Introduction
Topology optimization (TO) is a computational design method that aims to find the optimal distribution of material to improve part performance under the governing physical equilibrium equations [1]. It originates in structural mechanics [2] and has been applied to structures of continuum domains where designs are represented as distributed functions and to discrete truss structures where designs are of finite dimensions and represented via discrete parameters. In this article, we focus on the design problems of distributed parameter systems, a.k.a. continuum problems. Continuum TO has received growing attention in a variety of fields, including multiscale structures [3], fluid mechanics [4], [5], [6], [7], electromagnetics [8], [9], photonics [10], [11], quantum devices [12], and coupled multiphysics problems [13], [14], [15], [16]. Topologically optimized designs often exhibit complex free-form shapes, and additive manufacturing can produce parts of complex shapes. The recent wide adoption of additive manufacturing technologies has led to further development of TO for a host of applications [17], including imposing additive manufacturability constraints during the optimization process [18], [19].
As TO has been expanded from structural mechanics to general multiphysics, it is accompanied by quickly growing complexity and a vast searching space, resulting in grand computational challenges [12], [17], [20]. Quantum computing (QC) is emerging as a new computing paradigm that could be superior to classical computing for a variety of problems, optimization being one of them. In literature, only a few attempts have been made to apply quantum algorithms for solving TO problems. Maruo et al. [21] concerned the TO of electromagnetic devices. For that, a linear optimization problem was formulated, which can be directly cast into a quadratic unconstrained binary optimization (QUBO) formulation through the Biot–Savart's law. The formulated QUBO problem was then solved through simulated annealing. Their results have indicated that the proposed QUBO formulation can significantly reduce the number of iterations required for the optimization process compared with the commonly used classical method, the normalized Gaussian network [22]. Thereby, the overall computational cost can be significantly reduced, because much fewer forward evaluations of the objective functions are invoked, which usually dominate the computational cost. However, the resultant QUBO formulation was not implemented in quantum annealers. Thus, the additional overhead due to quantum embedding and the limitations of quantum hardware currently accessible were not addressed in the analysis of computational cost. Also, the benefit of quantum annealing to converge to the optimal or ground state with larger probability than simulated annealing was not utilized. The work by Sato et al. [23] reported the TO for a simple, discrete truss structure with three edges. The objective functions of all possible configurations composed of three edges were simultaneously evaluated through quantum entanglement. The optimal configuration with the minimum objective function value was determined by applying the variational quantum algorithms (VQAs) that are suitable for gate-based quantum processors. The VQAs and forward evaluations of the objective function via finite element analysis (FEA) were integrated together when implemented on a quantum computer, eliminating the need of amplitude estimation for the solution of FEA. In those two works, the proposed quantum algorithms can only be applied to linear TO problems, where the field variables such as magnetic flux density and temperature linearly depend on the design variables. Given that, the TO can be greatly simplified when the objective function is just a linear [23] or quadratic [22] function of the field variables. In addition, the volume or material usage constraint is not strictly imposed by any equality or inequality constraint; as a result, one cannot set a target optimal volume or material usage for the optimization, which is not common in practical TO applications. Further, discrete truss structures only represent a small portion of applications of TO. More generally, TO considers structures of continuum domains where designs are represented as distributed functions.
In the present work, we concern how to apply quantum algorithms to continuum, nonlinear TO, beyond the discrete truss structures or the linear TO problems tackled in literature [22], [23]. In addition, the volume constraint is strictly imposed by an equality or inequality constraint, such that any target optimal volume can be practically achieved. To determine the optimal distribution of materials, continuum TO essentially seeks the numerical solution of a partial differential equation (PDE)-constrained optimization problem. To solve it, the solution domain is first discretized with discrete nodes or elements. The distributed design problem hence involves a large number of binary (0/1) optimization variables, typically one variable per discrete node or element. As the field variables nonlinearly depend on the design variables, it results in a large-scale, mixed integer nonlinear programming (MINLP) problem, subject to equality and/or inequality constraints. This type of problem is intrinsically NP-hard [24], namely, unable to be solved in polynomial time, and hence is challenging to be handled using classical approaches [25], [26], [27], [28].
There are two main implementations of the noisy intermediate-scale quantum (NISQ) devices: quantum annealers and gate-based quantum processors. Quantum annnealers embrace quantum annealing techniques, and the gate-based quantum circuits provide universal quantum computation capabilities. Due to the noisy nature of quantum gates and the high overhead cost of noise reduction and error mitigation, the quantum circuit that can be reliably executed must be limited to short. Thus, hybrid quantum-classical algorithms such as VAQs must be developed for solving optimization on gate-based quantum circuits, where only a short parameterized circuit is executed, and classical optimizers are leveraged to train the quantum circuit by optimizing the expectation value over the circuit's parameters [29]. However, in light of the noise-induced barren plateau [30], [31], [32], the training process calls for an exponentially scaled computational cost [30], and the classical optimization problems corresponding to the training of parameters are NP-hard [33]. Therefore, in the present work, we exploit quantum annealing for solving the TO problem of our interest, which has been demonstrated for its advantages and speedups in solving a variety of combinatorial optimization problems [34], [35], [36], [37], [38]. For an n-qubit system, it works in the binary discrete space, with the operator defined as a Hamiltonian that can correspond to the objective function of an optimization problem [39], [40]. Quantum annealing can effectively escape local minima by tunneling through the barriers when exploring the objective function's landscape and seeking the global minimum [39], [40]. Thus, it can efficiently solve large-scale nonlinear optimization problems over binary discrete spaces. The final state of the n-qubit system at the end of annealing could be the ground state of the Hamiltonian, with each qubit a classical object standing for the solution to the binary optimization problem.
The currently accessible quantum computers that are designed to implement quantum annealing are the D-Wave systems [41]. To embed problems on the D-Wave quantum annealer, they need to be formulated in the form of QUBO. However, the optimization problem corresponding to the TO of our interest cannot be directly cast into QUBO formulations. Therefore, we first decompose the original problem into a sequence of mixed integer linear programs (MILPs) via the generalized Benders' decomposition (GBD). Next, in each MILP the continuous variables are represented by a series of auxiliary binary variables, such that each MILP can be mapped onto a binary optimization problem and then through the penalty method transformed into the QUBO formulation. Finally, from the solution of quantum annealing, we obtain the solution of each MILP, and in turn the ultimate optimal material layout after the sequence of MILPs are completely solved. If sufficient, fully connected quantum qubits are available [42], quantum advantages in computational efficiency can be straightforwardly obtained and demonstrated, owing to the efficiency of quantum annealing for solving QUBO. However, currently accessible quantum hardware only offers a small number of qubits with sparse connectivity. To solve practical TO problems in current and near-term quantum annealing devices, we further develop a splitting approach, by which a reduced problem that needs to be solved on a quantum annealer is separated from other subproblems that can be efficiently handled by classical computers. The resultant, greatly reduced QUBO problem is then implemented on a D-Wave Advantage quantum processing unit (QPU). Both the solution quality and computational cost (including the cost for embedding the problem onto the quantum computer, the time of quantum annealing, and the number of optimization iterations) are analyzed and assessed systematically in a series of problems of variable sizes.
Preliminary
A. Continuum Topology Optimization Problem
Continuum TO optimizes the material layout
\begin{equation*}
\begin{aligned} \min _{\rho } & \quad \int _{\Omega } F[\mathbf {u}(\rho (\mathbf {x}))] \mathrm{d} \Omega \\
\text{s.t.} & \quad G_{0}(\rho (\mathbf {x})) = \frac{\int _{\Omega } \rho (\mathbf {x}) \mathrm{d} \Omega }{\int _{\Omega } \mathrm{d} \Omega } - V_{T} = 0 \\
& \quad \mathcal {L}(\mathbf {u}(\rho (\mathbf {x}))) + \mathbf {b} = 0 & & \forall \mathbf {x} \in \Omega \\
& \quad \mathbf {u}(\rho (\mathbf {x})) = \mathbf {u}_\Gamma (\mathbf {x}) & & \forall \mathbf {x} \in \Gamma _{D} \\
& \quad \mathbf {n} \cdot \nabla \mathbf {u}(\rho (\mathbf {x})) = \mathbf {h}_\Gamma (\mathbf {x}) & & \forall \mathbf {x} \in \Gamma _{N} \\
& \quad G_{j}(\mathbf {u}(\rho (\mathbf {x})), \rho (\mathbf {x})) = 0, \quad j = 1, \dots m \\
& \quad H_{k}(\mathbf {u}(\rho (\mathbf {x})), \rho (\mathbf {x})) \leq 0, \quad k = 1,\ldots, n \\
\end{aligned} \tag{1}
\end{equation*}
where
To solve the continuum TO problem as in (1), the design domain
\begin{equation*}
\begin{aligned} \min _{\mathbf {u}, \boldsymbol{\rho }} & \quad \mathbf {f}^\intercal \mathbf {u} \\
\text{s.t.} & \quad \mathbf {K}(\boldsymbol{\rho }) \mathbf {u} = \mathbf {f} \\
& \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V_{T} \\
& \quad \mathbf {u} \in \mathbb {R}^{n_{u}}, \boldsymbol{\rho } \in \lbrace 0, 1 \rbrace ^{n_{\rho }} \end{aligned} \tag{2}
\end{equation*}
\begin{equation*}
\mathbf {K}(\boldsymbol{\rho }) = \mathbf {K}_{0} + \sum _{i=1}^{n_{\rho }} \rho _{i} \mathbf {K}_{i}(E, \nu) \; \tag{3}
\end{equation*}
\begin{equation*}
\mathbf {K}_{0} = \varepsilon \sum _{i = 1}^{n_{\rho }} \mathbf {K}_{i} \; \tag{4}
\end{equation*}
The challenge for solving the optimization problem defined in (2) lies in the fact that while
In the first category, the most widely used method is the solid isotropic material with penalization (SIMP) method [25]. Its strategy is to relax the binary variable into a continuous variable such that the MINLP problem can be converted into a continuous optimization problem. Specifically, the SIMP method approximates the binary design variable with a high-order polynomial function of a continuous variable that smooths out the discontinuity of the binary variable [25]. Hence, the density
The second category of methods separate the binary and continuous variables into different subproblems, by iteratively determining one of them with the other fixed in a subproblem. The representative methods include the discrete variable topology optimization via canonical relaxation algorithm (DVTOCRA) [26], the topology optimization of binary structures (TOBS) method [27], and the GBD method [28]. The DVTOCRA method employs a sequential linear/quadratic approximation to separate the binary and continuous variables into different subproblems [26]. The subproblem associated with the binary variable is a constrained, quadratic integer programming problem. Due to its NP-hardness, the canonical relaxation algorithm is applied, by which new continuous variables are introduced to replace the binary variable through an approximation function (different from that used in the SIMP method), and the binary optimization programming problem is transformed into a convex, continuous optimization problem in terms of the newly introduced continuous variables. Similarly to the SIMP method, the resultant continuous optimization problems are not suitable for quantum annealing to solve. The TOBS method utilizes sequential linear programming to separate the binary and continuous variables into different subproblems [27]. The resultant subproblem involving the binary variable is a linear integer programming problem. With the uniform mesh discretization and the exclusion of nonvolumetric constraints [i.e.,
The GBD method follows a different route. It decomposes the original problem into a sequence of MILPs. In contrast to the other methods that have no guarantee of convergence, the GBD method provides a deterministic optimality criterion to warrant convergence within a finite number of iterations [24]. It is also anticipated to converge faster than the SIMP or TOBS method, because the GBD formulation permits to use all the material layouts generated from previous iterations in each new iteration, while other methods like the SIMP or TOBS can only take into account the material layout yield from the last iteration. The GBD method was first introduced to TO by Muñoz et al. [28] for solving discrete TO problems. In the present article, we extend it to address a continuum TO problem, as discussed in Section II-B. In the subproblem related to the binary design variable after decomposition, additional constraints (associated with the PDEs), other than the volumetric constraint, are introduced and can bring NP-hardness to the optimization problem. Such a subproblem is well suited to be accelerated by quantum annealing, and hence we focus on the GBD method in the present article to investigate and demonstrate how QC, particularly via quantum annealing, can be leveraged for solving a continuum, nonlinear TO problem as stated in (2).
B. Generalized Benders' Decomposition
The field variable
At the
\begin{equation*}
\begin{aligned} \min _{\mathbf {u}} & \quad \mathbf {f}^\intercal \mathbf {u} \\
\text{s.t.} & \quad \mathbf {K}(\boldsymbol{\rho }^{k}) \mathbf {u} = \mathbf {f} \\
& \quad \mathbf {u} \in \mathbb {R}^{n_{u}}. \end{aligned} \tag{5}
\end{equation*}
\begin{equation*}
\mathbf {u}^{k} = \mathbf {K}^{-1}(\boldsymbol{\rho }^{k}) \mathbf {f} \;. \tag{6}
\end{equation*}
Next, we derive the subproblem with respect to
\begin{equation*}
\begin{aligned} \min _{\mathbf {u}, \boldsymbol{\rho }} & \quad \mathbf {f}^\intercal \mathbf {u} + \bm {\lambda }^\intercal \mathbf {K}(\bm {\rho }) \mathbf {u} \\
\text{s.t.} & \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \mathbf {u} \in \mathbb {R}^{n_{u}}, \boldsymbol{\rho } \in \lbrace 0, 1 \rbrace ^{n_{\rho }}. \end{aligned} \tag{7}
\end{equation*}
\begin{equation*}
\bm {\lambda }^\intercal \mathbf {K} (\bm {\rho }) = -\mathbf {f}^\intercal \;. \tag{8}
\end{equation*}
\begin{equation*}
\begin{aligned} \min _{\bm {\rho }, \mathbf {u}, \eta } & \quad \eta \\
\text{s.t.} & \quad \mathbf {f}^\intercal \mathbf {u} + \bm {\lambda }^\intercal \mathbf {K}(\bm {\rho }) \mathbf {u} \leq \eta \\
& \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \mathbf {u} \in \mathbb {R}^{n_{u}}, \boldsymbol{\rho } \in \lbrace 0, 1 \rbrace ^{n_{\rho }}. \end{aligned} \tag{9}
\end{equation*}
\begin{equation*}
\mathbf {f}^\intercal \mathbf {u}^{k} + \sum _{i = 1}^{n_{\rho }} {(\bm {\lambda }^{k})}^\intercal \mathbf {K}_{i} \mathbf {u}^{k} (\rho _{i} - \rho _{i}^{k}) \leq \mathbf {f}^\intercal \mathbf {u} \leq \eta. \tag{10}
\end{equation*}
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }, \eta } & \quad \eta \\
\text{s.t.} & \quad \mathbf {f}^\intercal \mathbf {u}^{j} + \sum _{i=1}^{n_{\rho }} {(\bm {\lambda }^{j})}^\intercal \mathbf {K}_{i} \mathbf {u}^{j} (\rho _{i} - \rho ^{j}_{i}) \leq \eta \\
& \qquad \text{j = 1},\ldots\text{, k} \\
& \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{n_{\rho }}. \end{aligned} \tag{11}
\end{equation*}
Since the continuum TO is essentially a PDE-constrained optimization problem [as in (1)], the solution quality also depends on the accuracy of the numerical approximation of the differential operator (e.g., gradient) in the PDE, which becomes especially challenging at the interface between solid (
Effect of filtering on eliminating the checkerboard artifact. The two subfigures report the resulting optimal material layouts with the same problem setup, using the same optimization procedure, and corresponding to the same region of the solution domain. (a) Optimal material layout yield without filtering exhibits a checkerboard pattern. (b) After applying filtering, the checkerboard artifact is effectively eliminated. (a) Material layout with “checkerboard” artifact. (b) Material layout after filtering.
To remedy for that, filtering is required for the sensitivity, following the literature [27], [45], as
\begin{equation*}
\widetilde{w}_{i}^{j} = \left\lbrace \begin{array}{l} \frac{\sum _{l \in \mathcal {N}_{i}^{r}} h_{i, l}^{r} (\mathbf {u}^{j})^\intercal \mathbf {K}_{l} \mathbf {u}^{j}}{\sum _{l \in \mathcal {N}_{i}^{r}} h_{i, l}^{r}}, \quad \rho _{i}^{j} = 1 \\
\varepsilon \frac{\sum _{l \in \mathcal {N}_{i}^{r}} h_{i, l}^{r} (\mathbf {u}^{j})^\intercal \mathbf {K}_{l} \mathbf {u}^{j}}{\sum _{l \in \mathcal {N}_{i}^{r}} h_{i, l}^{r}}, \quad \rho _{i}^{j} = 0 \end{array} \right. \tag{12}
\end{equation*}
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }, \eta } & \quad \eta \\
\text{s.t.} & \quad \mathbf {f}^\intercal \mathbf {u}^{j} - \sum _{i=1}^{n_{\rho }} \widetilde{w}_{i}^{j} (\rho _{i} - \rho ^{j}_{i}) \leq \eta \\
& \qquad \text{j = 1},\ldots\text{, k} \\
& \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{n_{\rho }}. \end{aligned} \tag{13}
\end{equation*}
As the iterations proceed, the number of inequality constraints (
\begin{equation*}
\mathcal {P}(k) = \lbrace j | \mathbf {f}^\intercal \mathbf {u}^{j} \leq \mathbf {f}^\intercal \mathbf {u}^{k}\quad \ \forall j = 1, \dots k \rbrace \; \tag{14}
\end{equation*}
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }, \eta } & \quad \eta \\
\text{s.t.} & \quad \mathbf {f}^\intercal \mathbf {u}^{j} - \sum _{i=1}^{n_{\rho }} \widetilde{w}_{i}^{j} (\rho _{i} - \rho ^{j}_{i}) \leq \eta \quad \ \quad \forall j \in \mathcal {P}(k) \\
& \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{n_{\rho }}. \end{aligned} \tag{15}
\end{equation*}
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }} & \quad \mathbf {f}^\intercal \mathbf {u}^{j} - \sum _{i=1}^{n_{\rho }} \widetilde{w}_{i}^{j} (\rho _{i} - \rho ^{j}_{i}), \quad j \in \mathcal {P}(k) \\
\text{s.t.} & \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{n_{\rho }} \end{aligned} \tag{16}
\end{equation*}
In summary, for any given volume constraint
Algorithm 1: ToGbdSub(V , \bm {\rho }^{1} ).
Input: Volume fraction
Output: Optimal material layout
Utility: Find an optimal material layout with a given volume fraction
for
Initialize the upper bound
Employ the linear system solver to obtain
if
end if
Generate the Pareto optimal cuts
if
Solve the master problem (16) for the minimizer
else
Solve the master problem (15) for the minimizer
end if
if
break
end if
end for
Return:
In the following, we explain how to obtain
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }} & \quad \mathbf {f}^\intercal \mathbf {u}^{0} - \sum _{i=1}^{n_{\rho }} \widetilde{w}_{i}^{0} (\rho _{i} - \rho ^{0}_{i}) \\
\text{s.t.} & \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{n_{\rho }} \end{aligned} \tag{17}
\end{equation*}
Algorithm 2: ToGbd(V_{T} , \Delta V ).
Input: Target volume fraction
Output: Optimal material layout
Initialize the design variables as
for
Employ the linear system solver to obtain
Solve the problem (17) to obtain
Prepare for the next iteration with
end for
Return:
Quantum Algorithm
A. From Topology Optimization to QUBO
As discussed above, the original MINLP problem, as in (2), is relaxed into a series of MILP problems, as in (15). In this section, we establish the conversion of MILP into QUBO, such that the master problem (15) can be solved on quantum annealers.
First, a slack variable
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }, \eta, \bm {\alpha }} & \quad \eta \\
\text{s.t.} & \quad \mathbf {f}^\intercal \mathbf {u}^{j} - \sum _{i=1}^{n_{\rho }} \widetilde{w}_{i}^{j} (\rho _{i} - \rho ^{j}_{i}) + \alpha ^{j} = \eta \quad \ \quad \forall j \in \mathcal {P}(k) \\
& \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{n_{\rho }} \\
& \quad \alpha ^{j} \geq 0\quad \ \quad \forall j \in \mathcal {P}(k). \\
\end{aligned} \tag{18}
\end{equation*}
Next, the continuous variables
\begin{equation*}
\begin{aligned} \eta \approx & \widetilde{\eta }(\mathbf {e}) = \frac{U}{2^{n_{\eta }}} e_{0} + \sum _{i = 1}^{n_{\eta }} \frac{U}{2^{i}} e_{i} \\
\alpha ^{j} \approx & \widetilde{\alpha }^{j}(\mathbf {a}^{j}) = \frac{U}{2^{n_{\alpha ^{j}}}} a_{0}^{j} + \sum _{i = 1}^{n_{\alpha ^{j}}} \frac{U}{2^{i}} a^{j}_{i} \; \end{aligned}
\end{equation*}
\begin{equation*}
\begin{aligned} \mathbf {e} \in \lbrace 0, 1\rbrace ^{n_{\eta } + 1}, \quad \mathbf {a}^{j} \in \lbrace 0,1 \rbrace ^{n_{\alpha ^{j}} + 1} \;. \end{aligned}
\end{equation*}
\begin{align*}
& \min _{\boldsymbol{\rho }, \mathbf {e}, \mathbf {a^{j}}} \quad \widetilde{\eta }(\mathbf {e}) \\
&\rm{s.t.} \quad \mathbf {f}^\intercal \mathbf {u}^{j} - \sum _{i=1}^{n_{\rho }} \widetilde{w}_{i}^{j} (\rho _{i} - \rho ^{j}_{i}) + \widetilde{\alpha }^{j} (\mathbf {a}^{j}) \\
& \qquad\qquad \rm{= \widetilde{\eta }(\mathbf {e})\quad \forall j \in \mathcal {P}(k)} \\
& \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{n_{\rho }} \;.\tag{19}
\end{align*}
\begin{equation*}
\begin{aligned} \widetilde{\eta }(\mathbf {e}) & + \sum _{j \in \mathcal {P}(k)} A^{j} \left[ W^{j} - \left(\sum _{i = 1}^{n_{\rho }} \widetilde{w}_{i}^{j} \rho _{i} \right) + \widetilde{\alpha }^{j}(\mathbf {a}^{j}) - \widetilde{\eta }(\mathbf {e}) \right]^{2} \\
& + B \left(\sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} - V \right)^{2} \; \end{aligned} \tag{20}
\end{equation*}
\begin{equation*}
W^{j} = \mathbf {f}^\intercal \mathbf {u}^{j} + \sum _{i = 1}^{n_{\rho }} \widetilde{w}_{i}^{j} \rho _{i}^{j} \;.
\end{equation*}
Finally, we determine the penalty factors. Theoretically, the penalty factors in (20) should be gradually increased until the optimal solution converges. However, due to the limited machine precision of currently accessible quantum computers (e.g., up to
B. Splitting Approach for Problem Reduction
Solving the QUBO problem in (20) requires
To resolve this issue, we further develop a splitting approach in this section. We note that a binary programming problem like (16) or (17) can be efficiently solved by a classical optimizer, as discussed below. In fact, by taking the continuous relaxation of the design variable as
\begin{equation*}
P = \left\lbrace (\bar{\rho }_{1},\ldots, \bar{\rho }_{n_{\rho }}) \left| \sum _{i=1}^{n_{\rho }} \bar{\rho }_{i} = n_{\rho } V, \quad 0 \leq \bar{\rho }_{i} \leq 1\quad \ \forall i \right. \right\rbrace \;
\end{equation*}
By leveraging this observation, we propose the following procedure to split the problem (15) into two parts. The first part consists of several subproblems like (16), which will be solved on classical computers. The second part is a problem similar to (15) but with greatly reduced variables, which will be solved on a quantum computer. By taking only one of the
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }} & \quad \mathbf {f}^\intercal \mathbf {u}^{j} - \sum _{i=1}^{n_{\rho }} \widetilde{w}_{i}^{j} (\rho _{i} - \rho ^{j}_{i}) \ \quad \forall j \in \mathcal {P}(k) \\
\text{s.t.} & \quad \sum _{i=1}^{n_{\rho }} \frac{\rho _{i}}{n_{\rho }} = V \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{n_{\rho }} \;. \end{aligned} \tag{21}
\end{equation*}
\begin{equation*}
\widetilde{\rho }_{i}^{j_{1}} = \widetilde{\rho }_{i}^{j_{2}} \ \quad \forall i \in \mathcal {I} \subseteq \lbrace 1,\ldots, n_{\rho } \rbrace \ \quad \forall j_{1}, j_{2} \in \mathcal {P}(k) \;. \tag{22}
\end{equation*}
Thus, a reduced problem of (15) can be formed as
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }, \eta } & \quad \eta \\
\text{s.t.} & \quad R^{j} - \sum _{i \in \mathcal {I^{C}}} \widetilde{w}_{i}^{j} (\rho _{i} - \rho ^{j}_{i}) \leq \eta \ \quad \forall j \in \mathcal {P}(k) \\
& \quad \sum _{i \in \mathcal {I^{C}}} \frac{\rho _{i}}{n_{\rho }} = V - \sum _{i \in \mathcal {I}} \frac{\widetilde{\rho }_{i}}{n_{\rho }} \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{|\mathcal {I^{C}}|} \; \end{aligned} \tag{23}
\end{equation*}
\begin{equation*}
R^{j} = \mathbf {f}^\intercal \mathbf {u}^{j} - \sum _{i \in \mathcal {I}} \widetilde{w}_{i}^{j} (\widetilde{\rho }_{i}^{j} - \rho _{i}^{j})\;. \tag{24}
\end{equation*}
By introducing slack variables
\begin{equation*}
\begin{aligned} \min _{\boldsymbol{\rho }, \mathbf {e}, \mathbf {a}^{j}, \mathbf {b}} & \quad \widetilde{\eta }(\mathbf {e}) \\
\rm{s.t.} & \quad R^{j} - \sum _{i \in \mathcal {I^{C}}} \widetilde{w}_{i}^{j} (\rho _{i} - \rho ^{j}_{i}) + \widetilde{\alpha }^{j}(\mathbf {a}^{j}) \\
& \qquad \qquad \rm{ = \widetilde{\eta }(\mathbf {e}), \quad j \in \mathcal {P}(k)} \\
& \quad \sum _{i \in \mathcal {I^{C}}} \frac{\rho _{i}}{n_{\rho }} = \widetilde{V} \\
& \quad \boldsymbol{\rho } \in \lbrace 0, 1\rbrace ^{|\mathcal {I^{C}}|}, \mathbf {e} \in \lbrace 0, 1\rbrace ^{n_{\eta } + 1} \\
& \quad \mathbf {a}^{j} \in \lbrace 0,1 \rbrace ^{n_{\alpha ^{j}} + 1} \; \end{aligned} \tag{25}
\end{equation*}
\begin{equation*}
\begin{aligned} \widetilde{\eta }(\mathbf {e}) = & \frac{U}{2^{n_{\eta }}} e_{0} + \sum _{i = 1}^{n_{\eta }} \frac{U}{2^{i}} e_{i} \; \\
\widetilde{\alpha }^{j}(\mathbf {a}^{j}) = & \frac{U}{2^{n_{\alpha ^{j}}}} a_{0}^{j} + \sum _{i = 1}^{n_{\alpha ^{j}}} \frac{U}{2^{i}} a^{j}_{i} \; \\
\widetilde{V} = & V - \sum _{i \in \mathcal {A}} \frac{\widetilde{\rho }_{i}}{n_{\rho }} \;. \end{aligned}
\end{equation*}
\begin{equation*}
\begin{aligned} \widetilde{\eta }(\mathbf {e}) + & \sum _{j \in \mathcal {P}(k)} A^{j} \left[ R^{j} - \left(\sum _{i \in \mathcal {I^{C}}} \widetilde{w}_{i}^{j} \rho _{i} \right) + \widetilde{\alpha }_{j}(\mathbf {a}_{j}) - \widetilde{\eta }(\mathbf {e}) \right]^{2} \\
+ & B \left[ \sum _{i \in \mathcal {I^{C}}} \frac{\rho _{i}}{n_{\rho }} - \widetilde{V} \right]^{2} \;. \\
\end{aligned} \tag{26}
\end{equation*}
As
The proposed splitting approach is summarized in Algorithm 3.
Algorithm 3: ToGbdSubSplitting(V , \bm {\rho }^{1} ).
Input: Volume fraction
Output: Optimal material layout
Utility: Find an optimal material layout with a given volume fraction
for
Initialize the upper bound
Employ the linear solver to obtain
if
end if
Generate the Pareto optimal cuts
if
Solve the master problem (16) for the minimizer
else
Solve the subproblem (21) for the minimizer
Generate the index sets
Solve the reduced QUBO problem (26) for the minimizer
end if
if
break
end if
end for
Return:
Numerical Results
A. Solving a Toy Problem for Validation
To validate that the method discussed in Section III-A permits a quantum computer to effectively solve the MILP problem like (15), we consider the following mixed-integer programming problem with linear inequality and equality constraints:
\begin{equation*}
\begin{aligned} \min _{u, v, w, t} \quad & v + w + t + (u - 2)^{2} \\
\text{s. t. } \quad & v + \text{2}\,w + t + u \leq 3 \\
\quad & v + w + t \geq 1 \\
\quad & v + w = 1 \\
\quad & v, w, t \in \lbrace 0, 1 \rbrace, u \in \mathbb {R} \;. \end{aligned} \tag{27}
\end{equation*}
According to the proposed methodology, we first transform this problem into a binary programming problem, and then rewrite it in the formulation of QUBO. By noting that
\begin{equation*}
\widetilde{u}(\mathbf {e}) = \frac{3}{2^{n_{u}}} e_{0} + \sum _{i = 1}^{n_{u}} \frac{3}{2^{i}} e_{i} \;.
\end{equation*}
\begin{equation*}
\begin{aligned} \min _{\mathbf {e}, \mathbf {a}^{1}, \mathbf {a}^{2}, v, w, t} \quad & v + w + t + (\widetilde{u} - 2)^{2} \\
\text{s. t. } \quad & v + \text{2}\,w + t + \widetilde{u} + \widetilde{\alpha }^{1} = 3 \\
\quad & -v - w - t + \widetilde{\alpha }^{2} = -1 \\
\quad & v + w = 1 \\
\quad & v, w, t \in \lbrace 0, 1 \rbrace \\
\quad & \mathbf {e} \in \lbrace 0, 1\rbrace ^{n_{u} + 1} \\
\quad & \mathbf {a}^{1} \in \lbrace 0, 1\rbrace ^{n_{\alpha ^{1}} + 1}, \mathbf {a}^{2} \in \lbrace 0, 1\rbrace ^{2} \; \end{aligned} \tag{28}
\end{equation*}
\begin{equation*}
\begin{aligned} \widetilde{\alpha }^{1}(\mathbf {a}^{1}) = & \frac{3}{2^{n_{\alpha ^{1}}}} a^{1}_{0} + \sum _{i = 1}^{n_{\alpha ^{1}}} \frac{3}{2^{i}} a^{1}_{i} \\
\widetilde{\alpha }^{2}(\mathbf {a}^{2}) = & a_{0}^{2} + a_{1}^{2} \;. \end{aligned}
\end{equation*}
\begin{align*}
& v + w + t + \left[ \widetilde{u} - 2 \right]^{2} + A \left[ v + \text{2}\,w + t + \widetilde{u} - 3 + \widetilde{\alpha }^{1} \right]^{2} \\
& + A\left[ -v - w - t + 1 + \widetilde{\alpha }^{2} \right]^{2} + A\left[ v + w - 1 \right]^{2} \; \tag{29}
\end{align*}
where the penalty factor is set as
The problem (29) was directly submitted to the D-Wave Advantage QPU. The annealing time was set with 20
B. Solving the Minimal Compliance Problem
In this section, we concern a canonical TO problem, the minimum compliance, as formulated in (2). The minimum compliance problem denotes a kind of topology optimization problem for finding an optimal material layout exerted with static external loads with a given volume constraint when doing static analysis for structures. The optimal material layout can minimize the displacement under unit given external load—the compliance, and in turn maximize the load the structure can sustain. Therefore, the optimal solution can give out the most efficient way to utilize the given material to withstand external loads. The problem setup follows that in [25]. More specifically, we consider a rectangular material domain with the external loading of a unit point force
Schematic of the problem setup for the minimum compliance. The yellow region denotes the design domain discretized by a uniform quadrilateral mesh with the resolution of
To solve this TO problem, we followed the proposed methodology, consisting of the GBD, conversion of the MILP into QUBO, and the splitting approach. More specifically, Algorithm 2 was executed, but with ToGbdSub in Line 5 replaced by ToGbdSubSplitting defined in Algorithm 3 to integrate the proposed splitting approach. While the binary programming problems (16), (17), and (21) were solved by invoking the classical MILP solver Gurobi [49], the binary programming problem (25) was solved on the quantum annealer provided by D-Wave. From the validation test in Section IV-A and considering the accuracy of the current quantum devices and the required number of qubits, the parameters for the binary approximations were set as:
To implement the problem (25) on the quantum annealer provided by D-Wave, we examined two different ways. In the first way, we directly embedded the QUBO problem (26) formulated from (25) in the QPU, which is named as “GBD-splitting-direct.” In the second way, we solved the problem (25) by taking advantage of the hybrid solver provided by D-Wave, particularly the constrained quadratic model (CQM), which is hence referred to as “GBD-splitting-CQM.” We compare the results in terms of solution quality and wall time, as summarized in Table 2, where the discretization resolution is chosen as
In GBD-splitting-direct, each sampling is set with
With the discretization resolution of
To tackle this challenge, we pursued the second way of implementation, i.e., GBD-splitting-CQM, where the problem (25) was submitted, and the CQM, provided by D-Wave, was called for embedding and solving the problem. All default setups were used when employing the CQM hybrid solver in all the numerical tests. As indicated by the wall time
Based on the above comparison and findings, when we scaled up the minimum compliance problem with increasing numbers of design variables (with finer discretization resolutions), we employed the implementation of GBD-splitting-CQM, owing to its better performance in terms of the solution quality, wall time, and capacity to embed larger problems. In particular, we considered two different shapes of material domains:
Resultant optimal material layouts obtained by different methods, with the design space of (
Resultant optimal material layouts obtained by different methods, with the design space of (
To verify the QC-based solutions and to assess the benefits of QC for TO, we further compare with the state-of-the-art classical solvers, as discussed in the next section.
C. Comparison With Classical Solvers
The classical solvers considered include the SIMP method [25], the TOBS method [27], and the DVTOPCRA method [26]. For the SIMP method, we followed the implementation in [25] with the penalty factor
The results obtained by GBD-splitting-CQM and by the three classical methods are compared in Table 4 and Figs. 3 and 4, for different shapes of material domains with different discretization resolutions. In Table 4,
By comparison, the proposed solution strategy exhibits the following advantages. First, it turns out requiring the least number of iterations to reach convergence in almost all the test cases. This can be related to the multicuts generated by the GBD iterations and the guarantee of finite iterations by the convexity of the problem. Whereas, the other classical methods solve nonconvex subproblems and in turn have no guarantee for convergence within finite iterations. As the discretization resolution increases, the SIMP method with Heaviside projection cannot meet the convergence criterion within the set maximum number of iterations, and the other two methods, TOBS and DVTOPCRA, tend to require more iterations to reach convergence, as shown in Table 4. Second, the minimum values of the objective function found by GBD-splitting-CQM are generally lower than those by the other methods, especially SIMP and DVTOPCRA. Lastly, in terms of the wall time, GBD-splitting-CQM exhibits only a slow growth in computing time and hence a good scaling performance, in contrast to the other three methods. Thus, for solving larger-scale problems, e.g., with the discretization resolutions of
The resultant optimal material layouts obtained by different methods are shown and compared in Figs. 3 and 4, where two different shapes of the design space with the finest discretization resolution are presented. Unlike the nonconvergent SIMP designs with gray regions as shown in Figs. 3(b) and 4(b), the GBD-splitting-CQM designs have sharp contrast with clear 0/1 in density. We hence demonstrate the advantage of GBD-splitting-CQM in generating optimal material layouts with clear boundaries, compared against the designs by the method like SIMP that relaxes the binary design variable into a continuous variable. In addition, when compared with the designs yield from the TOBS method, the designs optimized by GBD-Splitting-CQM also exhibit better minimal length control. Furthermore, the layouts generated by the DVTOPCRA method display crooked internal structures, which can be unfavorable with respect to yield strength.
Conclusion
We have established a quantum annealing-based solution method for solving continuum, nonlinear TO problems. Its key ingredients include the GBD, conversion of MILP into QUBO, a splitting approach, as well as the implementation on the quantum computer through the hybrid solver, CQM, provided by D-Wave. Through solving a canonical TO problem, the minimum compliance, we have assessed its accuracy and efficiency. By comparing with the state-of-the-art classical methods commonly used in the field of TO, we have demonstrated the advantages of our proposed QC-based methodology, with respect to both solution quality and computational efficiency. Considering the run-time penalty for embedding due to the hardware limitations of current quantum annealing machines [50] and given their future improvements in both the number of qubits and the long-range couplers between qubits, the advantage of the proposed QC-based solution method in terms of computing time can be remarkable. We have also shown its computational efficiency for TO in the sense that the number of FEA runs has been reduced significantly when compared with the classical SIMP method. Future work would exploit this computational efficiency advantages for larger-scale TO problems where FEA cost dominates the optimization run.
This work also presents a new application paradigm for QC and expands the application horizon of QC to include TO, enabling more efficient designs of topology for broad applications.
With increasing complexity, the continuum TO problems can become more challenging. For example, involving the constraints like
The splitting approach developed in the present work has been shown effective for greatly reducing the size of the problem to be embedded in quantum computers, while keeping the solution quality for the optimal material layout. Other approaches, e.g., the multilevel hybrid framework introduced for generic combinatorial optimization [54] and the hierarchical mesh refinement approach [55], can be explored in the future and integrated into the GBD framework to further reduce the cost for solving the problems like (15) on quantum computers.