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

Fully embedded photonic crystal cavity with Q=0.6 million fabricated within a full-process CMOS multiproject wafer

Open Access Open Access

Abstract

A two dimensional photonic crystal (PhC) resonator, based on a recent design concept, entirely embedded in Silica, is fabricated in a CMOS full-process multiproject wafer, including additional steps such as implantation, metalization, Germanium deposition and planarization. A large loaded Q-factor (5.9 × 105) is achieved without removal of the silica cladding. A statistical analysis over 56 devices leads to an average value for the loaded Q of 4 × 105, in close agreement with calculations. An upper boundary for the fabrication disorder is estimated to 1.2 nm.

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

1. Introduction

Optical resonators, as well as waveguides, interferometers or splitters, are essential ingredients of photonic integrated circuits (PIC). As they allow spectral filtering and wavelength selectivity, they improve the response of modulators and photodetectors and provide the feedback for lasing. The enhancement of the light-matter interaction and, therefore, energy efficiency, is related to the spatial distribution of the electric field and to the quality factor of the cavity, rather than to the physical size of optical cavity. Ultra-low power miniaturized devices such as lasers [1], optical random access memories (RAM) [2] have achieved femtoJoule/microWatt energy/power levels, owing to high-Q small modal volume optical resonators based on two-dimensional Photonic Crystal (PhC) [3]. Moreover, enhanced light-matter interaction allowed by this kind of resonator is the centerpiece of breakthrough research in the field of quantum optics [4], atom optics [5] and also optomechanics [6].

The continuous improvement of both design and fabrication [7] has led to the achievement of a very large Q-factor (> 107) in wavelength scale mode volumes (V ≈ (λ/n)3) [8]. This combination of large Q-factor and small mode volume is unique, clearly motivating the implementation in a standard photonic foundry, specifically CMOS, in the perspective of the photonic-electronic convergence. The challenge here is the resolution of the lithographic process, compared to the electron-beam writer, which is known to be crucial in sub-wavelength photonic structures [9]. Nonetheless, high-Q (2 × 106) PhC resonators have been reported [10], still based on a suspended membrane with a 2μm air gap underneath. This involves an additional processing step after the CMOS fabrication consisting in removing the Buried Oxide (BOX) through chemical etching. It is however desirable to keep the photonic layer protected by silica, more precisely the planarization layer allowing stacking additional elements, the full CMOS process, or also allowing the hybrid integration of other materials, e.g. active III–V semiconductors. Keeping a truly all-solid cladding structure, without any void, is challenging, but experimental Q-factors ranging between 1 × 105 and 1 million have been achieved [11–14]. We point out that there the e-beam lithography has been used. Full solid cladding PhC cavities fabricated in a CMOS foundry have been demonstrated, with Q-factors about 4.5 × 104 [15] for 1D cavities and 2.2 × 105 [16] for 2D cavities, without removing BOX, meaning that they can be readily integrated into larger photonic circuits.

In this paper we demonstrate a high-Q PhC resonator fabricated through the standard full-process CMOS fabrication line at IMEC Leuven (Belgium). Our devices have been included on the same mask than other functional silicon Photonic Integrated Circuits (PIC) and fabricated within a MultiProject Wafer (MPW) run. Despite numerous additional fabrication steps required by MPW users compared to a PhC-only run, the largest Q-factor (5.9×105) achieved demonstrates the role of the cavity design. A statistical analysis over 56 resonators has been carried out leading to estimates of the disorder induced losses, which are believed to remain the dominant limitation to the Q-factor.

2. Design and fabrication

The Photonic Integrated Circuit containing the PhC cavities (Fig. 1(a)) has been fabricated by IMEC using the iSiPP25G pilot line, within a full-process multi-project wafer run. A 193 nm deep immersion lithography is used to pattern a 220 nm thick silicon slab on a 2 μm thick BOX layer. The wafer also included a variety of photonic devices, both active (Germanium photodiodes, thermo-optic switches, ...) and passive (shallow and deep etched waveguides, MMIs, ...). The full process used here consisted in three etch steps: a complete 220 nm etch, a partial 70 nm etch and another partial 160 nm etch. In our case the deep etch has been used, except for the grating couplers which have been partially etched. As part of a multi-project wafer run, the process flow also included implantation steps, germanium (Ge) growth, planarization, metal deposition and oxide cladding.

 figure: Fig. 1

Fig. 1 a) Optical image of the chip corresponding to our user area within the multi-project wafer. The 8 PhC Cavities are located within the red rectangle. b) Optical image of one resonator, with the complete PhC area delimited by the blue line. Deep-etched ridge waveguides are used to connect to grating couplers. c) 3D representation of the PhC.

Download Full Size | PDF

The planarizing step is very important. It is normally used to isolate PIC devices from back-end steps, namely metalization levels (vias, PADs, ...). The PhC cavitiy (Fig. 1(b)) has been designed to comply as much as possible with the rules defined for this process. Under these conditions, it is very likely that the holes are entirely filled with Silicon Oxide, which is a necessary condition for achieving high Q-factor, and therefore the PhC structure is as represented in Fig. 1(c).

Embedding a PhC structure in a solid cladding has several implications, which generally tend to increase internal losses, mainly radiative leakage. Radiative leakage is understood when representing the distribution of the field of the mode in the reciprocal space [17] and it is associated with waves with in-plane wavevector |k//| < ωnclad/c (see Fig. 2(b)). The issue with solid cladding is that a larger fraction of the field is coupled to radiation and therefore radiative losses increase. This can be reformulated as favored leakage by the reduced index contrast. A practical guideline to improve the radiation-limited Q-factor is to avoid as much as possible abrupt changes in the spatial distribution of the field by introducing a suitable tapering of the PhC [18]. It has been shown that tapered1D periodic structures are very effective in this respect, in particular when the index contrast is reduced, such as in diamond PhCs [19] or with all-solid cladding resonators [20].

 figure: Fig. 2

Fig. 2 a) Design of a PhC using a bichromatic lattice. In black is the main triangular PhC lattice of period a. In red is the cavity, formed by a missing line defect surrounded by 2 rows with a slightly different period a′ = 0.98a. In blue are the 2 coupling waveguides (same parameter a as the main lattice). b) Representation of the fundamental mode in the reciprocal space, with the circle corresponding to |k//| < ωnclad/c. c) Modal distribution of the electric field (Ey) for the fundamental mode.

Download Full Size | PDF

Here we rather focus on the design of the cavity itself, which is shown on Fig. 2(a). The main idea is to implement an effective photonic potential through a combination of two photonic lattices with slightly different periods [21]. It has been shown theoretically that an arbitrarily large radiation-limited Q-factor can be achieved, yet with a much simpler design. We note that the principle of bichromatic PhC cavities has already been implemented in Silicon suspended membranes [22] and more recently in III–V suspended alloys [23]. Here we use the same cavity design as we used in [23], which is a variation of the original design proposed in [21]. Three parameters are in fact enough to describe the structure: the periods of the two lattices a and a′ (see Fig. 2(a)), as well as the radius of the holes r. Here the set of parameters is a = 405 nm, a′ = 0.98a = 397 nm and r = 134 nm. The cavity is formed by a 35 periods long missing line defect (see Fig. 2(a)) in the main triangular lattice. The photonic potential is obtained here by changing the period a′ = 0.98a of the first row of holes next to the missing line defect. The cavity is then evanescently coupled to two PhC waveguides located symmetrically with respect to the cavity center. The PhC waveguides terminate into a wire waveguide [24] and are finally connected through grating couplers to standard single-mode fibers. This PhC design has been adapted from the suspended InGaP structure in [23] to the embedded Silicon structure by merely adjusting the period and the size of the holes and it has been integrated directly on GDS mask files of the main project.

Using an in-house implementation of the FDTD method with grid resolution a/20 = 20 nm and subpixel smoothing, we have calculated the intrinsic Q-factor (radiation limited) for an uncoupled resonator, and found that even for our all solid-cladding structures Q is as large as 1.4 × 107 for the fundamental mode and 6 × 106 for the second order mode. Such large Q denotes minimized out of plane scattering, which is consistent with the mode distribution in the reciprocal space (Fig. 2(b)) [17]. As shown on Fig. 1(a), eight sets of cavities have been fabricated, with varying holes diameter, cavity length and waveguide coupling strength. However only two sets of cavities exhibited measurable resonances, only differing by the waveguide to cavity distance, respectively m = 6 and m = 7 lattice periods along the M direction (see Fig. 2(a)) (Other parameters are identical and mentioned above). We will refer to these two geometries as “strong” and “weak coupling” respectively. The corresponding calculated loaded Q-factor is 3.5 × 105 for coupling strength m = 6 and 6 × 105 for m = 7, for the fundamental (Fig. 2(c)) mode. With these parameters the expected resonance of the fundamental mode is at 194 THz (1545 nm) and the second order mode is separated by 420 GHz. In the CMOS MPW process, the unit cell (i.e. Fig. 1(a)) has been replicated resulting into 28 nominally identical dies, hence 56 individual measurable resonators.

3. Results

The resonances are detected by analyzing the complex transmission spectrum obtained through coherent detection, namely Optical Coherent Tomography (OCT) based on a sweep laser source (Santec TSL 510). We have shown that our implementation of this technique provides a spectral resolution of about 20 MHz [23]. The temperature of the sample is stabilized to 25 °C with a controller providing 0.01 °C accuracy. On Fig. 3(a) we report the narrowest lineshape measured, namely 330 MHz, corresponding to Qexp = 5.9 × 105. We note that this result exceeds previous reports on all-solid cladding CMOS cavities [14–16] by a factor of 2 to 3.

 figure: Fig. 3

Fig. 3 a) Measured transmission (black dots) using a coherent detection technique (OCT) revealing the spectral lineshape of the fundamental order mode of the PhC cavity along with a Lorentzian fit. b) temperature dependence of the frequency of the first two modes, and their spectral separation Δν = ν1ν0.

Download Full Size | PDF

The resonance follows the expected dependence on the temperature (Fig. 3(b)). Interestingly, we detect a temperature dependence of the frequency separation (the free spectral range FSR). The measurement error is estimated from the deviation from the linear fit δ ≈ 30MHz, which is consistent with the expected accuracy of our setup. From the same measurements we extract the uncertainty for the Q-factor ΔQ = 3 × 104 (hence 5%).

All the 28 dies and 56 resonators (all functional) have been characterized and the results have been analyzed statistically. For each resonator we consider the first two resonances. The distribution of the values of the Free Spectral Range (FSR, ν1ν0), shown in Fig. 4(a), is centered at 460 GHz and is approximately following a Gaussian law, as it would be expected if the source of the fluctuations were a random change of the structure, namely because of the surface roughness due to fabrication imperfections. Interestingly FSR and the frequency offset (2δν = ν1 + ν0 − 〈ν1 + ν0〉) are weakly correlated (Fig. 4(b)). The linear regression gives FSR = 458 ± 5GHz+(0.04 ± 0.01)δν, hence an increase of the frequency also results into an increase of the frequency spacing. This is an hint that the same mechanism affecting the FSR is also involved in the fluctuations of the center frequency. This point would have been trivial in more conventional multi-mode resonators, such as rings, but it is not obvious here as each mode has a specific field distribution and is therefore sensitive to different areas of the sample. These considerations are of great importance when dealing for example with light-matter interactions such as parametric interactions [23]. It is also very desirable to have a perfect control over the tunability of the resonances when using PhC resonators for filtering. We also note that δFSR/δν = 0.04 is about an order of magnitude larger than their temperature dependence (0.004), as shown in Fig. 3(b).

 figure: Fig. 4

Fig. 4 a) Measured free spectral range (FSR) of 56 resonators. Histogram with Gaussian fit. b) FSR as a function of the frequency offset δν (linear regression with standard error).

Download Full Size | PDF

The measured FSR strengthen the assumption that holes are completely filled with SiO2, as calculations with unfilled holes predict a FSR too small to be consistent. As shown on Fig. 5(a), the frequency offset is correlated to the position of the die on the wafer. This is inherent to the MPW process which replicates the mask over the whole wafer thus scattering our devices all over the wafer. Therefore it has to be accounted that the observation of Fig. 4(b) is also linked to this spatial dispersion of the cavities. We present a map of this dispersion of the resonance frequencies over the wafer on Fig. 5(b). Quantitatively, accounting for this dependence reduces the standard deviation of ν from 0.3643 THz to 0.1890 THz. This can be explained by a decrease of the slab thickness (or an increase of the size of the holes) from the center to the edge of the wafer. These inhomogeneities are well known in the CMOS process. It can be taken into account in our simulations and, assuming that the variation is imputable mainly on the slab thickness, it leads to a variation coefficient of −0.176 GHz/nm. The corresponding thickness variation is shown on Fig. 5(a), and is in average around 5 nm, which is consistent with what is expected from this kind of SOI wafer.

 figure: Fig. 5

Fig. 5 a) Correlation of the frequency shift of the first two modes with respect to the distance R from the center of the process cylindric symmetry, and the corresponding variation of the Si layer thickness. b) Physical wafer map of the frequency shift of the resonance frequencies towards the minimum resonance frequency. One can notice that the process symmetry is not exactly centered on the wafer physical center.

Download Full Size | PDF

We now consider the distribution of the measured Q-factors (see Fig. 6(a) for devices with m=7, similar result is observed for m=6 devices). First of all, we note that we have not found any correlation with the position, contrary to the resonance frequencies. Interestingly, there is a correlation between the Q-factor of the first and second order modes (Fig. 6(b)). Importantly, the Q-factor of the fundamental mode for the resonator with weaker coupling is 4.1 × 105 in average, which is substantially larger than earlier reports [14–16]. This confirms that the standard CMOS process can be used to achieve high-Q nanocavities with a good reproducibility. As a consequence, a sufficiently large amount of cavities is tested and a statistical analysis of the devices can be performed [10,25,26], which is more insightful than a single outstanding device based study.

 figure: Fig. 6

Fig. 6 a) Distribution of the Q-factor of the first two modes over the samples for m=7. b) Correlation between the Q (mode 1) and Q (mode 0).

Download Full Size | PDF

As shown in [10,25], the statistical fluctuations of the Q-factor provide hint on the structural disorder. Hence we consider the model:

1Qexp=1Qint+1Qdis
where the measured cavity loss 1/Qexp has two contributions: the former is intrinsic to the design (radiation limited) and the material (absorption) 1/Qint, while the latter is related to the structural disorder 1/Qdis. This latter term is responsible for the fluctuations. We assume that it follows a normal distribution, thus the distribution of δQ1=1Qexp1Qexp should be centered on 0, and would have the same shape, as it is representative of the fabrication disorder. This is applied with respect to each mode and coupling (4 data groups) and the resulting δQ−1 become then independent of these characteristics. Thus, all the measurements can be aggregated to form an ensemble of 112 independent values of δQ−1. Results are shown on Fig. 7. We first note that there is no clear correlation with the die, hence with the position (Fig. 7(a)). On the other hand, the histogram of δQ−1 clearly reveals a normal distribution, which provides an a-posteriori justification of our hypothesis (Fig. 7(b)). We now connect the width of the distribution with the structural disorder, following [10]. Namely, using the formula
δQ1=δQ2=A×σhole2
Here the standard deviation is δQ−1 = 5.97 × 10−7. This term is obtained from the raw data after correcting for the measurement uncertainty itself (δQ−1 = 2.08 × 10−7, as discussed above). The same proportionality constant A = 2.57 × 10−7 as in [10], could be assumed, on the ground that the PhC structure and the field distribution are very similar. This choice would lead to an estimate of the roughness of about σhole = 1.2 nm. There are however many reasons suggesting that the roughness is smaller, consistently with values reported recently characterizing the DUV lithography in the IMEC foundry, namely ≈ 1 nm [9,15, 26], as well as other foundries: 0.84nm [10] and 1.6nm [16].

 figure: Fig. 7

Fig. 7 Fluctuation of the inverse Q-factor 1/Q − 〈1/Q〉, covering the 112 resonances and for different mode order and coupling strengths. a) distributed on the die number, and, b) histogram with Gaussian fit (red line).

Download Full Size | PDF

First, the measured δQ−1 accounts for all kind of fabrication disorder, in particular imperfect filling of the holes. The method followed here, which is based on [10], only considers positional and size disorder of hole radius, leading to an overestimate of the roughness. Moreover, preliminary calculations made on our structure suggest that in our case A is about a factor 2 larger, because radiative leakage is more sensitive to disorder in silica cladded structures. This mechanically reduces the estimate for the roughness to a value below 1 nm. This suggests that holes must be perfectly filled, as expected for this process, otherwise disorder would have been much larger. [10,16,26]. The main point, in our view, is that full-process IMEC pilot-line has been used for fabricating 2D PhC nanocavities, and seems suited for these structures, which comes somehow as a surprise. Further work will be devoted to a thorough structural analysis, in particular cross-sectional SEM images as in [10,16,26].

4. Conclusions

We have demonstrated high-Q photonic crystal cavities fabricated in a standard silicon photonics full process, within a multi project wafer run. The largest Q-factor observed (Q = 5.9 × 105, table 1) is substantially larger than any other previous report also based on CMOS DUV lithography and all-solid SiO2 cladding. This is remarkable, as it is well known that a solid-cladding considerably increases radiative leakage and therefore reduces the Q-factor. This result is about a factor 4 from the best result in suspended membranes, still fabricated with DUV [10]. A “bichromatic” design of the PhC has been used which, besides minimizing radiative losses, almost complies with the design rules of the multiproject wafer process. It is also worth in mentioning that our results are not far away from what has been achieved with all-solid cladding resonators fabricated by e-beam [11–13], in spite of the much lower resolution of DUV. The robustness of our design against fabrication imperfections seems to contribute to the result. The second is that the process used seems to guarantee the filling of the holes with silica, which is important to reduce additional disorder. These results are corroborated by a full statistical analysis over our 56 devices, showing a very good reproducibility of the performances, giving a strong argument in favor of the mass-production. Another important point is that robustness is achieved while preserving a relatively small modal volume (V = 2.2(λ/n)3) which is comparable with other geometries.

Tables Icon

Table 1. Recent results regarding high-Q cavities. with and without (parenthesis) SiO2 cladding. a average and maximum value.

We have shown that compact and high-Q resonators can be included in a standard integrated photonic circuit using the most general and relatively less expensive process. These cavities could for instance be used for photonic filtering in complex photonic processing chips, in the context of 5G or microwave applications. Tunability could easily be implemented using integrated heaters. Also, these resonators could be used in the context of the quantum information in integrated circuits.

Funding

European Union Horizon 2020 research and innovation programme under grant agreement No 780848 (FunComp) and the project “Symphonie” funded by the French Research Agency (ANR).

References

1. G. Crosnier, D. Sanchez, S. Bouchoule, P. Monnier, G. Beaudoin, I. Sagnes, R. Raj, and F. Raineri, “Hybrid indium phosphide-on-silicon nanolaser diode,” Nat. Photonics 11, 297–3002017.

2. K. Nozaki, A. Shinya, S. Matsuo, Y. Suzaki, T. Segawa, T. Sato, Y. Kawaguchi, R. Takahashi, and M. Notomi, “Ultralow-power all-optical RAM based on nanocavities,” Nat. Photonics 6, 248–2522012.

3. S. Noda, A. Chutinan, and M. Imada, “Trapping and emission of photons by a single defect in a photonic bandgap structure,” nature 407, 608 (2000). [CrossRef]   [PubMed]  

4. K. Hennessy, A. Badolato, M. Winger, D. Gerace, M. Atatüre, S. Gulde, S. Fält, E. L. Hu, and A. Imamoglu, “Quantum nature of a strongly coupled single quantum dot cavity system,” Nature 445, 896–8992007.

5. J. D. Thompson, T. G. Tiecke, N. P. de Leon, J. Feist, A. V. Akimov, M. Gullans, A. S. Zibrov, V. Vuletic, and M. D. Lukin, “Coupling a single trapped atom to a nanoscale optical cavity,” Science 340, 1202–12052013. [PubMed]  

6. M. Eichenfield, J. Chan, R. M. Camacho, K. J. Vahala, and O. Painter, “Optomechanical crystals,” Nature 462, 78–822009. [PubMed]  

7. H. Sekoguchi, Y. Takahashi, T. Asano, and S. Noda, “Photonic crystal nanocavity with a Q-factor of ∼9 million,” Opt. Express 22, 916 (2014). [CrossRef]   [PubMed]  

8. T. Asano, Y. Ochi, Y. Takahashi, K. Kishimoto, and S. Noda, “Photonic crystal nanocavity with a Q factor exceeding eleven million,” Opt. Express 25, 1769 (2017). [CrossRef]  

9. K. Welna, K. Debnath, T. F. Krauss, and L. O’Faolain, “High-Q photonic crystal cavities realised using deep ultraviolet lithography,” Electron. Lett. 51, 1277–12792014.

10. K. Ashida, M. Okano, M. Ohtsuka, M. Seki, N. Yokoyama, K. Koshino, M. Mori, T. Asano, S. Noda, and Y. Takahashi, “Ultrahigh-q photonic crystal nanocavities fabricated by cmos process technologies,” Opt. Express 25, 18165–18174 (2017). [CrossRef]   [PubMed]  

11. B.-S. Song, S.-W. Jeon, and S. Noda, “Symmetrically glass-clad photonic crystal nanocavities with ultrahigh quality factors,” Opt. letters 36, 91–93 (2011). [CrossRef]  

12. S.-W. Jeon, J.-k. Han, B.-S. Song, and S. Noda, “Glass-embedded two-dimensional silicon photonic crystal devices with a broad bandwidth waveguide and a high quality nanocavity,” Opt. express 18, 19361–19366 (2010). [CrossRef]   [PubMed]  

13. X. Gai, B. Luther-Davies, and T. P. White, “Photonic crystal nanocavities fabricated from chalcogenide glass fully embedded in an index-matched cladding with a high Q-factor (> 750,000),” Opt. express 20, 15503–15515 (2012). [CrossRef]   [PubMed]  

14. E. Kuramochi, H. Taniyama, T. Tanabe, K. Kawasaki, Y.-G. Roh, and M. Notomi, “Ultrahigh-Q one-dimensional photonic crystal nanocavities with modulated mode-gap barriers on SiO2 claddings and on air claddings,” Opt. express 18, 15859–15869 (2010). [CrossRef]   [PubMed]  

15. W. Xie, M. Fiers, S. Selvaraja, P. Bienstman, J. Van Campenhout, P. Absil, and D. Van Thourhout, “High-q photonic crystal nanocavities on 300 mm soi substrate fabricated with 193 nm immersion lithography,” J. Light. Technol. 32, 1457–1462 (2014). [CrossRef]  

16. Y. Ooka, T. Tetsumoto, A. Fushimi, W. Yoshiki, and T. Tanabe, “CMOS compatible high-Q photonic crystal nanocavity fabricated with photolithography on silicon photonic platform,” Sci. Reports 5, 11312 (2015). [CrossRef]  

17. Y. Akahane, T. Asano, B.-S. Song, and S. Noda, “High-Q photonic nanocavity in a two-dimensional photonic crystal,” Nature 425, 944–947 (2003). [CrossRef]   [PubMed]  

18. Y. Tanaka, T. Asano, and S. Noda, “Design of Photonic Crystal Nanocavity with Q-Factor of 109,” J. Light. Technol. 26, 1532–1539 (2008). [CrossRef]  

19. M. W. McCutcheon and M. Loncar, “Design of a silicon nitride photonic crystal nanocavity with a quality factor of one million for coupling to a diamond nanocrystal,” Opt. Express 16, 191362008. [PubMed]  

20. A. Bazin, R. Raj, and F. Raineri, “Design of silica encapsulated High-Q photonic crystal nanobeam cavity,” J. Light. Technol. 32, 952–9582014.

21. F. Alpeggiani, L. C. Andreani, and D. Gerace, “Effective bichromatic potential for ultra-high Q-factor photonic crystal slab cavities,” Appl. Phys. Lett. 107, 261110 (2015). [CrossRef]  

22. A. Simbula, M. Schatzl, L. Zagaglia, F. Alpeggiani, L. Andreani, F. Schäffler, T. Fromherz, M. Galli, and D. Gerace, “Realization of high-Q/v photonic crystal cavities defined by an effective aubry-andré-harper bichromatic potential,” APL Photonics 2, 056102 (2017). [CrossRef]  

23. S. Combrié, G. Lehoucq, G. Moille, A. Martin, and A. Rossi, “Comb of high-q resonances in a compact photonic cavity,” Laser & Photonics Rev. 11170099 (2017).

24. Y. A. Vlasov and S. J. McNab, “Coupling into the slow light mode in slab-type photonic crystal waveguides,” Opt. letters 31, 50–522006.

25. Y. Taguchi, Y. Takahashi, Y. Sato, T. Asano, and S. Noda, “Statistical studies of photonic heterostructure nanocavities with an average Q factor of three million,” Opt. Express 19, 11916 (2011). [CrossRef]   [PubMed]  

26. R. Benevides, F. G. Santos, G. O. Luiz, G. S. Wiederhecker, and T. P. M. Alegre, “Ultrahigh-q optomechanical crystal cavities fabricated in a cmos foundry,” Sci. reports 7, 2491 (2017). [CrossRef]  

27. G. Crosnier, D. Sanchez, A. Bazin, P. Monnier, S. Bouchoule, R. Braive, G. Beaudoin, I. Sagnes, R. Raj, and F. Raineri, “High Q factor InP photonic crystal nanobeam cavities on silicon wire waveguides,” Opt. letters 41, 579–582 (2016). [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 (7)

Fig. 1
Fig. 1 a) Optical image of the chip corresponding to our user area within the multi-project wafer. The 8 PhC Cavities are located within the red rectangle. b) Optical image of one resonator, with the complete PhC area delimited by the blue line. Deep-etched ridge waveguides are used to connect to grating couplers. c) 3D representation of the PhC.
Fig. 2
Fig. 2 a) Design of a PhC using a bichromatic lattice. In black is the main triangular PhC lattice of period a. In red is the cavity, formed by a missing line defect surrounded by 2 rows with a slightly different period a′ = 0.98a. In blue are the 2 coupling waveguides (same parameter a as the main lattice). b) Representation of the fundamental mode in the reciprocal space, with the circle corresponding to |k//| < ωnclad/c. c) Modal distribution of the electric field (Ey) for the fundamental mode.
Fig. 3
Fig. 3 a) Measured transmission (black dots) using a coherent detection technique (OCT) revealing the spectral lineshape of the fundamental order mode of the PhC cavity along with a Lorentzian fit. b) temperature dependence of the frequency of the first two modes, and their spectral separation Δν = ν1ν0.
Fig. 4
Fig. 4 a) Measured free spectral range (FSR) of 56 resonators. Histogram with Gaussian fit. b) FSR as a function of the frequency offset δν (linear regression with standard error).
Fig. 5
Fig. 5 a) Correlation of the frequency shift of the first two modes with respect to the distance R from the center of the process cylindric symmetry, and the corresponding variation of the Si layer thickness. b) Physical wafer map of the frequency shift of the resonance frequencies towards the minimum resonance frequency. One can notice that the process symmetry is not exactly centered on the wafer physical center.
Fig. 6
Fig. 6 a) Distribution of the Q-factor of the first two modes over the samples for m=7. b) Correlation between the Q (mode 1) and Q (mode 0).
Fig. 7
Fig. 7 Fluctuation of the inverse Q-factor 1/Q − 〈1/Q〉, covering the 112 resonances and for different mode order and coupling strengths. a) distributed on the die number, and, b) histogram with Gaussian fit (red line).

Tables (1)

Tables Icon

Table 1 Recent results regarding high-Q cavities. with and without (parenthesis) SiO2 cladding. a average and maximum value.

Equations (2)

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

1 Q exp = 1 Q int + 1 Q dis
δ Q 1 = δ Q 2 = A × σ hole 2
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.