor-deposition (LPCVD) from dichlorsilane (SiH2CL2, 30 sccm) and ammonia (NH3, 180 sccm). The Si3N4 growth rate is 2.75 nm/min at a temperature of 775°. To avoid excessive stress buildup and successive cracking of the Si3N4 film, the deposition is carried out in two equal deposition steps with intermediate cooldown to room temperature [22]. The dielectrics are etched from the backside of the wafer via reactive-ion-etching (RIE) to avoid buildup of static charges during the following electron beam lithography step. After electron beam lithography (using positive ZEP520A resist), the waveguide and resonator patterns are transferred into the waveguide core layer via RIE in SF6 and CH4 chemistry. The etch parameters have been carefully optimized to achieve low roughness while maintaining verticality and selectivity. The sidewall angle is approximately 78°. The ZEP520A is then stripped in an oxygen Plasma. Large remaining Si3N4 patches, which can lead to cracking during subsequent thermal annealing, are removed utilizing photolithography and an additional RIE step. After cleaning using Piranha etch the waveguides and resonators are thermally annealed for 3 hours in a nitrogen atmosphere at a temperature of 1200 °C to eliminate residual hydrogen from the Si3N4 structures and to reduce optical absorption in the near-infrared [1]. HfO2 with a thickness of 55 nm is grown on top of the waveguides and resonators by atomic layer deposition (ALD) from tetrakis(ethylmethylamino)hafnium (TEMAH) and water (c.f. Fig. 1) at 200 °C. The low deposition temperature in ALD is beneficial due to the reduced thermal stress on the thick Si3N4 structures. Finally, 3 μm thick low-temperature silicon dioxide (SiO2) is deposited on top on the structures. The critically coupled optical quality factors Q of the resonators are up to 1 million with HfO2 clad resonators having slightly smaller quality factors (up to Q ≈ 6 · 105) due to the formation of small irregularities on the surface (c.f. Fig. 1) during deposition. These are expected to vanish with further optimized ALD parameters and wafer preparation. The circumference of the chips is defined via an additional photolithography step and all dielectrics are etched with a single RIE step. To facilitate cleaving of the wafer and clear space to approach the coupling facet with tapered-lensed fibers for input and output coupling of light, a 130 μm deep anisotropic silicon RIE etch is performed. To reduce losses from fiber-to-chip coupling the bus waveguides are tapered [23] to a width of 150 nm and terminated 6 μm away from the etched coupling facet.

 figure: Fig. 1

Fig. 1 Scanning electron micrograph of the cross-section of a HfO2 (green) coated Si3N4 (yellow) resonator before the SiO2 top cladding deposition. The Si3N4 core has a base width of 1700 nm and a height of 750 nm. The HfO2 coating thickness is 55 nm. The detachment of the HfO2 film on the lower right side of the resonator was caused by the cleaving process. Inset: Optical micrograph of a 50 μm radius ring resonator and coupling waveguide.

Download Full Size | PDF

3. Experiment

The resonator dispersion is characterized employing frequency-comb-assisted-diode-laser spectroscopy [19] (c.f. Fig. 2a). Here, a mode-hop-free external cavity diode laser (ECDL) is scanned between 1545 nm and 1615 nm. The spectrum of the resonator in transmission is recorded while a beat note with a self-referenced, stabilized fiber-laser frequency comb (Menlo Systems GmbH) with repetition rate of 250 MHz is generated. The beat signal is split and bandpass filtered at 30 MHz and 75 MHz to create two sets of accurate frequency calibration markers. All traces are recorded on a fast oscilloscope in peak detection mode with 10 million sampling points. To extend the span of the measurement the frequency comb is amplified in a dispersion compensated erbium-doped-fiber amplifier (Menlo Systems GmbH) and broadened in SMF-28. The intensity of the scanning laser is actively stabilized using a Mach-Zehnder interferometric intensity modulator. The measurement span is limited by the mode-hop free scanning range of the ECDL and the signal-to-noise ratio of the beat note generation.

 figure: Fig. 2

Fig. 2 (a) Setup for frequency-comb-assisted-diode-laser spectroscopy; FPC, fiber polarization controller; MZI, Mach-Zehnder intensity modulator; BP, bandpass; DC-EDFA, dispersion compensated erbium-doped-fiber-amplifier. (b) Calibrated transmission spectrum of a 55 nm ALD-coated ring resonator with radius 301 μm and input polarization aligned to the TM11 mode. (c) Wavelength dependence of the free-spectral-range (FSR) over the measurement span of 8.7 THz, corresponding to 120 subsequent FSRs, based on 16 consecutive scans (symbols) in alternating direction. The red line is a linear fit from which both FSR D1/2π of 75.04 GHz and dispersion D2/2π of 350 kHz are determined.

Download Full Size | PDF

The resonance positions are fitted with symmetric double Lorentzian functions to account for modal coupling of counter-propagating modes [24]. The cross-section of the ring resonators are designed to feature anomalous dispersion at 1.55 μm and support both fundamental TE(TM)11 and first higher order TE(TM)21 modes. While in larger resonators (e.g. 301 micron radius) the excitation of higher transverse order modes is suppressed by the phase matching condition, these modes can be excited in smaller resonator geometries. Different mode families are identified via their FSR (c.f. Figs. 3a and 3b). These different FSR lead to periodic crossings of mode families [19] and to the formation of hybrid modes due to modal coupling [25]. These effects are observable as spectral variation of the resonance depth (c.f. Fig. 2b). Data points strongly affected by modal coupling to higher order modes are discarded before the dispersion D2 is extracted from a linear fit of the measured FSR over the full spectral span (c.f. Fig. 2c) [19]. Another systematic error arises from wavelength dependent waveguide transmission due to interference of higher order and fundamental waveguide modes and asymmetric resonance line shapes [26]. Each dispersion measurement corresponds to 16 scans in alternating scan direction to cancel thermal shifts of resonance positions [27] and reduce the statistical error from calibration and double-Lorentzian fitting. Multiple measurement points in Fig. 3 and Fig. 4 correspond to different resonators with different coupling waveguide separation.

 figure: Fig. 3

Fig. 3 (a) Simulations (lines) and measurements (symbols) of FSR D1/2π and dispersion D2/2π as function of wavelength of TE11 (a,c) and TM11 (b,d) modes for different HfO2 ALD coating thickness applied to 301 μm Si3N4 ring resonators with base width of 1700 nm and Si3N4 height of 750 nm. (e) Dispersion D1/2π at 1.55 μm as function of HFO2 coating thickness and Si3N4 thickness for TE11 mode. The black solid line marks the transition from normal to anomalous dispersion. (f) Same as (e) for TM11 mode.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Simulated resonator dispersion D2/2π at 1.55 μm for the TE11(a) and TM11(b) mode as function of ring radius and ring base width for Si3N4 ring resonators without HfO2 coating and Si3N4 height of 750 nm. The black solid line marks the transition from normal to anomalous dispersion. (c) Comparison of dispersion simulations (lines) and measurements (symbols) for TE11 and TM11 mode for 24 μm, 37 μm and 50 μm ring resonators with a base width of 1700 nm at wavelength 1.58 μm. Dashed lines show the resonator dispersion D2 if constant β2 is assumed (c.f. Eq. (2)). Deviations from simulations and measurements are due to additional normal dispersion introduced by the finite radius of the ring resonators. (d,e,f) Same as (a,b,c) but with 55 nm HfO2 coating.

Download Full Size | PDF

4. Simulations and discussion

We simulate the dispersion of integrated Si3N4 based ring resonators with a fully-vectorial axially-symmetric finite-element model [28] employing a commercial FEM solver (COMSOL Multiphysics). The model includes the angled sidewalls as well as the shape of the conformal coating on top of the waveguide edges (c.f. Fig. 1). The material dispersion is determined via spectral ellipsometry and included iteratively (similar to Ref. [19]) into the simulations. The resonator dispersion D2 directly follows from the calculated mode resonance frequencies. The results of the simulations for varied HfO2 coating thickness are shown in Fig. 3 and we find that both the magnitude and bandwidth of anomalous dispersion can be significantly increased. While the low wavelength zero-dispersion point (ZDP) is only weakly but quadratically shifted (TE11: 0.0021 nm/nm2, TM11: 0.0067 nm/nm2), the long wavelength ZDP strongly linearly shifts towards mid-infrared wavelength for increased HfO2 thickness (TE11: 26 nm/10 nm, TM11: 27 nm/10 nm). The simulations are compared to measurements of 301 μm ring resonators shown in Fig. 2c. We measure an increase of anomalous dispersion D2/2π at 1.58 μm from 340 kHz to 490 kHz for the TE11 mode for a coating thickness of 55 nm. The TM11 mode, i.e. the mode with the major electric field component out of the resonator plane, is more strongly affected by the coating. We measure that ring resonators without HfO2 coating have normal dispersion (−180 kHz) while similar resonators coated with 55 nm HfO2 feature anomalous dispersion (350 kHz). Figures 3e and 3f show simulations of the 301 μm ring resonator dispersion D2/2π at 1.55 μm as function of both HfO2 coating thickness and Si3N4 core thickness. To decouple the HfO2 coating thickness on the sidewalls and the base width of the resonator in Figs. 3e and 3f, the base width of the resonators is chosen such that the total base width after coating is 1700 nm. The interpretation of the increase of anomalous dispersion upon coating is as follows: For thin HfO2 coatings, with only a few percent of the optical power contained in the HfO2 coating, the increase of anomalous dispersion and anomalous dispersion bandwidth is mainly due to the increased waveguide cross section [29] of the now composite Si3N4/HfO2 resonator core. Comparable results could thus also be achieved utilizing thicker Si3N4 layers, which however are very difficult to fabricate due to the high intrinsic stress of Si3N4. For thick HfO2 coatings the material dispersion of HfO2, which has a zero dispersion point around 1.8 μm, more strongly influences the ring resonator dispersion especially towards mid-infrared wavelengths. Also, from both the quadratic dependence of the low wavelength ZDP as well as Fig. 3e, it is apparent that the formation of a HfO2 slab waveguide on the wafer introduces additional normal dispersion especially towards shorter wavelengths. Note that increased anomalous dispersion for a Si3N4 ring resonator of fixed thickness could also be achieved through partial or full removal of the SiO2 top cladding, which however would sacrifice full encapsulation of the device. Coating with HfO2 facilitates large and broadband anomalous dispersion without aforementioned drawbacks.

For a constant group-velocity-dispersion β2 it is expected that the total dispersion scales as D2/2π ∝ 1/R2 (c.f. Eq. (2)). Thus, the ring radius not only determines the FSR D1/2π of the resonator but is also crucial for obtaining large values of resonator dispersion D2/2π. For whispering-gallery-mode resonators it has been observed that smaller wavelengths travel closer to the outer rim of the resonator and additional normal dispersion is introduced even for large radii and negligible bending loss [19]. Figure 4 presents measurements and simulations, which show that this effect also applies for ring resonators with strong radial confinement, such as integrated waveguide rings, and its alteration for resonators coated with HfO2. Similar effects have been numerically studied for bent silicon waveguides [30]. In Figs. 4a and 4b, simulations of the dispersion D2/2π as function of waveguide width and ring radius are presented.

We find a maximum of anomalous dispersion as function of ring radius, due to the interaction of the ∝ 1/R2 scaling and the additional normal dispersion, and ring base width. The decrease toward small base width is due to decreased anomalous waveguide dispersion [18]. The decrease toward large base width is due to the transformation of the waveguide like resonator mode to a disc like resonator mode. We observe that coating with HfO2 leads to a strongly increased dispersion which is shifted towards narrower ring base widths and smaller ring radii. The shift towards smaller radii is due to the increased confinement and supplements the increased anomalous dispersion (c.f. Eq. (2)). The shift towards smaller base width is due to the 110 nm HfO2 on the waveguide sidewalls (TE11) as well as due to the increased confinement (TM11). The predictions of the numerical simulations are confirmed by measurements on ring resonators with a Si3N4 base width of 1700 nm and ring radii of 24 μm, 37 μm and 50 μm (c.f. Figs. 4c and 4f).

5. Conclusion

In summary, we have demonstrated both numerically and experimentally that dispersion of Si3N4 based waveguides can be engineered to be more anomalous through conformal coating with HfO2 which is in particular favorable for applications such as low phase noise Kerr-comb generation. To facilitate measurements, we have extended the measurement span of frequency comb assisted diode laser spectroscopy to more than 8 THz maintaining MHz resolution and accuracy to resolve the dispersion of integrated ring resonators (values as small as x100 below the cavity decay rate can be measured). Using this unique method of directly measuring the dispersion of traveling wave microresonators with large FSR we have been able to demonstrate that the total Si3N4-ring resonator dispersion for small ring radii is always normal. Our results also show how HfO2 coating can be used to extend the difficult fabrication of thick Si3N4 based waveguide devices effectively reducing the necessary thickness of high stress, low loss stochiometric Si3N4 layers. The low process temperature of ALD and the high conformity should make ALD a versatile tool in microresonator fabrication and also be compatible to a variety of materials. This includes materials with high intrinsic stress such as Si3N4 and with low softening or melting points such as chalcogenide glasses or even polymers. The large transparency window of HfO2 is especially promising for applications in the mid-infrared wavelength region.

Acknowledgments

This work was supported by the NanoTera program of the SNF under the acronym micro-Comb, by DARPA under the QuASAR program and by the Swiss National Science Foundation (SNF). K. H. acknowledges support under a Marie Curie IAPP with Menlo Systems GmbH. The samples used in this work were fabricated at the Center for Microfabrication (CMi) at EPFL.

References and links

1. C. H. Henry, R. F. Kazarinov, H. J. Lee, K. J. Orlowsky, and L. E. Katz, “Low loss Si3N4-SiO2 optical waveguides on Si,” Appl. Opt. 26, 2621–2624 (1987). [CrossRef]   [PubMed]  

2. J. S. Levy, A. Gondarenko, M. A. Foster, A. C. Turner-Foster, A. L. Gaeta, and M. Lipson, “CMOS-compatible multiple-wavelength oscillator for on-chip optical interconnects,” Nat. Photonics 4, 37–40 (2010). [CrossRef]  

3. J. F. Bauters, M. J. R. Heck, D. D. John, J. S. Barton, C. M. Bruinink, A. Leinse, R. G. Heideman, D. J. Blumenthal, and J. E. Bowers, “Planar waveguides with less than 0.1 dB/m propagation loss fabricated with wafer bonding,” Opt. Express 19, 24090–24101 (2011). [CrossRef]   [PubMed]  

4. M. A. Foster, A. C. Turner, J. E. Sharping, B. S. Schmidt, M. Lipson, and A. L. Gaeta, “Broad-band optical parametric gain on a silicon photonic chip,” Nature 441, 960–963 (2006). [CrossRef]   [PubMed]  

5. C. Koos, L. Jacome, C. Poulton, J. Leuthold, and W. Freude, “Nonlinear silicon-on-insulator waveguides for all-optical signal processing,” Opt. Express 15, 5976–5990 (2007). [CrossRef]   [PubMed]  

6. T. J. Kippenberg, R. Holzwarth, and S. A. Diddams, “Microresonator-based optical frequency combs,” Science 332, 555–559 (2011). [CrossRef]   [PubMed]  

7. M. A. Foster, J. S. Levy, O. Kuzucu, K. Saha, M. Lipson, and A. L. Gaeta, “Silicon-based monolithic optical frequency comb source,” Opt. Express 19, 14233–14239 (2011). [CrossRef]   [PubMed]  

8. P. Del‘Haye, O. Arcizet, A. Schliesser, R. Holzwarth, and T. J. Kippenberg, “Full stabilization of a microresonator-based optical frequency comb,” Phys. Rev. Lett. 101, 053903 (2008). [CrossRef]  

9. T. Steinmetz, T. Wilken, C. Araujo-Hauck, R. Holzwarth, T. W. Hänsch, L. Pasquini, A. Manescau, S. D’Odorico, M. T. Murphy, T. Kentischer, W. Schmidt, and T. Udem, “Laser frequency combs for astronomical observations,” Science 321, 1335–1337 (2008). [CrossRef]   [PubMed]  

10. J. Pfeifle, C. Weimann, F. Bach, J. Riemensberger, K. Hartinger, D. Hillerkuss, M. Jordan, R. Holtzwarth, T. J. Kippenberg, J. Leuthold, W. Freude, and C. Koos, “Microresonator-based optical frequency combs for high-bitrate WDM data transmission,” in Optical Fiber Communication Conference (OFC) (Optical Society of America, 2012), paper OW1C.4.

11. F. Ferdous, H. Miao, D. E. Leaird, K. Srinivasan, J. Wang, L. Chen, L. T. Varghese, and A. M. Weiner, “Spectral line-by-line pulse shaping of on-chip microresonator frequency combs,” Nat. Photonics 5, 770–776 (2011). [CrossRef]  

12. A. A. Savchenkov, A. B. Matsko, W. Liang, V. S. Ilchenko, D. Seidel, and L. Maleki, “Kerr combs with selectable central frequency,” Nat. Photonics 5, 293–296 (2011). [CrossRef]  

13. C. Y. Wang, T. Herr, P. Del’Haye, A. Schliesser, J. Hofer, R. Holzwarth, T. W. Hänsch, N. Picque, and T. J. Kippenberg, “Mid-infrared optical frequency combs based on crystalline microresonators,” arXiv:1109.2716 (2011).

14. P. Del’Haye, T. Herr, E. Gavartin, M. L. Gorodetsky, R. Holzwarth, and T. J. Kippenberg, “Octave spanning tunable frequency comb from a microresonator,” Phys. Rev. Lett. 107, 063901 (2011). [CrossRef]  

15. Y. Okawachi, K. Saha, J. S. Levy, Y. H. Wen, M. Lipson, and A. L. Gaeta, “Octave-spanning frequency comb generation in a silicon nitride chip,” Opt. Lett. 36, 3398–3400 (2011). [CrossRef]   [PubMed]  

16. A. B. Matsko, A. A. Savchenkov, W. Liang, V. S. Ilchenko, D. Seidel, and L. Maleki, “Mode-locked Kerr frequency combs,” Opt. Lett. 36, 2845–2847 (2011). [CrossRef]   [PubMed]  

17. T. Herr, K. Hartinger, J. Riemensberger, C. Wang, E. Gavartin, R. Holzwarth, M. L. Gorodetsky, and T. J. Kippenberg, “Universal formation dynamics and noise of Kerr-frequency combs in microresonators,” Nat. Photonics 6, 480–487 (2012). [CrossRef]  

18. A. C. Turner, C. Manolatou, B. S. Schmidt, and M. Lipson, “Tailored anomalous group-velocity dispersion in silicon channel waveguides,” Opt. Express 14, 4357–4362 (2006). [CrossRef]   [PubMed]  

19. P. Del’Haye, O. Arcizet, M. Gorodetsky, R. Holzwarth, and T. J. Kippenberg, “Frequency comb assisted diode laser spectroscopy for measurement of microcavity dispersion,” Nat. Photonics 3, 529–533 (2009). [CrossRef]  

20. X. Liu, W. M. J. Green, X. Chen, I.-W. Hsieh, J. I. Dadap, Y. A. Vlasov, and R. M. Osgood, “Conformal dielectric overlayers for engineering dispersion and effective nonlinearity of silicon nanophotonic wires,” Opt. Lett. 33, 2889–2891 (2008). [CrossRef]   [PubMed]  

21. T. Alasaarela, D. Korn, L. Alloatti, A. Säynätjoki, A. Tervonen, R. Palmer, J. Leuthold, W. Freude, and S. Honkanen, “Reduced propagation loss in silicon strip and slot waveguides coated by atomic layer deposition,” Opt. Express 19, 11529–11538 (2011). [CrossRef]   [PubMed]  

22. A. Gondarenko, J. S. Levy, and M. Lipson, “High confinement micron-scale silicon nitride high-Q ring resonator,” Opt. Express 17, 11366–11370 (2009). [CrossRef]   [PubMed]  

23. V. R. Almeida, R. R. Panepucci, and M. Lipson, “Nanotaper for compact mode conversion,” Opt. Lett. 28, 1302–1304 (2003). [CrossRef]   [PubMed]  

24. T. J. Kippenberg, S. M. Spillane, and K. J. Vahala, “Modal coupling in traveling-wave resonators,” Opt. Lett. 27, 1669–1671 (2002). [CrossRef]  

25. T. Carmon, H. G. L. Schwefel, L. Yang, M. Oxborrow, A. D. Stone, and K. J. Vahala, “Static envelope patterns in composite resonances generated by level crossing in optical toroidal microcavities,” Phys. Rev. Lett. 100, 103905 (2008). [CrossRef]   [PubMed]  

26. A. Chiba, H. Fujiwara, J. I. Hotta, S. Takeuchi, and K. Sasaki, “Fano resonance in a multimode tapered fiber coupled with a microspherical cavity,” Appl. Phys. Lett. 86, 261106 (2005). [CrossRef]  

27. T. Carmon, L. Yang, and K. Vahala, “Dynamical thermal behavior and thermal self-stability of microcavities,” Opt. Express 12, 4742–4750 (2004). [CrossRef]   [PubMed]  

28. M. Oxborrow, “Traceable 2-D finite-element simulation of the whispering-gallery modes of axisymmetric electromagnetic resonators,” IEEE Trans. Microwave Theory Tech. 55, 1209–1218 (2007). [CrossRef]  

29. D. T. H. Tan, K. Ikeda, P. C. Sun, and Y. Fainman, “Group velocity dispersion and self phase modulation in silicon nitride waveguides,” Appl. Phys. Lett. 96, 061101 (2010). [CrossRef]  

30. L. Zhang, Y. Yue, R G. Beausoleil, and A E. Willner, “Analysis and engineering of chromatic dispersion in silicon waveguide bends and ring resonators,” Opt. Express 19, 8102–8107 (2011). [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 (4)

Fig. 1
Fig. 1 Scanning electron micrograph of the cross-section of a HfO2 (green) coated Si3N4 (yellow) resonator before the SiO2 top cladding deposition. The Si3N4 core has a base width of 1700 nm and a height of 750 nm. The HfO2 coating thickness is 55 nm. The detachment of the HfO2 film on the lower right side of the resonator was caused by the cleaving process. Inset: Optical micrograph of a 50 μm radius ring resonator and coupling waveguide.
Fig. 2
Fig. 2 (a) Setup for frequency-comb-assisted-diode-laser spectroscopy; FPC, fiber polarization controller; MZI, Mach-Zehnder intensity modulator; BP, bandpass; DC-EDFA, dispersion compensated erbium-doped-fiber-amplifier. (b) Calibrated transmission spectrum of a 55 nm ALD-coated ring resonator with radius 301 μm and input polarization aligned to the TM11 mode. (c) Wavelength dependence of the free-spectral-range (FSR) over the measurement span of 8.7 THz, corresponding to 120 subsequent FSRs, based on 16 consecutive scans (symbols) in alternating direction. The red line is a linear fit from which both FSR D1/2π of 75.04 GHz and dispersion D2/2π of 350 kHz are determined.
Fig. 3
Fig. 3 (a) Simulations (lines) and measurements (symbols) of FSR D1/2π and dispersion D2/2π as function of wavelength of TE11 (a,c) and TM11 (b,d) modes for different HfO2 ALD coating thickness applied to 301 μm Si3N4 ring resonators with base width of 1700 nm and Si3N4 height of 750 nm. (e) Dispersion D1/2π at 1.55 μm as function of HFO2 coating thickness and Si3N4 thickness for TE11 mode. The black solid line marks the transition from normal to anomalous dispersion. (f) Same as (e) for TM11 mode.
Fig. 4
Fig. 4 Simulated resonator dispersion D2/2π at 1.55 μm for the TE11(a) and TM11(b) mode as function of ring radius and ring base width for Si3N4 ring resonators without HfO2 coating and Si3N4 height of 750 nm. The black solid line marks the transition from normal to anomalous dispersion. (c) Comparison of dispersion simulations (lines) and measurements (symbols) for TE11 and TM11 mode for 24 μm, 37 μm and 50 μm ring resonators with a base width of 1700 nm at wavelength 1.58 μm. Dashed lines show the resonator dispersion D2 if constant β2 is assumed (c.f. Eq. (2)). Deviations from simulations and measurements are due to additional normal dispersion introduced by the finite radius of the ring resonators. (d,e,f) Same as (a,b,c) but with 55 nm HfO2 coating.

Equations (2)

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

D 2 = ω 1 2 ω + ω + 1 ,
β 2 = D 2 2 π R D 1 3 ,
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.