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

Grating-assisted silicon-on-insulator racetrack resonator reflector

Open Access Open Access

Abstract

We experimentally demonstrate a grating-assisted silicon-on-insulator (SOI) racetrack resonator reflector with a reflect port suppression of 10.3 dB and no free spectral range. We use contra-directional grating couplers within the coupling regions of the racetrack resonator to enable suppression of all but one of the peaks within the reflect port spectrum as well as all but one of the notches within the through port spectrum.

© 2015 Optical Society of America

1. Introduction

Silicon photonics integrated devices provide the opportunity to substantially miniaturize optics-based filters as compared to fiber-optics-based filters, and numerous companies are involved in silicon photonics research, see for example [1–9]. Resonator reflectors have been proposed for sensing applications [10–15], optical communication applications [15–17], optical feedback applications [16, 18–21], and have been experimentally demonstrated in a variety of configurations [10–16, 18–21, 23–25]. The features desired in a ring resonator reflector are application-dependent. For sensing applications, it is important that the free spectral range (FSR) of a ring resonator reflector be large enough so that a change in the ring’s resonance shift does not exceed its FSR [26]. A narrow bandwidth (large Q-factor) is also a desirable feature for sensing, as it allows smaller resonance shifts to be detected [26, 27]. In optical communications applications, a larger FSR allows more channels to be utilized for multiplexing and demultiplexing signals [28]. Also, in optical communications applications the required bandwidth is dependent upon a number of considerations, such as the operating data-rate. Therefore, flexibility in the design of the resonator’s bandwidth is necessary. In optical feedback applications involving narrow linewidth, single wavelength lasers, use of ring resonators can provide narrow bandwidths and large FSRs [29–32]. Therefore, in these applications, it is advantageous to eliminate the FSR. However, many of the ring resonator reflectors reported to date have FSRs, see for example [10–16, 18–20, 23, 25].

Nevertheless, reflector designs that eliminate the FSR have been reported [17,21,22,33–35]. For example, an all-pass silicon nitride microring resonator with a distributed Bragg reflector (DBR) has been experimentally demonstrated which showed suppression of minor notches in the through port and of minor peaks in the reflect port (i.e., no FSR) [21] and designs employing Fabry-Pérot structures composed of silicon directional couplers with silicon DBRs have also been experimentally demonstrated which showed elimination of the FSR [34,35]. Also, silicon Bragg grating reflectors, without ring resonators, have been experimentally demonstrated [12, 13, 36, 37].

In this paper, we experimentally demonstrate a grating-assisted silicon-on-insulator (SOI) racetrack resonator reflector by placing the racetrack resonator within a Mach-Zehnder interferometer (MZI) and adding contra-directional grating couplers (contra-DCs) to effectively eliminate the FSR. Our device is similar to those presented in [11, 16, 18], except that their devices do not use contra-DCs and their devices have FSRs.

2. Design

In order to eliminate the FSR of our device, we use contra-DCs [22, 38, 39]. Contra-DCs have been demonstrated to have highly wavelength-dependent coupling as compared to the coupling of co-directional couplers without gratings [40–47], and silicon racetrack resonators with contra-DCs have been experimentally demonstrated [38, 39, 45, 48] and have shown that the removal of the FSR is possible [38, 39, 45].

Figure 1(a) shows a diagram of our grating-assisted SOI racetrack resonator reflector where Ls is the length of the straight section of each of the branches of the MZI, Lb is the length of each S-bend, r is the radius of the bend regions of the racetrack resonator, and Lcd is the length of the contra-DC. Figure 1(b) shows a section of our contra-DC, where Λ is the grating period, G is the gap distance, wa and wb are the average waveguide widths of waveguide “a” and waveguide “b” of the coupler, respectively, and Δwa and Δwb are the corrugation widths for the gratings on waveguide “a” and waveguide “b” in the coupler, respectively [39, 47]. Also, we have chosen to use anti-reflection gratings to suppress the intra-waveguide Bragg reflections [39,41,44,47,49]. The bends of the racetrack resonator had widths equal to wb. The input Y-branch of our device splits the light equally into the two MZI branches and the light couples into the racetrack resonator at wavelengths near the Bragg condition of the contra-DC while light at other wavelengths is not coupled into the ring but passes through the MZI and recombines at the output Y-branch. The major resonant peak will occur at that wavelength within the contra-DC passband that aligns with a resonance of the racetrack resonator [38, 39]. As a result, provided that only one resonance occurs within the passband of the contra-DC, all other resonant notches and peaks of the racetrack resonator can be suppressed. In this way, provided that the FSR of the racetrack resonator is sufficiently large, the FSR of our device can be effectively eliminated. The light that was coupled into the racetrack resonator from the top branch is coupled out of the racetrack resonator at the bottom branch and vice-versa. Hence, the light coupled out of the racetrack resonator in both branches is constructively recombined at the input Y-branch and can be detected at the reflect port [Fig. 1(a)].

 figure: Fig. 1

Fig. 1 (a) Diagram of our reflector. (b) Diagram showing a section of our contra-DC with anti-reflection gratings [39].

Download Full Size | PDF

For the fabricated device that we present here, our mask layout for the device had the following design parameters: wa was 450 nm and wb was 550 nm [22,39,44,47], Δwa was 30 nm and Δwb was 40 nm [39,44,47], the waveguide heights were 220 nm [22,39,44,47], the bends of the racetrack resonator had widths of 550 nm [22, 39], G was 300 nm, r was 3 μm, Lcd was 56.784 μm (the number of grating periods was 182 and Λ was 312 nm), Ls was 156.784 μm, and Lb was 10.705 μm. Each MZI S-bend had a waveguide width of 500 nm which was then tapered to 450 nm [the tapered waveguide regions are shown in blue in Fig. 1(a)]. The reflector design was fabricated using electron beam lithography [50] and the device was covered with an oxide cladding. Shallow-etched fiber grating couplers were used for coupling light into and out of the device [51]. For testing purposes, we added an extra Y-branch before the input port shown in Fig. 1(a) so that we could separate the reflected light from the light injected into the device and, thus, be able to measure the reflect port spectrum. The designs of the Y-branch junctions were based on [52, 53].

3. Principle of operation

In this section, we conceptually describe the principle of operation of our reflector design. Figures 2(a) and 2(b) show the theoretical contra-DC power coupling and power transmission factors versus wavelength, respectively, for the straight sections of the racetrack resonator. Figure 2(c) shows the theoretical spectrum for the reflect port when using contra-DCs, as well as when using co-directional couplers without gratings, to couple the light into and out of the racetrack resonator. We can clearly see that our device suppresses all but one of the peaks and, therefore, effectively eliminates the FSR. Figure 2(d) shows the theoretical through port spectrum when using contra-DCs, as well as when using co-directional couplers without gratings, which shows that the FSR is also effectively eliminated at the through port when contra-DCs are used. For these figures, we have assumed the following values: the coupling coefficient is 2000 m−1; the Y-branch junction loss is 0.28 dB [52]; and the propagation loss for the racetrack resonator and MZI branches is 3 dB/cm. It should be noted that these theoretical results will not exactly match the experimental results since we have assumed values for the coupling coefficient and losses and we have ignored all scattering, e.g., backscattering and scattering from dielectric discontinuities. For the contra-DCs, we have assumed no co-directional coupling due to the large difference in the waveguide widths of the contra-DCs [38, 39, 44, 54]. For example, [44] experimentally demonstrates a contra-DC (average waveguide widths of 450 nm and 550 nm) that has small co-directional coupling. Also, the phases of the MZI branches, including propagation through the contra-DCs, were calculated assuming waveguide widths of 450 nm. For the modeling of the reflector without gratings (i.e., using normal co-directional couplers), the co-directional power coupling factor was set to the maximum contra-DC power coupling factor that was used for the device with gratings [see red dot in Fig. 2(a)]. Also, the co-directional power transmission factor was set to the minimum contra-DC power transmission factor that was used for the device with gratings [see red dot in Fig. 2(b)]. The phases of the MZI branches, including the co-directional coupling regions, were calculated assuming waveguide widths of 550 nm. The modeling of the racetrack resonator with and without contra-DCs is similar to the model used in [39]. The effective indices of the waveguides were calculated using MODE Solutions by Lumerical Solutions, Inc., [39, 47, 55]. The reflect port transfer function (Eq. 8) and the through port transfer function (Eq. (15)) are derived and given in the appendix for the device with gratings.

 figure: Fig. 2

Fig. 2 (a) Theoretical contra-DC power coupling factor versus wavelength (the red dot indicates the value used for the co-directional power coupling factor of the reflector without gratings). (b) Theoretical contra-DC power transmission factor versus wavelength (the red dot indicates the value used for the co-directional power transmission factor of the reflector without gratings). Theoretical comparison of (c) the reflect port spectrum and (d) the through port spectrum of our reflector with contra-DCs and with co-directional couplers without gratings.

Download Full Size | PDF

4. Measurement results

Here we present experimental results for our device. Figures 3(a) and 3(b) show the experimental reflect port spectrum (normalized to its maximum power) measured at a temperature of 25°C, which show the elimination of the FSR. For the reflect port spectrum, the reflect port peak suppression is 10.3 dB and the 3-dB bandwidth is 0.147 nm. The red dot in Fig. 3(b) represents the wavelength at which data was sent through the device to its reflect port (as described below). Figure 4(a) shows the experimental through port spectrum [normalized to the maximum power within the wavelength region shown in Fig. 4(a)] measured at a temperature of 25°C, which also shows the elimination of the FSR. The extinction ratio of the major notch is about 7 dB and the extinction ratio of the largest minor notch is about 1 dB. Improved performance, in terms of increased reflect port suppression, is likely achievable by thermally tuning the contra-DCs and/or the bend regions of the racetrack resonator to compensate for fabrication variations (see experimental results in [39]). Although there is significant suppression of the minor notches, it is still important to determine the extent of the through port dispersion within the regions of the suppressed minor notches (for example, see [56]). Figure 4(b) shows the measured through port dispersion at wavelengths near the first minor notch to the right of the major notch. The green dot in Fig. 4(b) represents the wavelength at which data was sent through the device to its through port (as described below). The through port dispersion is the average of 300 measurements using an Optical Vector Analyzer™ STe by Luna Innovations, Inc., [47].

 figure: Fig. 3

Fig. 3 (a) Measured reflect port spectrum and (b) the spectrum at wavelengths near the major peak (the red dot indicates the wavelength at which the data in Fig. 5(a) was sent through the device).

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 (a) Measured through port spectrum at wavelengths containing the major notch and one of the minor notches. (b) Measured through port dispersion at wavelengths near the minor notch (the green dot indicates the wavelength at which the data in Fig. 5(b) was sent through the device).

Download Full Size | PDF

Previously, we demonstrated a method to determine the coupling coefficients of fabricated contra-DCs and verified the approach using contra-DCs fabricated on three separate fabrication runs [47]. One of our fabricated contra-DCs in [47] had the same as-designed average waveguide widths, waveguide corrugation widths, grating period, and gap distance as the contra-DC we use in this paper. Also, the same fabrication process was used for the contra-DCs in this paper as the one used in [47]. The average extracted coupling coefficient for contra-DCs with gap distances of 300 nm based on the results of the three fabrication runs in [47] is 4139 m−1. Using this extracted coupling coefficient, we have simulated the power coupling factor versus wavelength of the contra-DC as well as the reflect port spectrum of the reflector. The simulated 3-dB bandwidth of the contra-DC is 4.450 nm. The value for the maximum reflectivity, R, in dB [RdB = 10log10(R)] of the reflector based on the simulated results is −1.28 dB. However, the maximum, inherent RdB (removing all loss mechanisms) is 0.00 dB, therefore, as SOI processing methods improve, e.g., reductions in waveguide propagation losses [57], the reflectivity of our device should only improve and will be limited only by the losses in the Y-branch junctions and the radiation losses in the ring.

Next, data was sent through the device using an external MZI modulator and eye diagrams were obtained (we used a set-up similar to the ones in [44,56,58]) at the reflect port and through port of our device. Figure 5(a) shows an open eye diagram (extinction ratio, ER, is 13.0 dB) at the reflect port for a 12.5 Gbps NRZ PRBS-31 signal operating at a wavelength of 1535.08 nm. Next, we sent a 12.5 Gbps NRZ PRBS-31 signal operating at a wavelength of 1539.42 nm, corresponding to the center of one of the through port minor notches, through our device and measured the eye diagram at the through port. Figure 5(b) shows that an open eye diagram (ER = 14.0 dB) is achievable at the through port when operating at the center of the minor notch. Since the dispersion within the through port minor notch is highly wavelength-dependent [as shown in Fig. 4(b)], it is expected that a change in the operating wavelength will affect the extent of achievable eye opening (a similar trend has been experimentally demonstrated within one of the minor notches of a quadruple Vernier racetrack resonator [56]). As compared to the reflectors presented in [21, 34, 35] which also have no FSRs, here, we have verified that our reflector can be used to transmit data.

 figure: Fig. 5

Fig. 5 (a) Measured reflect port eye diagram for a 12.5 Gbps signal operating at a wavelength of 1535.08 nm. (b) Measured through port eye diagram for a 12.5 Gbps signal operating at a wavelength of 1539.42 nm.

Download Full Size | PDF

Also, of interest is investigating the residual signal at the major notch of the through port, i.e., when the signal is being reflected. Hence, we measured eye diagrams for a 12.5 Gbps NRZ PRBS-31 signal operating at the wavelength corresponding to the center of the major notch for both the through and reflect ports. Figures 6(a) and 6(b) show the two eye diagrams side-by-side. As expected, the eye diagram at the through port major notch is more closed as compared to the eye diagram at the reflect port major peak.

 figure: Fig. 6

Fig. 6 (a) Measured through port eye diagram for a 12.5 Gbps signal and (b) measured reflect port eye diagram for a 12.5 Gbps signal both operating at the wavelength corresponding to the center of the major notch and are on the same scale.

Download Full Size | PDF

We then measured the spectrum of our device at different temperatures by heating the entire device. Figure 7(a) shows the experimental results when the temperature is changed from 15°C to 55°C (each measurement was normalized to its maximum power). Since our device suppresses all but one of the major peaks, we could adjust the temperature such that the major peak shifts over a wavelength range larger than the wavelength range that would correspond to the FSR of our device without gratings, i.e., the racetrack resonator’s FSR which we call FSRrr [Fig. 7(b)]. The resonant peak wavelength temperature dependence of our device is 0.0703 nm/°C [Fig. 7(b)], which provides the potential for our device to be used in temperature sensing applications, see [10, 14, 59–63].

 figure: Fig. 7

Fig. 7 (a) The reflect port spectrum of our device that has a gap distance of 300 nm measured at different temperatures where from left to right the peaks correspond to 15°C, 25°C, 35°C, 45°C, and 55°C. (b) Resonant peak wavelength versus temperature (dots indicate the measured results and the red line is the fit).

Download Full Size | PDF

Table 1 shows a comparison between previously reported on FSR-free reflectors [21, 34–36] and our reflector. Our reflector is substantially smaller than the reflectors presented in [21] and [34]. The reflectors in [35] and [36] are more compact than our device, however, our 3-dB bandwidth is much smaller than theirs. Although our device does not have the smallest footprint, it can be re-designed to optimize the footprint by using contra-DC bus waveguides that have widths of 500 nm, which would enable us to remove the tapered waveguide sections and, therefore, our device could have a footprint of 80.8 μm × 8.1 μm. The reflectivities presented in [21] and [36] are higher than our device’s reflectivity, however, the reflect port suppression (suppression is defined as the difference between the maximum intensity of the major peak and the maximum intensity of the largest minor peak) of our device is larger and the bandwidth is smaller as compared to those in [21] and [36]. Also, our reflect port suppression is comparable to the suppressions seen in [34] and in [35]. Additionally, [21, 34–36] provide no theoretical or experimental results for the dispersion of their devices and, thus, it is not possible to determine the extent to which the signal may be degraded by these reflectors. We, however, have experimentally demonstrated the through port dispersion at one of the minor through port notches as well as verified that our reflector can operate on PRBS data.

Tables Icon

Table 1. Comparison of FSR-free reflectors.

5. Conclusion

We have presented conceptual, theoretical (see appendix), and experimental results that show that it is possible to effectively eliminate the FSR of a SOI racetrack resonator reflector by using contra-DCs. The experimental results show a reflect port suppression of 10.3 dB. Also, open eye diagrams were measured at the reflect port and the through port of our device. Our reflector can also be used as a temperature sensor with a temperature-dependent spectral shift of 0.0703 nm/°C.

Appendix

To determine the transfer functions of our reflector, we used Mason’s rule [64–68]. Our reflector has two loop gains, GL1 and GL2, which are shown graphically in Fig. 8 and given by,

GL1=GL2=tcd2Xr,
where tcd is the contra-DC field transmission factor [54] (we have included the loss and the phase due to the length of the coupler [39]) for the straight sections of the racetrack resonator. Xr = exp(− rLrαrLr) where βr is the propagation constant of the bend regions of the racetrack resonator, αr is the field loss coefficient of the bend regions of the racetrack resonator, and Lr is the length of the racetrack resonator excluding the lengths of the contra-DCs. tcd is given by [39, 54],
tcd=sejΔβ2Lcdscosh(sLcd)+jΔβ2sinh(sLcd)ejβbLcdαbLcd,
where Δβ = βa + βb − 2π/Λ [54], βa is the propagation constant of the portion of the contra-DC within the racetrack resonator’s bus waveguide, βb is the propagation constant of the portion of the contra-DC within the racetrack resonator, Λ is the grating period, s=(|κ|2Δβ2/4)12 [54], κ is the coupling coefficient, αb is the field loss coefficient of the portion of the contra-DC within the racetrack resonator, and Lcd is the length of the contra-DC. In our model, we have assumed κ is real-valued since a complex-valued κ would only result in additional phase terms in the numerator of the reflect port transfer function (Eq. 8).

 figure: Fig. 8

Fig. 8 Diagram showing the loop gains, GL1 and GL2.

Download Full Size | PDF

The determinant, Δ, of our reflector is given by,

Δ=1(GL1+GL2)+GL1GL2.
The reflect port has two forward path gains, GP1 and GP2, (shown graphically in Fig. 9) and two co-factors, ΔP1 and ΔP2,
GP1=GP2=κcd2Xr12Xmzi2K22,
ΔP1=1GL2,
ΔP2=1GL1,
where κcd is contra-DC field coupling factor [54], Xmzi = exp(− mziLmziαmziLmzi), βmzi is the propagation constant of the MZI branch, αmzi is the field loss coefficient of the MZI branch, Lmzi is the length of one of the MZI S-bends plus the length of one of the tapered waveguide regions of the MZI branch, and K accounts for the field reduction going through each of the Y-branch junctions. κcd is defined as [54],
κcd=jκsinh(sLcd)scosh(sLcd)+jΔβ2sinh(sLcd).

 figure: Fig. 9

Fig. 9 Diagrams showing the two forward path gains, GP1 and GP2, from input port to reflect port.

Download Full Size | PDF

Therefore, the reflect port transfer function is,

TFreflect=GP1ΔP1+GP2ΔP2Δ.
The through port has four forward path gains, GP3, GP4, GP5, and GP6, (shown graphically in Fig. 10) and four co-factors, ΔP3, ΔP4, ΔP5, and ΔP6,
GP3=GP4=tcdbusXmzi2K22,
ΔP3=ΔP4=Δ,
GP5=GP6=κcd2tcdXrXmzi2K22,
ΔP5=1GL2,
ΔP6=1GL1.
In Eq. (9), tcd–bus is defined as the contra-DC bus waveguide field transmission factor [54] (we have included the loss and the phase due to the length of the coupler [39]),
tcdbus=sejΔβ2Lcdscosh(sLcd)+jΔβ2sinh(sLcd)ejβaLcdαaLcd,
where αa is the field loss coefficient of the portion of the contra-DC within the racetrack resonator’s bus waveguide. In our model we have set αr = αmzi = αa = αb, βr = βb, and βmzi = βa.

 figure: Fig. 10

Fig. 10 Diagrams showing the four forward path gains, GP3, GP4, GP5, and GP6, from input port to through port.

Download Full Size | PDF

Therefore, the through port transfer function is,

TFthrough=GP3ΔP3+GP4ΔP4+GP5ΔP5+GP6ΔP6Δ.

Acknowledgments

We acknowledge the Natural Sciences and Engineering Research Council of Canada, CMC Microsystems, the SiEPIC program, Lumerical Solutions, Inc., and Mentor Graphics. We thank Anritsu for the use of their MP1800A Signal Quality Analyzer and Richard Bojko for fabrication of the device. We thank Dr. Wei Shi, Han Yun, Alex MacKay, Yun Wang, and Jonas Flueckiger for their help. Part of this work was conducted at the University of Washington Nanofabrication Facility, a member of the NSF National Nanotechnology Infrastructure Network.

References and links

1. H.-F. Liu, “Integrated silicon photonics links for high bandwidth data transportation,” in Optical Fiber Communication Conference, OSA Technical Digest (online) (Optical Society of America, 2014), paper Th1D.1.

2. A. Ramaswamy, J. E. Roth, E. J. Norberg, R. S. Guzzon, J. H. Shin, J. T. Imamura, B. R. Koch, D. K. Sparacin, G. A. Fish, B. G. Lee, R. Rimolo-Donadio, C. W. Baks, A. Rylyakov, J. Proesel, M. Meghelli, and C. L. Schow, “A WDM 4×28Gbps integrated silicon photonic transmitter driven by 32nm CMOS driver ICs,” in Optical Fiber Communication Conference Post Deadline Papers, OSA Technical Digest (online) (Optical Society of America, 2015), paper Th5B.5.

3. X. Zheng and A. V. Krishnamoorthy, “A WDM CMOS photonic platform for chip-to-chip optical interconnects,” in CLEO: 2014, OSA Technical Digest (online) (Optical Society of America, 2014), paper SM4O.3.

4. C.-H. Chen, T.-C. Huang, D. Livshit, A. Gubenko, S. Mikhrin, V. Mikhrin, M. Fiorentino, and R. Beausoleil, “A comb laser-driven DWDM silicon photonic transmitter with microring modulator for optical interconnect,” in CLEO: 2015, OSA Technical Digest (online) (Optical Society of America, 2015), paper STu4F.1.

5. M. Mazzini, M. Traverso, M. Webster, C. Muzio, S. Anderson, P. Sun, D. Siadat, D. Conti, A. Cervasio, S. Pfnuer, J. Stayt, M. Nyland, C. Togami, K. Yanushefski, and T. Daugherty, “25GBaud PAM-4 error free transmission over both single mode fiber and multimode fiber in a QSFP form factor based on silicon photonics,” in Optical Fiber Communication Conference Post Deadline Papers, OSA Technical Digest (online) (Optical Society of America, 2015), paper Th5B.3.

6. Luxtera Inc., http://www.luxtera.com

7. D. Mahgerefteh and C. Thompson, “Techno-economic comparison of silicon photonics and multimode VCSELs,” in Optical Fiber Communication Conference, OSA Technical Digest (online) (Optical Society of America, 2015), paper M3B.2.

8. D. M. Calhoun, Q. Li, C. Browning, N. C. Abrams, Y. Liu, R. Ding, L. P. Barry, T. Baehr-Jones, M. Hochberg, and K. Bergman, “Programmable wavelength locking and routing in a silicon-photonic interconnection network implementation,” in Optical Fiber Communication Conference, OSA Technical Digest (online) (Optical Society of America, 2015), paper Tu2H.3.

9. Y. Painchaud, M. Poulin, F. Pelletier, C. Latrasse, J.-F. Gagné, S. Savard, G. Robidoux, M.-J. Picard, S. Paquet, C.-A. Davidson, M. Pelletier, M. Cyr, C. Paquet, M. Guy, M. Morsy-Osman, M. Chagnon, and D. V. Plant, “Silicon-based products and solutions,” Proc. SPIE 8988, 89880L (2014). [CrossRef]  

10. H. Sun, A. Chen, and L. R. Dalton, “A reflective microring notch filter and sensor,” Opt. Express 17(13), 10731–10737 (2009). [CrossRef]   [PubMed]  

11. W. Shi, H. Yun, W. Zhang, C. Lin, T. K. Chang, Y. Wang, N. A. F. Jaeger, and L. Chrostowski, “Ultra-compact, high-Q silicon microdisk reflectors,” Opt. Express 20(20), 21840–21846 (2012). [CrossRef]   [PubMed]  

12. P. Muellner, R. Bruck, M. Karl, M. Baus, T. Wahlbrink, and R. Hainberger, “Silicon photonic wire Bragg grating for on-chip wavelength (de)multiplexing employing ring resonators,” in Advanced Photonics, OSA Technical Digest (CD) (Optical Society of America, 2011), paper IMF4. [CrossRef]  

13. P. Muellner, R. Bruck, M. Baus, M. Karl, T. Wahlbrink, and R. Hainberger, “Silicon photonic MZI sensor array employing on-chip wavelength multiplexing,” Opt. Quantum Electron. 44(12–13), 557–562 (2012). [CrossRef]  

14. W. Shi, R. Vafaei, M. Á. G. Torres, N. A. F. Jaeger, and L. Chrostowski, “Design and characterization of microring reflectors with a waveguide crossing,” Opt. Lett. 35(17), 2901–2903 (2010). [CrossRef]   [PubMed]  

15. H. Yun, W. Shi, X. Wang, L. Chrostowski, and N. A. F. Jaeger, “Dumbbell micro-ring reflector,” Proc. SPIE 8412, 84120P (2012). [CrossRef]  

16. G. T. Paloczi, J. Scheuer, and A. Yariv, “Compact microring-based wavelength-selective inline optical reflector,” IEEE Photon. Technol. Lett. 17(2), 390–392 (2005). [CrossRef]  

17. S. Vargas and C. Vazquez, “Optical reconfigurable demultiplexer based on Bragg grating assisted ring resonators,” Opt. Express 22(16), 19156–19168 (2014). [CrossRef]   [PubMed]  

18. X. Li, Y. Wan, T. Hu, P. Yu, H. Yu, J. Yang, and X. Jiang, “A tunable silicon ring reflector,” J. Opt. 44(1), 26–29 (2015). [CrossRef]  

19. L. S. Stewart and P. D. Dapkus, “In-plane thermally tuned silicon-on-insulator wavelength selective reflector,” in Integrated Photonics Research, Silicon and Nanophotonics and Photonics in Switching, OSA Technical Digest (CD) (Optical Society of America, 2010), paper IME2.

20. J. Scheuer, G. T. Paloczi, and A. Yariv, “All optically tunable wavelength-selective reflector consisting of coupled polymeric microring resonators,” Appl. Phys. Lett. 87(25), 251102 (2005). [CrossRef]  

21. A. Arbabi, Y. M. Kang, C.-Y. Lu, E. Chow, and L. L. Goddard, “Realization of a narrowband single wavelength microring mirror,” Appl. Phys. Lett. 99(9), 091105 (2011). [CrossRef]  

22. R. Boeck, M. Caverley, L. Chrostowski, and N. A. F. Jaeger, “An FSR-free silicon resonator reflector using a contra-directional coupler and a Bragg reflector,” in Proc. Photonics North 2015 (IEEE, 2015), in press.

23. J. Park, T. Lee, D. Lee, S. Kim, W. Hwang, and Y. Chung, “Widely tunable coupled-ring-reflector filter based on planar polymer waveguide,” IEEE Photon. Technol. Lett. 20(12), 988–990 (2008). [CrossRef]  

24. A. Samarelli, A. Canciamilla, G. Morea, F. Morichetti, R. M. De LaRue, A. Melloni, and M. Sorel, “Grating-assisted micro-ring resonators for silicon dual mode filters,” in CLEO/Europe and EQEC 2011 Conf. Digest, OSA Technical Digest (CD) (Optical Society of America, 2011), paper CK_P10.

25. C. Alonso-Ramos, F. Morichetti, A. Ortega-Moñux, I. Molina-Fernández, M. J. Strain, and A. Melloni, “Dual-mode coupled-resonator integrated optical filters,” IEEE Photon. Technol. Lett. 26(9), 929–932 (2014). [CrossRef]  

26. C.-Y. Chao and L. J. Guo, “Design and optimization of microring resonators in biochemical sensing applications,” J. Lightw. Technol. 24(3), 1395–1402 (2006). [CrossRef]  

27. K. De Vos, “Label-free silicon photonics biosensor platform with microring resonators,” PhD thesis, Universiteit Gent (2010).

28. O. Schwelb, “The nature of spurious mode suppression in extended FSR microring multiplexers,” Opt. Commun. 271(2), 424–429 (2007). [CrossRef]  

29. T. Segawa, W. Kobayashi, T. Nakahara, and R. Takahashi, “Wavelength-routed switching for 25-Gbit/s optical packets using a compact transmitter integrating a parallel-ring-resonator tunable laser and an InGaAlAs EAM,” IEICE Trans. Electron. E97(7), 719–724 (2014). [CrossRef]  

30. B. Liu, A. Shakouri, and J. E. Bowers, “Wide tunable double ring resonator coupled lasers,” IEEE Photon. Technol. Lett. 14(5), 600–602 (2002). [CrossRef]  

31. T. Chu, N. Fujioka, and M. Ishizaka, “Compact, lower-power-consumption wavelength tunable laser fabricated with silicon photonic-wire waveguide micro-ring resonators,” Opt. Express 17(16), 14063–14068 (2009). [CrossRef]   [PubMed]  

32. B. Liu, A. Shakouri, and J. E. Bowers, “Passive microring-resonator-coupled lasers,” Appl. Phys. Lett. 79(22), 3561–3563 (2001). [CrossRef]  

33. O. Schwelb and I. Chremmos, “Band-limited microresonator reflectors and mirror structures,” in Photonic microresonator research and applications, I. Chremmos, O. Schwelb, and N. Uzunoglu, eds. (Springer Series in Optical Sciences, 2010), pp. 139–163. [CrossRef]  

34. C. Alonso-Ramos, A. Ortega-Moñux, I. Molina-Fernández, A. Annoni, A. Melloni, M. Strain, M. Sorel, P. Orlandi, P. Bassi, and F. Morichetti, “Silicon-on-insulator single channel-extraction filter for DWDM applications,” in IEEE 11th International Conference on Group IV Photonics (GFP) (IEEE, 2014), pp. 219–220.

35. C. Alonso-Ramos, A. Annoni, A. Ortega-Moñux, I. Molina-Fernández, M. Strain, P. Orlandi, P. Bassi, F. Morichetti, and A. Melloni, “Narrow-band single-channel filter in silicon photonics,” in Advanced Photonics for Communications, OSA Technical Digest (online) (Optical Society of America, 2014), paper JT3A.32.

36. X. Wang, W. Shi, H. Yun, S. Grist, N. A. F. Jaeger, and L. Chrostowski, “Narrow-band waveguide Bragg gratings on SOI wafers with CMOS-compatible fabrication process,” Opt. Express 20(14), 15547–15558 (2012). [CrossRef]   [PubMed]  

37. X. Wang, “Silicon photonic waveguide Bragg gratings,” PhD thesis, University of British Columbia (2013).

38. W. Shi, X. Wang, W. Zhang, H. Yun, C. Lin, L. Chrostowski, and N. A. F. Jaeger, “Grating-coupled silicon microring resonators,” Appl. Phys. Lett. 100(12), 121118 (2012). [CrossRef]  

39. R. Boeck, W. Shi, L. Chrostowski, and N. A. F. Jaeger, “FSR-eliminated Vernier racetrack resonators using grating-assisted couplers,” IEEE Photon. J. 5(5), 2202511 (2013). [CrossRef]  

40. K. Ikeda, M. Nezhad, and Y. Fainman, “Wavelength selective coupler with vertical gratings on silicon chip,” Appl. Phys. Lett. 92(20), 201111 (2008). [CrossRef]  

41. W. Shi, H. Yun, C. Lin, M. Greenberg, X. Wang, Y. Wang, S. T. Fard, J. Flueckiger, N. A. F. Jaeger, and L. Chrostowski, “Ultra-compact, flat-top demultiplexer using anti-reflection contra-directional couplers for CWDM networks on silicon,” Opt. Express 21(6), 6733–6738 (2013). [CrossRef]   [PubMed]  

42. W. Shi, X. Wang, C. Lin, H. Yun, Y. Liu, T. Baehr-Jones, M. Hochberg, N. A. F. Jaeger, and L. Chrostowski, “Silicon photonic grating-assisted, contra-directional couplers,” Opt. Express 21(3), 3633–3650 (2013). [CrossRef]   [PubMed]  

43. J. St-Yves, H. Bahrami, S. LaRochelle, and W. Shi, “Widely bandwidth-tunable broadband optical filter on silicon,” in Optical Fiber Communication Conference, OSA Technical Digest (online) (Optical Society of America, 2015), paper Th1F.2.

44. M. Caverley, R. Boeck, L. Chrostowski, and N. A. F. Jaeger, “High-speed data transmission through silicon contra-directional grating coupler optical add-drop multiplexers,” in CLEO: 2015, OSA Technical Digest (online) (Optical Society of America, 2015), paper JTh2A.41.

45. W. Shi, X. Wang, W. Zhang, H. Yun, N. A. F. Jaeger, and L. Chrostowski, “Integrated microring add-drop filters with contradirectional couplers,” in Conference on Lasers and Electro-Optics 2012, OSA Technical Digest (Optical Society of America, 2012), paper JW4A.91.

46. W. Shi, X. Wang, W. Zhang, L. Chrostowski, and N. A. F. Jaeger, “Contradirectional couplers in silicon-on-insulator rib waveguides,” Opt. Lett. 36(20), 3999–4001 (2011). [CrossRef]   [PubMed]  

47. R. Boeck, M. Caverley, L. Chrostowski, and N. A. F. Jaeger, “Process calibration method for designing silicon-on-insulator contra-directional grating couplers,” Opt. Express 23(8), 10573–10588 (2015). [CrossRef]   [PubMed]  

48. P. Orlandi, P. Velha, M. Gnan, P. Bassi, A. Samarelli, M. Sorel, M. J. Strain, and R. M. De La Rue, “Microring resonator with wavelength selective coupling in SOI,” in Proceedings of 8th IEEE International Conference on Group IV Photonics (IEEE, 2011), pp. 281–283.

49. J.-P. Weber, “Spectral characteristics of coupled-waveguide Bragg-reflection tunable optical filter,” IEE Proc. J. Optoelectron. 140(5), 275–284 (1993). [CrossRef]  

50. R. J. Bojko, J. Li, L. He, T. Baehr-Jones, M. Hochberg, and Y. Aida, “Electron beam lithography writing strategies for low loss, high confinement silicon optical waveguides,” J. Vac. Sci. Technol. B 29(6), 06F309 (2011). [CrossRef]  

51. Y. Wang, J. Flueckiger, C. Lin, and L. Chrostowski, “Universal grating coupler design,” Proc. SPIE 8915, 89150Y (2013). [CrossRef]  

52. Y. Zhang, S. Yang, A. E.-J. Lim, G.-Q. Lo, C. Galland, T. Baehr-Jones, and M. Hochberg, “A compact and low loss Y-junction for submicron silicon waveguide,” Opt. Express 21(1), 1310–1316 (2013). [CrossRef]   [PubMed]  

53. Y. Zhang, S. Yang, and T. Baehr-Jones, “Compact and low loss Y-junction for submicron silicon waveguide,” U.S. Patent, US 2014/0178005 A1 (2014).

54. A. Yariv and P. Yeh, Photonics: optical electronics in modern communications (Oxford University, Inc., 2007).

55. L. Chrostowski and M. Hochberg, Silicon photonics design: from devices to systems (Cambridge University, 2015).

56. R. Boeck, M. Caverley, L. Chrostowski, and N. A. F. Jaeger, “Silicon quadruple series-coupled Vernier racetrack resonators: experimental signal quality,” in Optical Fiber Communication Conference, OSA Technical Digest (online) (Optical Society of America, 2015), paper W2A.8.

57. S. K. Selvaraja, P. De Heyn, G. Winroth, P. Ong, G. Lepage, C. Cailler, A. Rigny, K. K. Bourdelle, W. Bogaerts, D. Van Thourhout, J. Van Campenhout, and P. Absil, “Highly uniform and low-loss passive silicon photonics devices using a 300mm CMOS platform,” in Optical Fiber Communication Conference, OSA Technical Digest (online) (Optical Society of America, 2014), paper Th2A.33.

58. A. Melloni, M. Martinelli, G. Cusmai, and R. Siano, “Experimental evaluation of ring resonator filters impact on the bit error rate in non return to zero transmission systems,” Opt. Commun. 234(1–6), 211–216 (2004). [CrossRef]  

59. N. N. Klimov, T. Purdy, and Z. Ahmed, “On-chip silicon photonic thermometers: from waveguide Bragg grating to ring resonators sensors,” Proc. SPIE 9486, 948609 (2015). [CrossRef]  

60. N. Klimov, T. Purdy, and Z. Ahmed, “Fabry-Perrot cavity-based silicon photonic thermometers with ultra-small footprint and high sensitivity,” in Advanced Photonics 2015, OSA Technical Digest (online) (Optical Society of America, 2015), paper SeT4C.4.

61. G.-D. Kim, H.-S. Lee, C.-H. Park, S.-S. Lee, B. T. Lim, H. K. Bae, and W.-G. Lee, “Silicon photonic temperature sensor employing a ring resonator manufactured using a standard CMOS process,” Opt. Express 18(21), 22215–22221 (2010). [CrossRef]   [PubMed]  

62. J. F. Tao, H. Cai, Y. D. Gu, J. Wu, and A. Q. Liu, “Demonstration of a photonic-based linear temperature sensor,” IEEE Photon. Technol. Lett. 27(7), 767–769 (2015). [CrossRef]  

63. S.-L. Tsao, S.-G. Lee, P.-C. Peng, and M.-C. Chen, “Silicon-on-insulator optical waveguide Michelson interferometer sensor for temperature monitoring,” U.S. Patent, US 6,603,559 B2 (2003).

64. S. J. Mason, “Feedback theory-further properties of signal flow graphs,” Proc. IRE 44(7), 920–926 (1956). [CrossRef]  

65. C. Chaichuay, P. P. Yupapin, and P. Saeung, “The serially coupled multiple ring resonator filters and Vernier effect,” Opt. Appl. 39(1), 175–194 (2009).

66. R. Boeck, J. Flueckiger, L. Chrostowski, and N. A. F. Jaeger, “Experimental performance of DWDM quadruple Vernier racetrack resonators,” Opt. Express 21(7), 9103–9112 (2013). [CrossRef]   [PubMed]  

67. R. Boeck, “Silicon ring resonator add-drop multiplexers,” Master’s thesis, University of British Columbia (2011).

68. E. J. Klein, “Densely integrated microring-resonator based components for fiber-to-the-home applications,” PhD thesis, University of Twente (2007).

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 (10)

Fig. 1
Fig. 1 (a) Diagram of our reflector. (b) Diagram showing a section of our contra-DC with anti-reflection gratings [39].
Fig. 2
Fig. 2 (a) Theoretical contra-DC power coupling factor versus wavelength (the red dot indicates the value used for the co-directional power coupling factor of the reflector without gratings). (b) Theoretical contra-DC power transmission factor versus wavelength (the red dot indicates the value used for the co-directional power transmission factor of the reflector without gratings). Theoretical comparison of (c) the reflect port spectrum and (d) the through port spectrum of our reflector with contra-DCs and with co-directional couplers without gratings.
Fig. 3
Fig. 3 (a) Measured reflect port spectrum and (b) the spectrum at wavelengths near the major peak (the red dot indicates the wavelength at which the data in Fig. 5(a) was sent through the device).
Fig. 4
Fig. 4 (a) Measured through port spectrum at wavelengths containing the major notch and one of the minor notches. (b) Measured through port dispersion at wavelengths near the minor notch (the green dot indicates the wavelength at which the data in Fig. 5(b) was sent through the device).
Fig. 5
Fig. 5 (a) Measured reflect port eye diagram for a 12.5 Gbps signal operating at a wavelength of 1535.08 nm. (b) Measured through port eye diagram for a 12.5 Gbps signal operating at a wavelength of 1539.42 nm.
Fig. 6
Fig. 6 (a) Measured through port eye diagram for a 12.5 Gbps signal and (b) measured reflect port eye diagram for a 12.5 Gbps signal both operating at the wavelength corresponding to the center of the major notch and are on the same scale.
Fig. 7
Fig. 7 (a) The reflect port spectrum of our device that has a gap distance of 300 nm measured at different temperatures where from left to right the peaks correspond to 15°C, 25°C, 35°C, 45°C, and 55°C. (b) Resonant peak wavelength versus temperature (dots indicate the measured results and the red line is the fit).
Fig. 8
Fig. 8 Diagram showing the loop gains, GL1 and GL2.
Fig. 9
Fig. 9 Diagrams showing the two forward path gains, GP1 and GP2, from input port to reflect port.
Fig. 10
Fig. 10 Diagrams showing the four forward path gains, GP3, GP4, GP5, and GP6, from input port to through port.

Tables (1)

Tables Icon

Table 1 Comparison of FSR-free reflectors.

Equations (15)

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

G L 1 = G L 2 = t cd 2 X r ,
t cd = s e j Δ β 2 L cd s cosh ( s L cd ) + j Δ β 2 sinh ( s L cd ) e j β b L cd α b L cd ,
Δ = 1 ( G L 1 + G L 2 ) + G L 1 G L 2 .
G P 1 = G P 2 = κ cd 2 X r 1 2 X m z i 2 K 2 2 ,
Δ P 1 = 1 G L 2 ,
Δ P 2 = 1 G L 1 ,
κ cd = j κ sinh ( s L cd ) s cosh ( s L cd ) + j Δ β 2 sinh ( s L cd ) .
T F reflect = G P 1 Δ P 1 + G P 2 Δ P 2 Δ .
G P 3 = G P 4 = t c d b u s X m z i 2 K 2 2 ,
Δ P 3 = Δ P 4 = Δ ,
G P 5 = G P 6 = κ cd 2 t cd X r X m z i 2 K 2 2 ,
Δ P 5 = 1 G L 2 ,
Δ P 6 = 1 G L 1 .
t cd bus = s e j Δ β 2 L cd s cosh ( s L cd ) + j Δ β 2 sinh ( s L cd ) e j β a L cd α a L cd ,
TF through = G P 3 Δ P 3 + G P 4 Δ P 4 + G P 5 Δ P 5 + G P 6 Δ P 6 Δ .
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.