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

Expanding the adiabatic design toolbox – more modes, parameters and versatility

Open Access Open Access

Abstract

Adiabatic design principles can be used to improve the performance of many photonic components. The recently published adiabatic optimization method, MODALL, relies on a design rule that guarantees adiabaticity and enables optimization of adiabatic photonic components against multiple dimensions and radiation modes. In this work, MODALL is extended to enable optimization of multi-mode components, optimization against an extra degree of freedom and optimization of modal crosstalk. We present a derivation of these extensions starting from MODALL theory and verify them via the design, fabrication and characterization of a mode multiplexer with ultra-low crosstalk: worst-case <−38 dB and median <−45 dB. These design extensions will aid the adiabatic design optimization of many photonic components including splitters, polarization rotators, interlayer transitions and edge couplers.

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

1. Introduction

As the complexity of photonic circuits increases to meet the high data rates demanded by industry, the fabrication yield, insertion loss and footprint of photonic devices becomes crucial to the success of a photonic product [1,2]. Common photonic operations such as splitting, rotating and mode multiplexing light can be implemented via numerous techniques [36]. Adiabatic theory has the advantage of being inherently robust against typical fabrication variation as devices do not rely upon the relative phase difference between eigenmodes and they may be lengthened to increase robustness [7,8].

Adiabatic expressions for particular photonic devices have been derived in literature [913], though the complexity of devices that can be analyzed in this manner is limited. More generic adiabatic optimization methods have been developed including STA [14], FAQUAD [10], CTAP [15], SUSY [16] and RAC [17]. These methods are typically limited to one or two dimensions, one or two modes, do not optimize against radiation modes or are based on a heuristic adiabatic criterion. Recently, we presented multidimensional optimization of devices via adiabatic loss limiting (MODALL) [1824], where we derived an adiabatic criterion and demonstrated adiabatic optimization in two arbitrary dimensions, verifying the results via simulation. We later presented experimental results for devices including an interlayer photonic transition and an adiabatic 2$\times$2.

In this work, we further develop MODALL and apply it to the design and mass fabrication of a TE$_0$/TE$_1$ adiabatic mode multiplexer. We derive MODALL expressions that enable multi-parameter device optimization for an arbitrary number of modes, optimizing against an extra degree of freedom that does not increase the mode computation space and engineering the crosstalk between two modes. The theory presented in this work is applicable to many photonic devices that support one or more modes including tapers, polarization splitter rotators, polarization splitters, adiabatic 3 dB splitters, adiabatic 2$\times$2s, mode multiplexers, waveguide crossings, edge couplers and interlayer transitions.

2. Adiabatic design method

The design methodology originates from the multiparameter adiabatic design expression, Eq. (29) in [18]:

$$\epsilon_\mu^2 = 4\sum_{m \neq \mu} \left(\frac{\mathbf{\kappa}_{m \mu} (\mathbf{p}_t) \cdot \frac{\textrm{d}\mathbf{p}_t}{\textrm{dz}}}{\Delta \beta_{\mu m} (\mathbf{p}_t)} \right)^2,$$
which is derived from expressions for the local mode coupling between a mode that is the target of optimization, $\mu$, and and the remaining system modes, $m$. In Eq. (1), $\epsilon _\mu ^2$ is constant in $z$ and $\mathbf{\kappa }_{m \mu }$ is the coupling coefficient given by
$$\mathbf{\kappa}_{m \mu} = \left(\frac{\varepsilon_0}{\mu_0}\right)^2 \frac{k_0}{4} \frac{1}{\Delta \beta_{\mu m} (\mathbf{p}_t)} \int_A \mathbf{\hat{e}}_m^* \cdot \mathbf{\hat{e}}_\mu\,\nabla_{\mathbf{p}_t} \varepsilon_r \textrm{d} A,$$
where $\Delta \beta _{\mu m}$ is the propagation constant mismatch between the modes $\mu$ and $m$, and $\mathbf {p}_t$ is the vector of parameters we are optimizing in the optical propagation axis, $z$. The modulus has been dropped from Eq. (1) as we only consider optical materials with $\varepsilon _r \in \mathbb {R}$. A design that adheres to this adiabatic design method has a transmission from input mode $\mu$ to mode $\mu$ along the device of $T_{\mu \mu }(z) \in [1-\epsilon _\mu ^2, 1]$.

2.1 Extension to optimize for multiple modes

In many cases, components require transmitting two modes; for example, a mode multiplexer or adiabatic 2$\times$2 might transmit TE$_0$ and TE$_1$, and a polarization rotator will typically transmit TE$_0$ and convert TM$_0$ to TE$_1$. To handle these multimode structures, we should extend the analysis to include an adiabatic limit for an arbitrary number of modes rather than just a single-mode. Using Eq. (13) from [18] and the analysis preceding it, we generate adiabatic loss constraints for a set of modes, $v_i$

$${\epsilon }_{{v_i}}^2 \buildrel \Delta \over = 4\sum\limits_{m \ne {v_i}} {{{\left( {\frac{{{\mathbf{\kappa} _{m{v_i}}}({{\mathbf p}_t}) \cdot \frac{{\textrm{d}{{\mathbf p}_t}}}{\textrm{d}z}}}{{{\Delta }{\beta _{{v_i}m}}({{\mathbf p}_t})}}} \right)}^2}} \ge \eta _{{v_i}}^2(z),$$
where $\epsilon _{v_i}^2$ is not necessarily constant in $z$ and $\eta ^2_{v_i}(z)$ is the power lost from mode $v_i$. A loss limit for the set of modes is established:
$$\epsilon_v^2 = \max_i \epsilon_{v_i}^2 \ge \eta^2_{v_i}(z)\ \forall\ i,$$
where $\epsilon _v^2$ is constant in $z$—none of the $v_i$ modes will lose more than $\epsilon _v^2$ power anywhere along the device. We may now evaluate Eq. (4) to determine an optimized $\textrm{d} \mathbf {p}_t / \textrm{d} z$ profile. The transmission along a device for a $v_i$ mode is $T_{v_i v_i}(z) \in [1 - \epsilon _v^2, 1]$.

2.2 Extension to center shifting

In our design equation, $\mathbf {p}_t$ may represent an arbitrary set of design parameters; however, a special case occurs for design parameters that result in an invariant cross-section. In integrated optics, one such parameter could be structure center, $C$. When $C$ changes in $z$, all waveguide boundaries are shifted perpendicular to the $z$-axis by the same amount and the permittivity cross-section does not change, requiring a single eigenmode computation to fully solve a cross-section for all $C$. We let $\mathbf {p}_t = (\mathbf {p}, C)^T$ to introduce the lateral center parameter to our design expression and differentiate it from parameters, $\mathbf {p}$, that do not have the permittivity cross-sectional invariance property. Eq. (1) becomes

$$\epsilon^2_{v_i} = 4\sum_{m \neq v_i} \left(\frac{\mathbf{\kappa}^{(\mathbf{p})}_{m v_i}(\mathbf{p}) \cdot \frac{{\textrm{d}\mathbf{p}}}{\textrm{dz}} + \kappa_{m v_i}^{(C)}(\mathbf{p}) \frac{\textrm{d}C}{\textrm{dz}}}{\Delta \beta_{v_i m} (\mathbf{p})} \right)^2,$$
where $\kappa _{m v_i}^{(\mathbf {p})}$ and $\kappa _{m v_i}^{(C)}$ are the coupling coefficients resulting from a shift in center:
$$\begin{aligned}\kappa_{m v_i}^{(\mathbf{p})} = \left(\frac{\varepsilon_0}{\mu_0}\right)^2 \frac{k_0}{4} \frac{1}{\Delta \beta_{v_i m} (\mathbf{p})} \int_A \mathbf{\hat{e}}_m^* \cdot \mathbf{\hat{e}}_{v_i}\,\nabla_{\mathbf{p}} \varepsilon_r \textrm{d} A, \end{aligned}$$
$$\begin{aligned}\kappa_{m v_i}^{(C)} = \left(\frac{\varepsilon_0}{\mu_0}\right)^2 \frac{k_0}{4} \frac{1}{\Delta \beta_{v_i m} (\mathbf{p})} \int_A \mathbf{\hat{e}}_m^* \cdot \mathbf{\hat{e}}_{v_i} \frac{\partial \varepsilon_r}{\partial C} \textrm{d} A. \end{aligned}$$

In this section, we restrict ourselves to optimizing for two modes, $v_1$ and $v_2$, involving an arbitrary number of interacting modes, though it should be possible to extend the principles introduced in this section to optimize for more than two modes.

When $C$ is constant in $z$, the minimum $\epsilon ^2_v$ satisfying both target modes is $\epsilon _v^2 = \max _i \epsilon _{v_i}^2$ as shown in Section 2.1. For the general case when $\textrm{d} C / \textrm{d} z \neq 0$, satisfying the adiabatic design rule is more challenging as we need to find the $\textrm{d} C / \textrm{d} z = C'$ for each position along the device that satisfies $\epsilon _{v_1}^2$ and $\epsilon _{v_2}^2$, and minimizes $\epsilon _v^2$. To do so, we minimize $\epsilon _v^2$ in $\textrm{d} C / \textrm{d} z$ subject to the constraints $\epsilon _{v_1}^2 \leq \epsilon _v^2$ and $\epsilon _{v_2}^2 \leq \epsilon _v^2$ [25]. The Lagrange function for the minimization is

$$\mathcal{L}(C', \lambda_1, \lambda_2) = \epsilon_v^2 + \lambda_1 (\epsilon_{v_1}^2 - \epsilon_v^2 + s_1^2) + \lambda_2 (\epsilon_{v_2}^2 - \epsilon_v^2 + s_2^2)$$
where $\lambda _i$ are Lagrange multipliers and $s_i$ are slack variables used to convert the inequality constraints to equality constraints. Partial derivatives are then taken in $C'$, $\lambda _1$ and $\lambda _2$ with the results set to 0 yielding a system of three equations that may be solved. Three nontrivial cases arise and are summarized along with their corresponding transmission:
$$\begin{array}{l} \textrm{C}.\textrm{1}\;\textrm{Choose}\;\textrm{d}C/\textrm{d}z = \textrm{d}C_{{v_1}}^{(\textrm{min})}/\textrm{d}z\;\textrm{and}\;{\epsilon }_v^2 = {\epsilon }_{{v_1}}^2,\\ \textrm{C}.\textrm{2}\;\textrm{Choose}\;\textrm{d}C/\textrm{d}z = \textrm{d}C_{{v_2}}^{(\textrm{min})}/\textrm{d}z\;\textrm{and}\;{\epsilon }_v^2 = {\epsilon }_{{v_1}}^2,\textrm{and}\\ \textrm{C}.\textrm{3}\;\textrm{Choose}\;{\epsilon }_v^2 = {\epsilon }_{{v_1}}^2 ={\epsilon }_{{v_2}}^2\textrm{and}\;\textrm{d}C/\textrm{d}z \in \{ \textrm{d}{C_{{q_1}}}/\textrm{d}z,\textrm{d}{C_{{q_2}}}/\textrm{d}z\} \end{array}$$

In these cases, $\textrm{d} C_{v_1}^\mathrm {(min)} / \textrm{d} z$ and $\textrm{d} C_{v_2}^\mathrm {(min)} / \textrm{d} z$ are the shifts of center that minimize $\epsilon ^2_{v_1}$ and $\epsilon ^2_{v_2}$ respectively, and $\left \lbrace \textrm{d} C_{q_1} / \textrm{d} z,\textrm{d} C_{q_2} / \textrm{d} z\right \rbrace$ are the pair of solutions resulting from the $\epsilon ^2_{v_1}=\epsilon ^2_{v_2}$ equality condition. The prevailing case at each point in $z$ will be the one with the smallest $\epsilon _v^2$ and where $\lambda _1, \lambda _2, s_1, s_2 \geq 0$.

To resolve cases C.1 and C.2 we must first find the $\textrm{d} C / \textrm{d} z$ that minimizes $\epsilon _{v_1}^2$ and $\epsilon _{v_2}^2$ by taking the derivative in $C'(z)$ of both sides and setting $\mathrm {d} \epsilon _{v_i}^2 / \textrm{d} C'(z) = 0$ to find the extrema. The differential equation describing $\textrm{d} C_{v_i}^\mathrm {(min)} / \textrm{d} z$, the shift in center that minimizes $\epsilon ^2_{v_i}$ is:

$$\frac{\textrm{d} C_{v_i}^\mathrm{(min)}}{\textrm{d} z} = \frac{-\sum_{m \neq v_i}\left[\frac{\kappa_{m v_i}^{(C)}}{\Delta\beta_{v_i m}^2}\mathbf{\kappa}_{m v_i}^{(\mathbf{p})} \cdot \frac{\textrm{d} \mathbf{p}}{\textrm{d} z}\right]}{\sum_{m \neq v_i}\frac{\left(\kappa_{m v_i}^{(C)}\right)^2}{\Delta\beta_{v_i m}^2 }}$$
where $i \in \{1, 2\}$ and the explicit dependence of $\mathbf \kappa$ and $\Delta \beta$ on $\mathbf {p}$ has been dropped for simplicity. We can now use $\textrm{d} C_{v_i}^\mathrm {(min)} / \textrm{d} z$ to evaluate $\epsilon _{v_i}^2$ via Eq. (5).

Case C.3 requires solving

$$0 = \epsilon_{v_1}^2 - \epsilon_{v_2}^2$$
for $\textrm{d} C / \textrm{d} z$. The pair of solutions $\left \lbrace \textrm{d} C_{q_1} / \textrm{d} z,\textrm{d} C_{q_2} / \textrm{d} z\right \rbrace$ to Eq. (9) are computed using the quadratic formula by making use of Eq. (5) to calculate $\epsilon _v^2$—the element of $\left \lbrace \textrm{d} C_{q_1} / \textrm{d} z,\textrm{d} C_{q_2} / \textrm{d} z\right \rbrace$ that results in the lower $\epsilon _v^2$ is chosen and used to solve for $\textrm{d} \mathbf {p}_t / \textrm{d} z$ from Eq. (3) and Eq. (4).

2.3 Extension to crosstalk suppression

Our analysis so far has focused solely on minimizing the adiabatic loss of multiple target modes. When optimizing for two modes, it may be desirable for the scattering between the two modes of interest to be weighted more heavily than the scattering to other modes. For an adiabatic 2$\times$2, we would like to minimize scattering from the TE$_0$ and TE$_1$ modes to all modes; however, we may want to bias the calculation such that it favors minimizing scattering between the TE$_0$ and TE$_1$ modes themselves.

We introduce the crosstalk optimization parameters $F_\mathrm {XT}$ and $F_\mathrm {L}$ via the substitution

$$\sum_{m \neq v_1}f_{m v_1} \to F_\mathrm{XT}f_{v_2 v_1} + F_\mathrm{L}\!\!\sum_{m\neq\{v_1,v_2\}}\!\!f_{m v_1} \mathrm{\ \ \ where\ \ \ } F_\mathrm{XT},F_\mathrm{L} \in [0,1]$$
and
$${f_{m{v_1}}}\textrm{ } \buildrel \Delta \over = {\left( {\frac{{2{\mathbf{\kappa} _{m{v_1}}}({{\mathbf p}_t}) \cdot \frac{{{{\textrm d\mathbf p}_t}}}{{\textrm d z}}}}{{{\Delta }{\beta _{{v_1}m}}({{\mathbf p}_t})}}} \right)^2},$$
and $v_1$ and $v_2$ may be interchanged. This substitution may then be applied to the summation terms in Eq. (8) and Eq. (9) to calculate a $\textrm{d} C / \textrm{d} z$ with an optimized trade off between crosstalk and loss. Crosstalk may be reduced by setting $F_\mathrm {XT}=1$ and $F_\mathrm {L}<1$. The case where $F_\mathrm {XT}=1$ and $F_\mathrm {L}=0$ produces the canceling of coupling coefficients described in Eq. (33) of [18] and when $\theta =\theta _{RAC}$ in [17]. When $F_\mathrm {XT}=1$ and $F_\mathrm {L}=1$, the result in Section 2.2 is reproduced. Tuning these parameters allows for lower modal crosstalk at the expense of increased scattering to other unwanted modes.

We shall now outline some details surrounding what can and cannot be achieved using $F_\mathrm {L}$ tuning. The crosstalk is bounded as follows for the three non-trivial $T_{v_1 v_2}(z)$ cases:

  • 1. if $F_\mathrm {XT}=1$ and $F_\mathrm {L}=1$, $T_{v_1 v_2}(z) \in [0,\epsilon _v^2]$;
  • 2. if $F_\mathrm {XT}=1$ and $F_\mathrm {L}=0$, $T_{v_1 v_2}(z) = 0$; and
  • 3. if $F_\mathrm {XT}=0$ and $F_\mathrm {L}=1$, $T_{v_1 v_2}(z) \in [0,1]$.
If $F_\mathrm {L}>0$, we cannot predict the crosstalk beyond it being in $[0,\epsilon _v^2]$. We know $T_{v_1 v_2}(z)$ must be bound and will be 0 when $F_\mathrm {XT}=0$, but we cannot guarantee a monotonic decrease in crosstalk unless we know that $\mathbf {p}(z)$ is constant in $F_\mathrm {L}$. When maintaining the same path while adjusting $F_\mathrm {L}$, the worst-case crosstalk will decrease monotonically as $F_\mathrm {L}$ decreases, though a more optimal solution may be found by allowing the path to change. The amount the worst-case loss will increase as $F_\mathrm {L}$ decreases is unpredictable and highly dependent on the specific adiabatic transition under investigation—a designer is left only with the option of varying $F_\mathrm {L}$ and checking whether the corresponding adiabatic optimization meets the requirements.

For some devices, including edge couplers and polarization splitter rotators, crosstalk between the two modes of interest may be of little importance. In this case, one should set $F_\mathrm {XT}=0$ and $F_\mathrm {L}=1$ to ensure minimum scattering to unwanted radiation modes, while ensuring the optimization is not restricted by minimizing scattering between the $v_1$ and $v_2$ modes. In cases where crosstalk is critical for most two-mode devices, one should choose $F_\mathrm {XT}=1$ as well as an adequate $F_\mathrm {L} \in [0,1]$.

2.4 Discretization

The equations presented cannot, in general, be resolved analytically, and most useful problems require the equations to be discretized to enable numerical computation. Discretization is implemented by changing the derivative operator to its discrete counterpart, $\textrm{d} \to \Delta$. As examples, Eq. (5) becomes

$$\epsilon_{v_i}^2 = \frac{4}{\Delta z^2}\sum_{m \neq v_i} \left(\frac{1}{\Delta \beta_{v_i m} (\mathbf{p})} \left( \mathbf{\kappa}^{(\mathbf{p})}_{m v_i}(\mathbf{p})\cdot \Delta \mathbf{p} + \kappa_{m v_i}^{(C)}\!(\mathbf{p}) \Delta C \right) \right)^2,$$
and Eq. 8 becomes
$$\Delta C_{v_i}^\mathrm{(min)} = \frac{-\sum_{m \neq v_i}\left[\frac{\kappa_{m v_i}^{(C)}}{\Delta\beta_{v_i m}^2}\mathbf{\kappa}_{m v_i}^{(\mathbf{p})} \cdot \Delta \mathbf{p}\right]}{\sum_{m \neq v_i}\frac{\left(\kappa_{m v_i}^{(C)}\right)^2}{\Delta\beta_{v_i m}^2 }}$$
when discretized. Discretizing all presented equations allows computing an optimized $\mathbf {p}_t(z)$. This may be accomplished using a shortest path algorithm as described in Section 4.1 of [18] to find the path through the $\mathbf {p}$-space that minimizes the total $\Delta z$ sum.

3. Application to a low crosstalk mode multiplexer

To illustrate the adiabatic optimization design process, we apply all three extensions introduced in Section 2 to a mode multiplexer ("modemux") that transmits TE$_1$ to TE$_0$ on waveguide one and transmits TE$_0$ to TE$_0$ on waveguide two. We require the modemux to have a low insertion loss for both modes and have low modal crosstalk across the O-band. Low crosstalk is critical to applications such a two-mode Bragg-based wavelength demultiplexer, where demultiplexed light passes through the modemux twice (once in the forward direction, and once in the reverse direction) [26].

3.1 Design and fabrication

To demonstrate the design strategy, we use a photonic platform comprising 200 nm fully etched SiN clad with SiO$_2$ and fabricated at a foundry on 300 mm Si substrates. The modemux has a 1.7 µm wide input multimode waveguide and a 100 nm tip 200 nm away that adiabatically transition to two single-mode waveguides 900 nm and 1 µm wide separated by a 2.5 µm gap—the slight difference in waveguide widths is needed to avoid degeneracy between the TE$_0$ and TE$_1$ modes. From the modemux cartoon above Fig. 3(b), when TE$_0$ is launched at $E$, ideally all the optical power should transmit to $G$ (no power should transmit to $F$), and when TE$_1$ is launched at $E$, all the optical power should transmit to $F$ (no power should transmit to $G$). At the output side of the modemux, the two single-mode waveguides have a large enough gap not to couple light between them and induce further crosstalk.

The design begins by choosing the parameters to define and optimize the adiabatic structure. The parameters must enable a structure with the aforementioned waveguide width and gap endpoints. Figure 1(a) shows the relevant widths, gap and center used in the design: $p_{w_1}$ is the width of the narrower waveguide, $p_{w_2}$ is the width of the wider waveguide, $p_g$ is the gap between the waveguides and $C$ is the center point of the structure. We let $\mathbf {p}=[p_w,p_g]^T$ where $p_{w_1}=p_w$ and $p_{w_2}=-0.875 p_w + 1.7875$. These parameters let us solve a $\dim {\mathbf {p}} = 2$ ($\dim {\mathbf {p}_t} = 3$) problem that will satisfy our design endpoints.

 figure: Fig. 1.

Fig. 1. (a) Cartoon of the design structure with the critical parameters labeled. (b) The shortest path through the optimization space calculated using a discrete shortest path algorithm where the steps between nodes are minimized in $\mathrm {d}C/\mathrm {d}z$. (c) The parameters the shortest path in (b) correspond to along the length of the device. (d) The case that was used to minimize the path in $\mathrm {d}C/\mathrm {d}z$ at each node along the shortest path as defined in Section 2.2.

Download Full Size | PDF

Using Section 2.1, we set up adiabatic loss conditions for both modes of interest: $v_1=\mathrm {TE}_0$ and $v_2=\mathrm {TE}_1$. Section 2.2 allows us to add an extra dimension, $C$, to our design space to obtain a more optimized structure without performing additional eigenmode computations. Finally, Section 2.3 lets us introduce the parameter $F_\mathrm {L}$ that will allow us to further reduce the crosstalk at the cost of increased insertion loss for a given device length. We discretize Eqs. (5)–(10) and use cases C.1, C.2 and C.3 to determine $\mathbf {p}(z)$ and $C(z)$ via a shortest path algorithm similar to that described in Section 4.1 of [18]. We conservatively chose $F_\mathrm {L}=0.2$ for our design as it appeared to slightly improve the simulated crosstalk while negligibly impacting the simulated transmission.

The $p_w$$p_g$ space (where each node is minimized in $\textrm{d} C / \textrm{d} z$) is shown in Fig. 1(b), with the shortest path to traverse our two target endpoints shown in green. We note that to adhere to the foundry’s manufacturing requirements, the domain of the $p_w$$p_g$ space is restricted to $p_w \geq {100 \; \textrm{nm}}$ and $p_g \geq {150\; \textrm{nm}}$. The optimized parameters $p_w(z/L)$, $p_g(z/L)$ and $C(z/L)$ where $L$ is the length of the modemux are shown in Fig. 1(c) along with a spline fit represented by a lighter colored line. Figure 1(c) shows the $\textrm{d} C / \textrm{d} z$ case that was used at each position along the design: towards the start of the device, the TE$_1$ loss limit governs $\textrm{d} C / \textrm{d} z$ and towards the end of the device, equality between the TE$_0$ and TE$_1$ loss limits governs $\textrm{d} C / \textrm{d} z$. This is expected as TE$_1$ is more sensitive to bending than TE$_0$ in a multimode waveguide. Towards the end of the device, TE$_0$ and TE$_1$ become increasingly similar and have similar bending properties.

The results shown in Fig. 2 show the simulated TE$_1$ transmission and the simulated crosstalk of various modemux designs for various $F_\mathrm {L}$, as well as the case without center shifting. Figure 2(a) shows how $L_0$—a length that characterizes the adiabatic loss limit [18]—varies in $F_\mathrm {L}$, reveals decreasing $F_\mathrm {L}$ results in a longer design as expected. Figures 2(b) and 2(c) confirm that a smaller $F_\mathrm {L}$ has improved crosstalk but requires a longer device to maintain a high TE$_1$ transmission. For this particular design, the optimization in $\textrm{d} C / \textrm{d} z$ was critical to achieving a low crosstalk.

 figure: Fig. 2.

Fig. 2. Eigenmode expansion simulations of various modemuxes tuned using $F_\mathrm {XT}$. (a) Sweep of $F_\mathrm {L}$ showing how $L_0$ increases as $F_\mathrm {L} \to 0$. (b) Crosstalk between the TE$_0$ and TE$_1$ modes through the modemuxes. (c) Transmission of the TE$_1$ mode through the modemuxes.

Download Full Size | PDF

3.2 Measurements

Measurements were conducted across four 300 mm wafers each containing 64 reticles with conservatively sized modemuxes of lengths 190 µm, 280 µm and 400 µm. The 190 µm and 280 µm lengths approximately correspond to nulls in the length–crosstalk simulations shown in Fig. 2(b), and the 400 µm length was chosen as an extreme case against which the shorter length devices could be benchmarked. A fiber array angled at 8° was used to couple light into and out of each reticle via grating couplers spaced 127 µm. The grating couplers exhibited minimum reticle-to-reticle variation at 1310 nm, with variation increasing significantly closer to the edges of the O-band. Figure 3 presents transmission and crosstalk measurements for the 280 µm long modemux, though no significant performance differences were seen between the various modemux lengths as expected from simulation. A face-to-face modemux configuration was used to measure the transmission of a single modemux by calculating $(C+D)/2$ and the results are shown in Fig. 3(a). Figure 3(b) shows the crosstalk measurements made using just a single modemux calculated via $F/(F+G)$. The solid lines show the median crosstalk across all reticles in a particular wafer at each measured wavelength, while the boxplots show the spread in transmission for the center wavelengths of the four CWDM bands.

 figure: Fig. 3.

Fig. 3. Measurement macro description and measurement results for transmission and crosstalk across four wafers each containing 64 reticles. (a) Boxplots for the transmission for the TE$_0$ and TE$_1$ mode at 1310 nm. (b) Crosstalk measurements showing the TE$_0$ to TE$_1$ crosstalk with boxplots showing the the variation across the four CWDM bands. The solid lines show the median crosstalk in wavelength across the 64 reticles. The cartoons above (a) and (b) show the insertion loss and crosstalk macros respectively with the arrows to the left left indicating optical inputs (from a swept laser) and the arrows to the right indicating optical outputs (to off-chip photodetectors).

Download Full Size | PDF

Nearly every reticle measured had a crosstalk $<\! {-40}\; \textrm{db}$ across the O-band with the mean crosstalk being approximately −45 dB. Accurate measurement of modemux transmission was not a priority, but was measured at 1310 nm (to minimize reticle-to-reticle variation) to be $< {-0.1}\; \textrm{db}$ with the limitations being the test macro and measurement fiber alignment accuracy. The characterization results show that the modemux operates well with extremely low crosstalk across the whole O-band and does not have significant insertion loss for either the TE$_0$ or the TE$_1$ modes, verifying the theory outlined in Section 2. There is good agreement between the simulated crosstalk at 1310 nm shown in Fig. 2(b) (−46 dB at $L= {280}\; \mathrm{\mu m}$) and the median crosstalk measurements at 1310 nm that lie between −45 dB and −50 dB.

4. Conclusion

We extend MODALL theory to enable adiabatic device optimization of many modes, optimization against a particular dimension that does not require extra eigenmode computation and optimization of crosstalk in two-mode devices. The theory was applied to a TE$_0$/TE$_1$ adiabatic mode multiplexer designed using all the theoretical extensions introduced, mass produced at a photonic foundry and characterized across multiple 300 mm wafers. The adiabatic mode multiplexer demonstrated robust performance with 100% of reticles having $< {-38}\; \textrm{db}$ crosstalk at 1310 nm and median crosstalk at 1310 nm matching well with the simulation performed at 1310 nm. The proposed MODALL extensions pave the way towards highly optimized adiabatic photonic devices including power splitters, mode multiplexers and polarization rotators. Future work will focus on applying the theory presented in this work to other adiabatic devices critical to photonic transceivers.

Acknowledgment

The authors wish to acknowledge the support and feedback provided by Prakash Gothoskar and Mark Webster of Cisco Systems, Inc.

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. S. Y. Siew, B. Li, F. Gao, et al., “Review of silicon photonics technology and platform development,” J. Lightwave Technol. 39(13), 4374–4389 (2021). [CrossRef]  

2. N. Margalit, C. Xiang, S. M. Bowers, et al., “Perspective on the future of silicon photonics and electronics,” Appl. Phys. Lett. 118(22), 220501 (2021). [CrossRef]  

3. Y. Su, Y. Zhang, C. Qiu, et al., “Silicon photonic platform for passive waveguide devices: Materials, fabrication, and applications,” Adv. Mater. Technol. 5(8), 1901153 (2020). [CrossRef]  

4. X. Tu, C. Song, T. Huang, et al., “State of the art and perspectives on silicon photonic switches,” Micromachines 10(1), 51 (2019). [CrossRef]  

5. C. Doerr, “Silicon photonic integration in telecommunications,” Front. Phys. 3, 5 (2015). [CrossRef]  

6. J. E Bowers, T. Komljenovic, M. Davenport, et al., “Recent advances in silicon photonic integrated circuits,” in Next-Generation Optical Communication: Components, Sub-Systems, and Systems V, Vol. 9774, G. LiX. Zhou eds., International Society for Optics and Photonics (SPIE, 2016), p. 977402.

7. A. W. Snyder, “Mode propagation in a nonuniform cylindrical medium (correspondence),” IEEE Trans. Microwave Theory Tech. 19(4), 402–403 (1971). [CrossRef]  

8. A. W. Snyder and J. Love, Optical waveguide theory (Springer Science Business Media, 2012).

9. A. Milton and W. Burns, “Mode coupling in optical waveguide horns,” IEEE J. Quantum Electron. 13(10), 828–835 (1977). [CrossRef]  

10. Y. Wu, F. Liang, H. Chung, et al., “Adiabaticity engineering in optical waveguides,” Opt. Express 28(20), 30117–30129 (2020). [CrossRef]  

11. J. D. Love, “Application of a low-loss criterion to optical waveguides and devices,” IEE Proc.-J: Optoelectron. 136(4), 225–228 (1989). [CrossRef]  

12. J. D. Love, W. M. Henry, W. J. Stewart, et al., “Tapered single-mode fibres and devices. i. adiabaticity criteria,” IEE Proc.-J: Optoelectron. 138(5), 343–354 (1991). [CrossRef]  

13. R. J. Black, S. Lacroix, F. Gonthier, et al., “Tapered single-mode fibres and devices. ii. experimental and theoretical quantification,” IEE Proc.-J: Optoelectron. 138(5), 355–364 (1991). [CrossRef]  

14. S. Tseng, R. Wen, Y. Chiu, et al., “Short and robust directional couplers designed by shortcuts to adiabaticity,” Opt. Express 22(16), 18849–18859 (2014). [CrossRef]  

15. S. Longhi, G. Della Valle, M. Ornigotti, et al., “Coherent tunneling by adiabatic passage in an optical waveguide system,” Phys. Rev. B 76(20), 201101 (2007). [CrossRef]  

16. G. Queraltó, V. Ahufinger, and J. Mompart, “Integrated photonic devices based on adiabatic transitions between supersymmetric structures,” Opt. Express 26(26), 33797–33806 (2018). [CrossRef]  

17. J. M. F. Cabanillas and M. A. Popovic, “Fast adiabatic mode evolution based on geometry-induced suppression of nearest-mode crosstalk,” in 2018 Conference on Lasers and Electro-Optics, (2018).

18. D. F. Siriani and J. Tambasco, “Adiabatic guided wave optics - a toolbox of generalized design and optimization methods,” Opt. Express 29(3), 3243–3257 (2021). [CrossRef]  

19. D. F. Siriani and J. Tambasco, “Crosstalk suppression in adiabatic 2x2 couplers via device perturbative reshaping,” in 2021 IEEE 17th International Conference on Group IV Photonics (GFP), (2021), pp. 1–2.

20. J. Tambasco and D. F. Siriani, “Optimization of tapers and interlayer transitions via adiabatic loss limiting,” in 2021 IEEE 17th International Conference on Group IV Photonics (GFP), (2021), pp. 1–2.

21. J. Tambasco, “Bilayer photonic 3db y-splitter,” (U.S. Patent Appl. US2023/0034508 A1, Feb. 2023).

22. J. Tambasco, “Bilayer silicon nitride polarization splitter and rotator,” (U.S. Patent Appl. US2023/0266531 A1, Aug. 2023).

23. J. Tambasco, “Simultaneous polarization splitter rotator,” (U.S. Patent No. US 11,698,491 B2, Jul. 2023).

24. J. Tambasco, “Polarization beam splitter,” (U.S. Patent No. US 11,714,235, Aug. 2023).

25. J. S. Arora, Introduction to Optimum Design (Academic Press, Boston, 2017) 4th ed.

26. K. Chung, T. Shih, J. Lee, et al., “Ultra-low crosstalk and fabrication-tolerant silicon-nitride o-band (de)multiplexer using bragg grating-assisted contra-directional coupler,” IEEE Photonics J. 15, 1–9 (2023). [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 (3)

Fig. 1.
Fig. 1. (a) Cartoon of the design structure with the critical parameters labeled. (b) The shortest path through the optimization space calculated using a discrete shortest path algorithm where the steps between nodes are minimized in $\mathrm {d}C/\mathrm {d}z$. (c) The parameters the shortest path in (b) correspond to along the length of the device. (d) The case that was used to minimize the path in $\mathrm {d}C/\mathrm {d}z$ at each node along the shortest path as defined in Section 2.2.
Fig. 2.
Fig. 2. Eigenmode expansion simulations of various modemuxes tuned using $F_\mathrm {XT}$. (a) Sweep of $F_\mathrm {L}$ showing how $L_0$ increases as $F_\mathrm {L} \to 0$. (b) Crosstalk between the TE$_0$ and TE$_1$ modes through the modemuxes. (c) Transmission of the TE$_1$ mode through the modemuxes.
Fig. 3.
Fig. 3. Measurement macro description and measurement results for transmission and crosstalk across four wafers each containing 64 reticles. (a) Boxplots for the transmission for the TE$_0$ and TE$_1$ mode at 1310 nm. (b) Crosstalk measurements showing the TE$_0$ to TE$_1$ crosstalk with boxplots showing the the variation across the four CWDM bands. The solid lines show the median crosstalk in wavelength across the 64 reticles. The cartoons above (a) and (b) show the insertion loss and crosstalk macros respectively with the arrows to the left left indicating optical inputs (from a swept laser) and the arrows to the right indicating optical outputs (to off-chip photodetectors).

Equations (15)

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

ϵ μ 2 = 4 m μ ( κ m μ ( p t ) d p t dz Δ β μ m ( p t ) ) 2 ,
κ m μ = ( ε 0 μ 0 ) 2 k 0 4 1 Δ β μ m ( p t ) A e ^ m e ^ μ p t ε r d A ,
ϵ v i 2 = Δ 4 m v i ( κ m v i ( p t ) d p t d z Δ β v i m ( p t ) ) 2 η v i 2 ( z ) ,
ϵ v 2 = max i ϵ v i 2 η v i 2 ( z )     i ,
ϵ v i 2 = 4 m v i ( κ m v i ( p ) ( p ) d p dz + κ m v i ( C ) ( p ) d C dz Δ β v i m ( p ) ) 2 ,
κ m v i ( p ) = ( ε 0 μ 0 ) 2 k 0 4 1 Δ β v i m ( p ) A e ^ m e ^ v i p ε r d A ,
κ m v i ( C ) = ( ε 0 μ 0 ) 2 k 0 4 1 Δ β v i m ( p ) A e ^ m e ^ v i ε r C d A .
L ( C , λ 1 , λ 2 ) = ϵ v 2 + λ 1 ( ϵ v 1 2 ϵ v 2 + s 1 2 ) + λ 2 ( ϵ v 2 2 ϵ v 2 + s 2 2 )
C . 1 Choose d C / d z = d C v 1 ( min ) / d z and ϵ v 2 = ϵ v 1 2 , C . 2 Choose d C / d z = d C v 2 ( min ) / d z and ϵ v 2 = ϵ v 1 2 , and C . 3 Choose ϵ v 2 = ϵ v 1 2 = ϵ v 2 2 and d C / d z { d C q 1 / d z , d C q 2 / d z }
d C v i ( m i n ) d z = m v i [ κ m v i ( C ) Δ β v i m 2 κ m v i ( p ) d p d z ] m v i ( κ m v i ( C ) ) 2 Δ β v i m 2
0 = ϵ v 1 2 ϵ v 2 2
m v 1 f m v 1 F X T f v 2 v 1 + F L m { v 1 , v 2 } f m v 1       w h e r e       F X T , F L [ 0 , 1 ]
f m v 1   = Δ ( 2 κ m v 1 ( p t ) d p t d z Δ β v 1 m ( p t ) ) 2 ,
ϵ v i 2 = 4 Δ z 2 m v i ( 1 Δ β v i m ( p ) ( κ m v i ( p ) ( p ) Δ p + κ m v i ( C ) ( p ) Δ C ) ) 2 ,
Δ C v i ( m i n ) = m v i [ κ m v i ( C ) Δ β v i m 2 κ m v i ( p ) Δ p ] m v i ( κ m v i ( C ) ) 2 Δ β v i m 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.