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

Study of resonant mode coupling in the transverse-mode-conversion based resonator with an anti-symmetric nanobeam Bragg reflector

Open Access Open Access

Abstract

The traveling-wave like Fabry-Perot (F-P) resonators based on transverse-mode-conversion have been extensively studied as on-chip filters. However, the incomplete transverse mode conversion will lead to the coupling between two degenerated resonant modes, which brings additional loss and may further induce the resonance splitting. In this paper, we take the transverse-mode-conversion based resonator with anti-symmetric nanobeam Bragg reflector as an example and study the resonant mode coupling in both the direct-coupled and side-coupled resonators. The coupled mode equations are used to model the incomplete transverse mode conversion of Bragg reflector. The resonant mode coupling can be effectively suppressed by carefully designing the phase shifter length and adding the tapered holes. The insertion loss of less than −1 dB can be achieved in the simulation using the two methods. This work is believed to benefit the design of mode-conversion based resonators with low insertion loss and non-splitting line shape.

© 2023 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Silicon photonics has become one of the most commonly used integration platforms due to the advantages of high bandwidth, low cost, high integration density, and compatibility with complementary metal-oxide-semiconductor (CMOS) fabrication technology [1,2]. Optical resonators play an important role in the success of silicon photonics because the high refractive index contrast of silicon waveguide makes the optical resonators extremely compact [3,4]. The silicon optical resonators can be employed in many applications including optical filters, electro-optical modulators, add-drop multiplexers, optical delay lines and label-free biosensors. As a traveling-wave resonator, microring resonator (MRR) is the most popular geometry for its advantages including simple structure, small footprint and no intrinsic reflection [5]. However, the bending loss of waveguide becomes large when the radius is small. Therefore, it’s hard for a single MRR to achieve a wide spectrum range (FSR) with high quality factor (Q-factor) [57]. On the other hand, the Fabry-Perot (F-P) resonator based on the photonic crystal structures or Bragg gratings can easily realize a wide FSR or an FSR-free response with little effect on the internal loss. But the conventional F-P resonator is a standing-wave resonator so the backward wave will cause the reflection to the input port [810]. Recently, the multi-mode Bragg reflectors enabling the coupling between two different transverse modes have attracted great interest and have been well studied [1115]. For example, the Bragg reflector with anti-symmetric periodic structure can make the incident fundamental transverse-electric mode (TE0) reflected and converted to the first order transverse-electric mode (TE1), and vice versa [11,12]. Using such mode-conversion based Bragg reflectors, several traveling-wave-like F-P resonators have been proposed and demonstrated as optical filters, add-drop multiplexers and modulators [1630]. Taking the F-P resonator with anti-symmetric Bragg reflectors as an example, when the input light is directly coupled to the resonator, the reflected TE1 mode can be radiated by adding an adiabatic taper [19,29], or dropped by adding an asymmetric directional coupler [26,27]. When the input light in the bus waveguide is side-coupled to the forward TE1 mode in the resonator, the backward TE0 mode in the resonator cannot be coupled to the bus waveguide due to the large effective index difference [23,24]. Therefore, this type of F-P resonator can avoid the reflection to the input port. It is an attractive configuration to realize an FSR-free response with high Q-factor.

However, it’s still challenging to realize a narrow bandwidth, i.e., high Q-factor, for the transverse-mode-conversion based F-P resonator because of the coupling between the two degenerated resonant modes, which is caused by the incomplete transverse mode conversion, i.e., part of the incident transverse mode is reflected to the same mode. In this case, the resonator becomes a second order filter. The coupling between the resonant modes will bring additional loss and reflection to the input port, and even cause resonance splitting, which has been observed in both simulated and experimental results [12,16,19,2124]. This phenomenon becomes much more severe when using the Bragg reflector with strong perturbation of refractive index, i.e., large coupling coefficient, such as etching circular holes in the multimode waveguide. Up to now, there is few research on the suppression of coupling between the resonant modes [15]. In this paper, the coupled mode theory (CMT) is used to model the incomplete transverse mode conversion in the Bragg reflector. It can be found that the coupling between the resonant modes can be greatly suppressed by carefully designing the phase shifter length of the resonator and adding smaller or tapered holes at the interface between the phase shifter and Bragg reflector. It should be noted that the focus of this paper is on the resonant mode coupling caused by the structure itself rather than the fabrication error.

2. Principle

2.1 Description of the structure

The F-P resonator discussed in this paper utilizes anti-symmetric nanobeam Bragg reflectors to realize the coupling between TE0 and TE1 modes, which is one of the most common configurations of transverse-mode-conversion based resonators [12,19,20,23,27]. It consists of one phase shifter and two Bragg reflectors, as shown in Fig. 1. The input light can be directly coupled to the resonator or side-coupled to the resonator through the directional coupler.

 figure: Fig. 1.

Fig. 1. Schematics of the (a) direct-coupled and (b) side-coupled F-P resonators with anti-symmetric nanobeam Bragg reflector. The inserted figure is the zoom-in view of the Bragg reflector. Here w, r and Λ denote the multimode waveguide width, hole radius and grating period, respectively.

Download Full Size | PDF

The schematic of the direct-coupled resonator is shown in Fig. 1(a), which can be used as a narrow pass-band filter. The input and through ports with single mode waveguides are connected to the resonator with multimode waveguide via two adiabatic tapers. The input light will excite the resonance formed by the conversion between the forward TE0 mode and backward TE1 mode. The light at resonant wavelength can nearly fully travel through the resonator. It should be noted that the reflected TE1 mode is radiated in the adiabatic taper and single mode waveguide so the reflection to the input port can be avoided.

The schematic of the side-coupled resonator is shown in Fig. 1(b), the resonator with multimode waveguide is side-coupled to two single mode waveguides so as to form an add-drop filter. The taper on each end of the resonator is used to radiate the residual transmitted light and avoid reflection at the facet. In addition, the waveguide widths should be designed following the phase match condition that the effective index of the TE0 mode in the single waveguide is equal to that of the TE1 mode in the multimode waveguide to make the two modes efficiently coupled. The input light will excite the resonance formed by the conversion between the forward TE1 mode and the backward TE0 mode. It should be noted that the backward TE0 mode cannot be coupled to the single mode waveguide due to the large effective index difference. The light at resonant wavelength can be removed from the bus and nearly fully dropped.

2.2 Coupled mode analysis

Next, we discuss the coupling of transverse modes in the Bragg reflector. In ideal condition that the perturbation is very weak, we can assume that the modal profile of each transverse mode remains constant. Thus only the contra-directional coupling between the TE0 and TE1 modes exists in the Bragg reflector. However, in practice the index variation in the Bragg reflector is large so the modal profiles severely change in the propagation direction. The mismatches of modal profiles lead to the other unwanted coupling terms, i.e., the contra-directional coupling between the forward and backward TE0 modes, the contra-directional coupling between the forward and backward TE1 modes, as well as the directional coupling between the TE0 and TE1 modes. As a consequence, the coupled mode equations can be written as [3133]

$$\frac{{d{F_0}(z)}}{{dz}} = ({\alpha _0} - i{\delta _0}){F_0} - i{\kappa _{00}}{R_0} - i{\kappa _{01}}{R_1} + i{\mu _{01}}{F_1}$$
$$- \frac{{d{R_0}(z)}}{{dz}} = ({\alpha _0} - i{\delta _0}){R_0} - i{\kappa _{00}}{F_0} - i{\kappa _{01}}{F_1} + i{\mu _{01}}{R_1}$$
$$\frac{{d{F_1}(z)}}{{dz}} = ({\alpha _1} - i{\delta _1}){F_1} - i{\kappa _{11}}{R_1} - i{\kappa _{01}}{R_0} + i{\mu _{01}}{F_0}$$
$$- \frac{{d{R_1}(z)}}{{dz}} = ({\alpha _1} - i{\delta _1}){R_1} - i{\kappa _{11}}{F_1} - i{\kappa _{01}}{F_0} + i{\mu _{01}}{R_0}$$
where the subscripts ‘0’ and ‘1’ refer to the TE0 and TE1 modes, respectively, F and R are the amplitudes of the forward and backward propagating waves, respectively, α is the modal loss, δ is the detuning factor, μ is the directional coupling coefficient, and κ is the contra-directional coupling coefficient. The detuning factor of TEi mode can be expressed as
$${\delta _i} = {n_{eff{\kern 1pt} i}}\frac{{2\pi }}{\lambda } - \frac{\pi }{\Lambda }$$
where neff i is the average effective index of TEi mode, λ is the wavelength and Λ is the grating period. The central wavelength of transverse mode conversion, also known as Bragg wavelength (λB), is determined by δ0 + δ1 = 0, i.e., λB = (neff 0+ neff 1) Λ [14,29].

The solution of the coupled mode equations can be expressed in the following form [31,34]:

$$\left[ {\begin{array}{*{20}{c}} {{F_0}(z + \Delta z)}\\ {{R_0}(z + \Delta z)}\\ {{F_1}(z + \Delta z)}\\ {{R_1}(z + \Delta z)} \end{array}} \right] = {e^{A\Delta z}}\left[ {\begin{array}{*{20}{c}} {{F_0}(z)}\\ {{R_0}(z)}\\ {{F_1}(z)}\\ {{R_1}(z)} \end{array}} \right]$$
where A is a 4 × 4 matrix which can be expressed as
$$A = \left[ {\begin{array}{cccc} {{\alpha_0} - i{\delta_0}}&{ - i{\kappa_{00}}}&{i{\mu_{01}}}&{ - i{\kappa_{01}}}\\ {i{\kappa_{00}}}&{ - {\alpha_0} + i{\delta_0}}&{i{\kappa_{01}}}&{ - i{\mu_{01}}}\\ {i{\mu_{01}}}&{ - i{\kappa_{01}}}&{{\alpha_1} - i{\delta_1}}&{ - i{\kappa_{11}}}\\ {i{\kappa_{01}}}&{ - i{\mu_{01}}}&{i{\kappa_{01}}}&{ - {\alpha_1} + i{\delta_1}} \end{array}} \right]$$

Thus the response of Bragg reflector can be obtained using the transfer matrix method (TMM).

Due to the presence of κ00, κ11 and μ01, the transverse mode conversion becomes incomplete and part of the incident transverse mode is reflected to the same transverse mode, which will lead to the coupling between the two degenerated resonant modes. One resonant mode contains the forward TE1 and backward TE0 modes (denoted as RM1), and the other contains the forward TE0 and backward TE1 modes (denoted as RM2), as shown in Fig. 2(a). Here rij denotes the amplitude reflection coefficient between TEi and TEj modes. The resonant wavelength λr satisfies the following phase match condition:

$$2{\phi _{01}} + ({n_{eff0}} + {n_{eff1}})\frac{{2\pi }}{{{\lambda _r}}}{L_p} = m \cdot 2\pi$$
where ϕ01 is the reflection phase of r01, Lp is the phase shifter length and m is an integer. Considering the resonant mode coupling induced by the reflection between the forward and backward TE0 modes, we can assume that the coupling occurs in the middle of the resonator, as shown in Fig. 2(b). Then the coupling coefficient can be expressed as
$${k_{00}} = |{{k_{00}}} |\exp (i{\varphi _{00}}) = |{{r_{00}}} |\exp [i({\phi _{00}} + {n_{eff0}}\frac{{2\pi }}{{{\lambda _r}}}{L_p})]$$
where φ00 is the phase of k00 and ϕ00 is the phase of r00. Similarly, the resonant mode coupling coefficient induced by the reflection between the forward and backward TE1 modes (shown in Fig. 2(c)) can be expressed as
$${k_{11}} = |{{k_{11}}} |\exp (i{\varphi _{11}}) = |{{r_{11}}} |\exp [i({\phi _{11}} + {n_{eff1}}\frac{{2\pi }}{{{\lambda _r}}}{L_p})]$$
where φ11 is the phase of k11 and ϕ11 is the phase of r11. Then the total resonant mode coupling coefficient krm can be expressed as
$${k_{rm}} = |{{k_{00}} + {k_{11}}} |= \sqrt {{{|{{k_{00}}} |}^2} + {{|{{k_{11}}} |}^2} + 2|{{k_{00}}} ||{{k_{11}}} |\cos ({\varphi _{00}} - {\varphi _{11}})}$$
When φ00φ11 = , where m is odd, krm reaches the minimum value of ||r00|−|r11||, and when m is even, krm reaches the maximum value of ||r00|+|r11||. Therefore, we can adjust the phase shifter length Lp to control the value of (φ00φ11) and further achieve a smaller krm.

 figure: Fig. 2.

Fig. 2. Schematics of (a) the coupling between two resonant modes, (b) the resonant mode coupling induced by the reflection between the forward and backward TE0 modes and (c) the resonant mode coupling induced by the reflection between the forward and backward TE1 modes.

Download Full Size | PDF

3. Resonators with uniform anti-symmetric Bragg reflector

3.1 Reflection spectrum of Bragg reflector

The reflection spectrum of the Bragg reflector is firstly investigated. The length, waveguide width, hole radius and grating period used in the simulation are 10 μm, 837 nm, 40 nm and 326 nm, respectively. The effective indices and coupling coefficients in the Bragg reflector are calculated by simulating the Bragg grating with one period via finite-difference time domain (FDTD) method with the perfectly matched layer (PML) boundaries, as shown in Fig. 3. When the TE0 mode is incident, we record the phase response of the transmitted TE0 mode (Δϕ0), and then the effective index of TE0 mode (neff 0) can be deduced by $\Delta {\phi _0} = {n_{eff0}}\frac{{2\pi }}{\lambda }\Lambda $. Similarly, we can obtain neff 1 when the TE1 mode is incident. The amplitude reflection and transmission coefficients including |t01|, |r00| and |r01| can be obtained when the TE0 mode is incident, as shown in Fig. 3(a). |t01|, |r11| and |r01| can also be obtained when the TE1 mode is incident, as shown in Fig. 3(b). Then the coupling coefficients can be deduced by |t01| = μ01Λ, |r00| = κ00Λ, |r01| = κ01Λ and |r11| = κ11Λ, respectively. The calculated Δϕ0, Δϕ1, |t01|, |r00|, |r01| and |r11| using FDTD method and the deduced neff 0, neff 1, μ01, κ00, κ01 and κ11 at 1550 nm are shown in Table 1. Such large values of κ00, κ11 and μ01 cannot be neglected in the coupled mode analysis. Figure The effective indices and coupling coefficients width different wavelength are also calculated, as shown in Fig. 4(a) and (b), respectively.

 figure: Fig. 3.

Fig. 3. Schematics of the calculation of effective indices and coupling coefficients in the Bragg reflector when (a) the TE0 mode is incident and (b) the TE1 mode is incident.

Download Full Size | PDF

 figure: Fig. 4.

Fig. 4. Calculated (a) effective indices of TE0 and TE1 modes in the Bragg reflector and strip waveguide, and (b) coupling coefficients of Bragg reflector.

Download Full Size | PDF

Tables Icon

Table 1. Calculated effective indices and coupling coefficients with 40-nm hole radius at λ = 1550 nm

The normalized reflection spectra calculated by the FDTD method and CMT are shown in Fig. 5. Here the modal loss is neglected in the calculation using CMT. It can be seen the two curves in each spectrum are well matched, which validates that CMT can be used to model the incomplete transverse mode conversion in the Bragg reflector, though the perturbation is very strong. |r01|2 reaches 0.99 at the Bragg wavelength of 1550 nm. Moreover, |r00|2 and |r11|2 are almost equal within the stopband of Bragg reflector. Both of them are dramatically increased around 1520 nm, which indicates the incomplete transverse mode conversion is much more severe for the shorter wavelengths.

 figure: Fig. 5.

Fig. 5. Calculated normalized reflection spectra of (a) |r01|2, (b) |r00|2 and (c) |r11|2 using 3D FDTD method and CMT.

Download Full Size | PDF

The impacts of the hole radius on the coupling coefficients are also investigated. Figure 6(a) shows the coupling coefficients at 1550 nm with different hole radius. Here the grating period is varied with the hole radius to keep the Bragg wavelength at 1550 nm. Figure 6(b) and (c) show the calculated |r01|2 and |r00|2 with different hole radius using CMT, respectively. For easy comparison, the length of Bragg reflector (LB) is varied with the hole radius to keep κ01 LB = 3.5. As the hole radius increases, |r00|2 increases and the maximum value of |r01|2 slightly decreases. Moreover, when the hole radius is 50 μm and 60 μm, the maximum point of |r01|2 is apparently red shifted due to the severe incomplete mode conversion. Therefore, to increase |r01|2 and decrease |r00|2, the phase shifter length of the resonator should be designed to make the resonant wavelength red shifted from the Bragg wavelength when using a large hole radius. Using a smaller hole radius can also effectively decrease |r00|2. This is because the perturbation is weakened, which reduces the impact of the unwanted couplings between the transverse modes. However, a small hole radius also leads to a small κ01, which requires a large Bragg reflector length to achieve a sufficient power reflection of |r01|2. Therefore, the hole radius is chosen to be 40 nm in the following simulations.

 figure: Fig. 6.

Fig. 6. Calculated (a) coupling coefficients with different hole radius at 1550 nm, as well as (b) |r01|2 and (c) |r00|2 with different hole radius using CMT when keeping κ01 LB = 3.5.

Download Full Size | PDF

3.2 Performance of the resonator

The resonant wavelength and the resonant mode coupling coefficient (krm) with different phase shifter length (LP) are calculated by Eqs. (8) and (11), respectively, as shown in Fig. 7. The Bragg reflector length used in the calculation is 10 μm. The resonant wavelength equals 1550 nm when Lp equals m·0.96Λ, where m is an integer, since the effective indices of the TE0 and TE1 modes in the phase shifter are higher than that in the Bragg reflector. When LP equals 0, the value of (φ00φ11) is −0.1π, so the value of krm is up to 0.1. When LP increases to 1.88 μm (6 × 0.96Λ), the value of (φ00φ11) is 1.05π, and krm decreases to 0.01. When LP further increases to near 2.0 μm, the resonant wavelength is red shifted and krm can decrease to a small value close to 0. However, |r01|2 may decrease when the resonant wavelength deviates from the Bragg wavelength, which will lead to the decrease of Q-factor.

 figure: Fig. 7.

Fig. 7. Calculated resonant mode coupling coefficient (krm) and resonant wavelength versus phase shifter length (Lp).

Download Full Size | PDF

The spectra of direct-coupled and side-coupled resonators are simulated using 3D FDTD method. In the side-coupled resonator, the two bus waveguides are made in arc shape in the directional coupling region with the same bending radius of 10 μm and minimum waveguide spacing of 250 nm. The Bragg reflector lengths of the direct-coupled and side-coupled resonators are 6.5 μm and 10 μm, respectively. Such difference of Bragg reflector length is used to keep the bandwidths or the Q-factors of the two resonators approximately equal. When LP is 0, the calculated |r00|2 and |r01|2 of the 6.5-μm long Bragg reflector are both about 2.6 × 10−3, and the calculated (φ00φ11) is about −0.1π. These values are very close to that of the 10-μm long Bragg reflector. Therefore, the calculated resonant mode coupling coefficients (krm) shown in Fig. 7 is also suitable for the direct-coupled resonator with 6.5-μm long Bragg reflectors.

The calculated spectra of the direct-coupled resonator with different LP are shown in Fig. 8. Due to the large value of krm, an apparent resonance splitting can be observed when LP is 0, as shown in Fig. 8(a). When LP is 1.57 μm (5 × 0.96Λ), the resonance splitting becomes weak, as shown in Fig. 8(b). The resonance splitting vanishes when LP is 1.88 μm (6 × 0.96Λ), the insertion loss and 3-dB bandwidth of through port are 0.51 dB and 0.89 nm, respectively, as shown in Fig. 8(c). The corresponding Q-factor is 1.74 × 103. When LP is 1.96 μm (6.25 × 0.96Λ), the resonant wavelength is shifted to around 1565 nm, the insertion loss and 3-dB bandwidth of through port are 0.23 dB and 1.21 nm, respectively, as shown in Fig. 8(d). The corresponding Q-factor is 1.29 × 103.

 figure: Fig. 8.

Fig. 8. Calculated spectra of the direct-coupled resonators with uniform Bragg reflector at the two ports when Lp are (a) 0, (b) 1.57 μm (5 × 0.96Λ), (c) 1.88 μm (6 × 0.96Λ) and (d) 1.96 μm (6.25 × 0.96Λ), respectively.

Download Full Size | PDF

The calculated spectra of the side-coupled resonator with different LP are shown in Fig. 9. Resonance splitting can be observed when LP is 0 and 1.57 μm, as shown in Fig. 9(a) and (b), respectively. The resonance splitting vanishes when LP is 1.88 μm. The extinction ratio of through port, as well as the insertion loss and 3-dB bandwidth of drop port are −16.6 dB, −1.31 dB and 0.91 nm, respectively, as shown in Fig. 9(c). The corresponding Q-factor is 1.70 × 103. Besides, the transmission at add port and the reflection at input port are both decreased to around −22 dB. When LP is 1.96 μm, the resonant wavelength is shifted to around 1565 nm, the extinction ratio, insertion loss and 3-dB bandwidth are 16.0 dB, 1.48 dB and 1.03 nm, respectively, as shown in Fig. 9(d). The corresponding Q-factor is 1.52 × 103. The transmission at add port and the reflection at input port are further decreased to around −28 dB.

 figure: Fig. 9.

Fig. 9. Calculated spectra of the side-coupled resonator at the four ports when Lp are (a) 0, (b) 1.57 μm (5 × 0.96Λ), (c) 1.88 μm (6 × 0.96Λ) and (d) 1.96 μm (6.25 × 0.96Λ), respectively.

Download Full Size | PDF

From the simulations we can see that the coupling between the two resonant modes can severely affect the shape, linewidth and insertion loss of the spectrum, and lead to crosstalk at other ports. The resonant mode coupling can be greatly suppressed by carefully designing the phase shifter length, which agrees with the calculation of krm shown in Fig. 7.

4. Resonators with tapered anti-symmetric Bragg reflector

Adding tapered holes in the initial position of Bragg reflector can effectively suppress the unwanted reflections of |r00|2 and |r11|2, and further decrease the coupling between the two resonant modes [15,16,22]. However, the previously reported papers didn’t explain whether and how the optimal solution of the tapered holes is obtained. Here we add the holes with 20-nm radius and holes with 30-nm radius, with period numbers of n1 and n2, respectively, as shown in the inserted figure of Fig. 10(a). The length of the holes with 40-nm radius is 10 μm. The periods of the holes with 20-nm radius and 30-nm radius are 316.5 nm and 320.5 nm respectively to keep the Bragg wavelength at 1550 nm. The TMM based on CMT is used to calculate |r00|2 at 1550 nm with different n1 and n2, as shown in Fig. 10(a). It has much higher efficiency than FDTD method due to the lower time consumption. In order to achieve a low |r00|2 with a shorter length of tapered holes, n1 and n2 are both chosen to be 3, which is marked with a red dot in Fig. 10(a). Then |r00|2 and |r11|2 are calculated using FDTD method, as shown in Fig. 10(b). It can be seen that |r00|2 and |r11|2 at 1550 nm decreases to 3.2 × 10−5 and 2.7 × 10−5 (∼ −45 dB), respectively. According to Eq. (11), the maximum value of krm at resonant wavelength of 1550 nm is only 0.012, which is much lower than that of the resonator with uniform Bragg reflector. Due to the tapered holes, the resonant wavelength equals 1550 nm when Lp= 20 nm + m·0.96Λ, where m is an integer and Λ is the period of holes with 40-nm radius. The calculated (φ00φ11) is about −0.3π when LP is 20 nm.

 figure: Fig. 10.

Fig. 10. (a) Calculated |r00|2 at 1550 nm with different n1 and n2 using TMM bases on CMT, and (b) calculated |r00|2 and |r11|2 spectra using FDTD method when n1 = n2 = 3, which is marked with a red dot in (a).

Download Full Size | PDF

Except for the added smaller holes, the other structures of the direct-coupled and side-coupled resonators are the same as that of the resonators with uniform anti-symmetric Bragg reflector. Considering the strongest and weakest coupling between the resonant modes, the two cases that LP equals 0.65 μm (20 nm + 2 × 0.96Λ) and 2.22 μm (20 nm + 7 × 0.96Λ) are calculated, with the corresponding (φ00φ11) of 0.08π and 1.03π, respectively. The calculated spectra of the direct-coupled and side-coupled resonators with different LP are shown in Fig. 11 and Fig. 12, respectively.

 figure: Fig. 11.

Fig. 11. Calculated spectra of the direct-coupled resonators with tapered Bragg reflector at the two ports when Lp are (a) 0.65 μm (20 nm + 2 × 0.96Λ) and (b) 2.22 μm (20 nm + 7 × 0.96Λ), respectively.

Download Full Size | PDF

 figure: Fig. 12.

Fig. 12. Calculated spectra of the side-coupled resonators with tapered Bragg reflector at the four ports when Lp are (a) 0.65 μm (20 nm + 2 × 0.96Λ) and (b) 2.22 μm (20 nm + 7 × 0.96Λ), respectively.

Download Full Size | PDF

For the direct-coupled resonator, when LP is 0.65 μm, the insertion loss and 3-dB bandwidth are −2.12 dB and 0.63 nm, respectively, as shown in Fig. 11(a). The corresponding Q-factor is 2.38 × 103. The reflection at the input port is −8.29 dB. When LP is 2.22 μm, a better performance with insertion loss of −0.29 dB and 3-dB bandwidth of 0.36 nm is achieved, as shown in Fig. 11(b). The corresponding Q-factor is 4.31 × 103. The reflection to the input port is also greatly decreased, which indicates the coupling between the resonant modes is effectively suppressed.

For the side-coupled resonator, when LP is 0.65 μm, the extinction ratio, insertion loss and 3-dB bandwidth are −19.7 dB, −0.89 dB and 0.98 nm, respectively, as shown in Fig. 12(a). The corresponding Q-factor is 1.58 × 103. The transmission at add port and the reflection at input port are both around −15 dB. When LP is 2.22 μm, the extinction ratio, insertion loss and 3-dB bandwidth are −22.5 dB, −0.55 dB and 0.65 nm, respectively, as shown in Fig. 12(b). The corresponding Q-factor is 2.38 × 103. The transmission at add port and the reflection at input port are both decreased to around −28 dB.

From the simulations we can see that adding the tapered holes is an effective way to suppress the coupling between the resonant modes. Due to the small values of |r00|2 and |r11|2, the resonance splitting is not observed even in the worst case that (φ00φ11) is close to 0. However, an optimal response still requires the careful design of LP to further suppress the coupling between the resonant modes.

5. Conclusion

In summary, we have studied the coupling between the two degenerated resonant modes in the transverse-mode-conversion based resonator with anti-symmetric nanobeam Bragg reflector, which is induced by the incomplete mode-conversion of the Bragg reflector. Such incomplete mode-conversion can be modeled by the coupled mode equations in which all the coupling coefficients are considered. We have also explained that the coupling between the two resonant modes can be effectively suppressed by carefully designing the phase shifter length and adding the tapered holes. This study is believed to benefit the design of transverse-mode-conversion based resonator filters which have great potential for applications in wavelength-division multiplexing networks.

Funding

National Natural Science Foundation of China (61975075, 52202241); the Special Foundation for State Major Basic Research Program of China (2018YFE0201200, 2017YFA0206401, 2020YFB2205800).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. R. Soref, “The past, present, and future of silicon photonics,” IEEE J. Sel. Top. Quantum Electron. 12(6), 1678–1687 (2006). [CrossRef]  

2. J. You, Y. Luo, J. Yang, J. Zhang, K. Yin, K. Wei, X. Zheng, and T. Jiang, “Hybrid/integrated silicon photonics based on 2D materials in optical communication nanosystems,” Laser & Photon. Rev. 14(12), 2000239 (2020). [CrossRef]  

3. W. Bogaerts, P. De Heyn, T. Van Vaerenbergh, K. De Vos, S. Kumar Selvaraja, T. Claes, P. Dumon, P. Bienstman, D. Van Thourhout, and R. Baets, “Silicon microring resonators,” Laser Photonics Rev. 6(1), 47–73 (2012). [CrossRef]  

4. D. Liu, H. Xu, Y. Tan, Y. Shi, and D. Dai, “Silicon photonic filters,” Microw. Opt. Technol. Lett. 63(9), 2252–2268 (2021). [CrossRef]  

5. Q. Xu, D. Fattal, and R. G. Beausoleil, “Silicon microring resonators with 1.5-microm radius,” Opt. Express 16(6), 4309–4315 (2008). [CrossRef]  

6. D. Liu, C. Zhang, D. Liang, and D. Dai, “Submicron-resonator-based add-drop optical filter with an ultra-large free spectral range,” Opt. Express 27(2), 416–422 (2019). [CrossRef]  

7. A. M. Prabhu, A. Tsay, H. Zhanghua, and V. Vien, “Ultracompact SOI microring add–drop filter with wide bandwidth and wide FSR,” IEEE Photonics Technol. Lett. 21(10), 651–653 (2009). [CrossRef]  

8. C. Henry, R. Kazarinov, H. Lee, N. Olsson, and K. Orlowsky, “A narrow-band Si3N4-SiO2 resonant optical reflector,” IEEE J. Quantum Electron. 23(9), 1426–1428 (1987). [CrossRef]  

9. C. Manolatou, M. J. Khan, S. Fan, P. R. Villeneuve, H. A. Haus, and J. D. Joannopoulos, “Coupling of modes analysis of resonant channel add-drop filters,” IEEE J. Quantum Electron. 35(9), 1322–1331 (1999). [CrossRef]  

10. R. Kazarinov, C. Henry, and N. Olsson, “Narrow-band resonant optical reflectors and resonant optical transformers for laser stabilization and wavelength division multiplexing,” IEEE J. Quantum Electron. 23(9), 1419–1425 (1987). [CrossRef]  

11. H. Qiu, J. Jiang, T. Hu, P. Yu, J. Yang, X. Jiang, and H. Yu, “Silicon add-drop filter based on multimode Bragg sidewall gratings and adiabatic couplers,” J. Lightwave Technol. 35(9), 1705–1709 (2017). [CrossRef]  

12. P. Yu, H. Qiu, R. Cheng, L. Chrostowski, and J. Yang, “High-Q antisymmetric multimode nanobeam photonic crystal cavities in silicon waveguides,” Opt. Express 26(20), 26196–26204 (2018). [CrossRef]  

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

14. H. Qiu, Y. Su, F. Lin, J. Jiang, P. Yu, H. Yu, J. Yang, and X. Jiang, “Silicon add-drop filter based on multimode grating assisted couplers,” IEEE Photon. J. 8(6), 1–8 (2016). [CrossRef]  

15. Q. Huang, K. Jie, Q. Liu, Y. Huang, Y. Wang, and J. Xia, “Ultra-compact, broadband tunable optical bandstop filters based on a multimode one-dimensional photonic crystal waveguide,” Opt. Express 24(18), 20542–20553 (2016). [CrossRef]  

16. Q. Huang, Q. Liu, and J. Xia, “Traveling wave-like Fabry-Perot resonator-based add-drop filters,” Opt. Lett. 42(24), 5158–5161 (2017). [CrossRef]  

17. R. Soref, F. De Leonardis, and V. M. N. Passaro, “Compact resonant 2 × 2 crossbar switch using three coupled waveguides with a central nanobeam,” Opt. Express 29(6), 8751–8762 (2021). [CrossRef]  

18. Y. Wu, Y. Shi, Y. Zhao, L. Li, P. Wu, P. Dai, T. Fang, and X. Chen, “On-chip optical narrowband reflector based on anti-symmetric Bragg grating,” Opt. Express 27(26), 38541–38552 (2019). [CrossRef]  

19. M. Mendez-Astudillo, H. Okayama, and H. Nakajima, “Silicon optical filter with transmission peaks in wide stopband obtained by anti-symmetric photonic crystal with defect in multimode waveguides,” Opt. Express 26(2), 1841–1850 (2018). [CrossRef]  

20. M. Mendez-Astudillo, H. Okayama, and T. Kita, “Fabry-Perot Cavity Using Two Row Photonic Crystal in a Multimode Waveguide,” in Conference on Lasers and Electro-Optics (2019), STh1H.6.

21. M. W. Pruessner, J. B. Khurgin, T. H. Stievater, W. S. Rabinovich, R. Bass, J. B. Boos, and V. J. Urick, “Demonstration of a mode-conversion cavity add–drop filter,” Opt. Lett. 36(12), 2230–2232 (2011). [CrossRef]  

22. Q. Liu, D. Zeng, C. Mei, H. Li, Q. Huang, and X. Zhang, “Integrated photonic devices enabled by silicon traveling wave-like Fabry-Perot resonators,” Opt. Express 30(6), 9450–9462 (2022). [CrossRef]  

23. P. Yu, H. Qiu, T. Dai, R. Cheng, B. Lian, W. Li, H. Yu, and J. Yang, “Ultracompact channel add-drop filter based on single multimode nanobeam photonic crystal cavity,” J. Lightwave Technol. 39(1), 162–166 (2021). [CrossRef]  

24. Y. Zhao, Y. Shi, X. Xiong, L. Hao, S. Liu, R. Xiao, P. Dai, J. Lu, T. Fang, and X. Chen, “Single wavelength resonator based on π phase-shifted antisymmetric Bragg grating,” IEEE Photonics Technol. Lett. 31(16), 1339–1342 (2019). [CrossRef]  

25. H. Okayama, Y. Onawa, D. Shimura, H. Takahashi, H. Yaegashi, and H. Sasaki, “Wavelength filter using twin one-dimensional photonic crystal cavity silicon waveguides,” Electron. Lett. 55(2), 107–109 (2019). [CrossRef]  

26. R. Xiao, Y. Shi, J. Li, P. Dai, C. Ma, M. Chen, Y. Zhao, and X. Chen, “Integrated Bragg grating filter with reflection light dropped via two mode conversions,” J. Lightwave Technol. 37(9), 1946–1953 (2019). [CrossRef]  

27. P. Dong, C. Liu, L. Zhang, D. Dai, and Y. Shi, “Reconfigurable add-drop filter based on an antisymmetric multimode photonic crystal nanobeam cavity in a silicon waveguide,” Opt. Express 30(10), 17332–17339 (2022). [CrossRef]  

28. Y. Ma, Y. Zhao, Y. Shi, L. Hao, Z. Sun, Z. Hong, and X. Chen, “Silicon add-drop multiplexer based on π phase-shifted antisymmetric Bragg grating,” IEEE J. Quantum Electron. 57(4), 1–8 (2021). [CrossRef]  

29. H. Okayama, Y. Onawa, D. Shimura, H. Yaegashi, and H. Sasaki, “Silicon wire waveguide TE0/TE1 mode conversion Bragg grating with resonant cavity section,” Opt. Express 25(14), 16672–16680 (2017). [CrossRef]  

30. Y. Zhao, Y. Shi, P. Dai, S. Liu, L. Hao, Y. Ma, and X. Chen, “Side-coupled Fabry-Perot resonator filter based on dual-waveguide Bragg grating,” J. Lightwave Technol. (2022).

31. C. Riziotis and M. N. Zervas, “Design considerations in optical add/drop multiplexers based on grating-assisted null couplers,” J. Lightwave Technol. 19(1), 92–104 (2001). [CrossRef]  

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

33. A. Yariv, “Coupled-mode theory for guided-wave optics,” IEEE J. Quantum Electron. 9(9), 919–933 (1973). [CrossRef]  

34. Y. Zhao, Y. Shi, P. Dai, R. Xiao, S. Liu, L. Hao, T. Fang, J. Lu, J. Zheng, and X. Chen, “A hybrid transverse mode resonance DFB semiconductor laser,” IEEE J. Sel. Top. Quantum Electron. 25(6), 1–9 (2019). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (12)

Fig. 1.
Fig. 1. Schematics of the (a) direct-coupled and (b) side-coupled F-P resonators with anti-symmetric nanobeam Bragg reflector. The inserted figure is the zoom-in view of the Bragg reflector. Here w, r and Λ denote the multimode waveguide width, hole radius and grating period, respectively.
Fig. 2.
Fig. 2. Schematics of (a) the coupling between two resonant modes, (b) the resonant mode coupling induced by the reflection between the forward and backward TE0 modes and (c) the resonant mode coupling induced by the reflection between the forward and backward TE1 modes.
Fig. 3.
Fig. 3. Schematics of the calculation of effective indices and coupling coefficients in the Bragg reflector when (a) the TE0 mode is incident and (b) the TE1 mode is incident.
Fig. 4.
Fig. 4. Calculated (a) effective indices of TE0 and TE1 modes in the Bragg reflector and strip waveguide, and (b) coupling coefficients of Bragg reflector.
Fig. 5.
Fig. 5. Calculated normalized reflection spectra of (a) |r01|2, (b) |r00|2 and (c) |r11|2 using 3D FDTD method and CMT.
Fig. 6.
Fig. 6. Calculated (a) coupling coefficients with different hole radius at 1550 nm, as well as (b) |r01|2 and (c) |r00|2 with different hole radius using CMT when keeping κ01 LB = 3.5.
Fig. 7.
Fig. 7. Calculated resonant mode coupling coefficient (krm) and resonant wavelength versus phase shifter length (Lp).
Fig. 8.
Fig. 8. Calculated spectra of the direct-coupled resonators with uniform Bragg reflector at the two ports when Lp are (a) 0, (b) 1.57 μm (5 × 0.96Λ), (c) 1.88 μm (6 × 0.96Λ) and (d) 1.96 μm (6.25 × 0.96Λ), respectively.
Fig. 9.
Fig. 9. Calculated spectra of the side-coupled resonator at the four ports when Lp are (a) 0, (b) 1.57 μm (5 × 0.96Λ), (c) 1.88 μm (6 × 0.96Λ) and (d) 1.96 μm (6.25 × 0.96Λ), respectively.
Fig. 10.
Fig. 10. (a) Calculated |r00|2 at 1550 nm with different n1 and n2 using TMM bases on CMT, and (b) calculated |r00|2 and |r11|2 spectra using FDTD method when n1 = n2 = 3, which is marked with a red dot in (a).
Fig. 11.
Fig. 11. Calculated spectra of the direct-coupled resonators with tapered Bragg reflector at the two ports when Lp are (a) 0.65 μm (20 nm + 2 × 0.96Λ) and (b) 2.22 μm (20 nm + 7 × 0.96Λ), respectively.
Fig. 12.
Fig. 12. Calculated spectra of the side-coupled resonators with tapered Bragg reflector at the four ports when Lp are (a) 0.65 μm (20 nm + 2 × 0.96Λ) and (b) 2.22 μm (20 nm + 7 × 0.96Λ), respectively.

Tables (1)

Tables Icon

Table 1. Calculated effective indices and coupling coefficients with 40-nm hole radius at λ = 1550 nm

Equations (11)

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

d F 0 ( z ) d z = ( α 0 i δ 0 ) F 0 i κ 00 R 0 i κ 01 R 1 + i μ 01 F 1
d R 0 ( z ) d z = ( α 0 i δ 0 ) R 0 i κ 00 F 0 i κ 01 F 1 + i μ 01 R 1
d F 1 ( z ) d z = ( α 1 i δ 1 ) F 1 i κ 11 R 1 i κ 01 R 0 + i μ 01 F 0
d R 1 ( z ) d z = ( α 1 i δ 1 ) R 1 i κ 11 F 1 i κ 01 F 0 + i μ 01 R 0
δ i = n e f f i 2 π λ π Λ
[ F 0 ( z + Δ z ) R 0 ( z + Δ z ) F 1 ( z + Δ z ) R 1 ( z + Δ z ) ] = e A Δ z [ F 0 ( z ) R 0 ( z ) F 1 ( z ) R 1 ( z ) ]
A = [ α 0 i δ 0 i κ 00 i μ 01 i κ 01 i κ 00 α 0 + i δ 0 i κ 01 i μ 01 i μ 01 i κ 01 α 1 i δ 1 i κ 11 i κ 01 i μ 01 i κ 01 α 1 + i δ 1 ]
2 ϕ 01 + ( n e f f 0 + n e f f 1 ) 2 π λ r L p = m 2 π
k 00 = | k 00 | exp ( i φ 00 ) = | r 00 | exp [ i ( ϕ 00 + n e f f 0 2 π λ r L p ) ]
k 11 = | k 11 | exp ( i φ 11 ) = | r 11 | exp [ i ( ϕ 11 + n e f f 1 2 π λ r L p ) ]
k r m = | k 00 + k 11 | = | k 00 | 2 + | k 11 | 2 + 2 | k 00 | | k 11 | cos ( φ 00 φ 11 )
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.