[go: up one dir, main page]

 
 

Simulation-Based Optimization: Methods and Applications in Engineering Design

A special issue of Algorithms (ISSN 1999-4893).

Deadline for manuscript submissions: closed (1 July 2022) | Viewed by 21947

Special Issue Editors


E-Mail Website
Guest Editor
CNR-INM, National Research Council-Institute of Marine Engineering, Via di Vallerano 139, 00128 Rome, Italy
Interests: simulation-based design optimization; machine learning; dimensionality reduction; surrogate-modelling; multi-fidelity methods; optimization algorithms; uncertainty quantification; application of computational fluid dynamics

E-Mail Website
Guest Editor
Institute of Marine Engineering, Italian National Research Council, 00128 Rome, Italy
Interests: simulation-based design; optimization; surrogate models; variable fidelity; fluid-structure interaction

Special Issue Information

Dear Colleagues,

In the engineering design context, the demand for efficient products is constantly increasing. These products must respond to an ever-increasing number of specific and complex requirements. Over the last thirty years, engineering design has radically transformed thanks to the exponential development of IT and digital resources. This has allowed the transformation of the classical approach of design, built, and test toward a more efficient simulation-based design optimization (SBDO) process, integrating numerical solvers, design modification methods, optimization algorithms, and also uncertainty quantification methods. The results obtained through the SBDO process are often a compromise between its efficiency (speed in achieving the optimum) and effectiveness (accurate simulations, requiring high-fidelity/computationally expensive solvers). Despite the advancement of computational resources, the challenge is to improve the SBDO framework (as a whole or its single components) in order to efficiently achieve accurate optimal solutions in solving complex engineering design problems.

The aim of this Special Issue is to collect state-of-the-art research on simulation-based optimization methods and their applications to complex engineering design problems. Relevant topics, methods, and applications are included in (but not limited to) the list below

  • Single- and multiobjective optimization algorithms;
  • Multidisciplinary optimization;
  • Metamodeling and machine learning in SBDO;
  • Multi-fidelity methods;
  • Dimensionality reduction;
  • Optimization under uncertainty;
  • Design modification methods;
  • Engineering design of aeronautical, aerospace, electrical, mechanical, naval applications.

Dr. Andrea Serani
Dr. ‪Riccardo Pellegrini
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Algorithms is an international peer-reviewed open access monthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 1600 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (8 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Editorial

Jump to: Research

2 pages, 157 KiB  
Editorial
Overview on the Special Issue on “Simulation-Based Optimization: Methods and Applications in Engineering Design”
by Riccardo Pellegrini and Andrea Serani
Algorithms 2023, 16(4), 191; https://doi.org/10.3390/a16040191 - 30 Mar 2023
Viewed by 1127
Abstract
The simulation-based design optimization (SBDO) paradigm is a well-known approach that has assisted, assists, and will continue to assist designers to develop ever-improving systems [...] Full article

Research

Jump to: Editorial

18 pages, 6796 KiB  
Article
Research of Flexible Assembly of Miniature Circuit Breakers Based on Robot Trajectory Optimization
by Yan Han, Liang Shu, Ziran Wu, Xuan Chen, Gaoyan Zhang and Zili Cai
Algorithms 2022, 15(8), 269; https://doi.org/10.3390/a15080269 - 31 Jul 2022
Cited by 5 | Viewed by 2230
Abstract
This paper is dedicated to achieving flexible automatic assembly of miniature circuit breakers (MCBs) to resolve the high rigidity issue of existing MCB assembly by proposing a flexible automatic assembly process and method with industrial robots. To optimize the working performance of the [...] Read more.
This paper is dedicated to achieving flexible automatic assembly of miniature circuit breakers (MCBs) to resolve the high rigidity issue of existing MCB assembly by proposing a flexible automatic assembly process and method with industrial robots. To optimize the working performance of the robot, a time-optimal trajectory planning method of the improved Particle Swarm Optimization (PSO) with a multi-optimization mechanism is proposed. The solution uses a fitness switch function for particle sifting to improve the stability of the acceleration and jerk of the robot motion as well as to increase the computational efficiency. The experimental results show that the proposed method achieves flexible assembly for multi-type MCB parts of varying postures. Compared with other optimization algorithms, the proposed improved PSO is significantly superior in both computational efficiency and optimization accuracy. Compared with the standard PSO, the proposed trajectory planning method shortens the assembly time by 6.9 s and raises the assembly efficiency by 16.7%. The improved PSO is implemented on the experimental assembly platform and achieves smooth and stable operations, which proves the high significance and practicality for MCB fabrication. Full article
Show Figures

Figure 1

Figure 1
<p>Internal structure of an MCB.</p>
Full article ">Figure 2
<p>Models of the parts to be assembled.</p>
Full article ">Figure 3
<p>Posture adjustment platform.</p>
Full article ">Figure 4
<p>Robot flexible multi-gripper claw.</p>
Full article ">Figure 5
<p>Auxiliary adjustment rack for the parts.</p>
Full article ">Figure 6
<p>Flexible assembly processes for the parts: (<b>a</b>) arc extinguishing cover posture adjustment; (<b>b</b>) position adjustment of other parts.</p>
Full article ">Figure 7
<p>Positioning carrier with adjusted parts.</p>
Full article ">Figure 8
<p>Schematic diagram of the robot connecting rod structure.</p>
Full article ">Figure 9
<p>Improve PSO process.</p>
Full article ">Figure 10
<p>Schematic diagram of robot segmentation.</p>
Full article ">Figure 11
<p>Joint motions before optimization: (<b>a</b>) Position curve; (<b>b</b>) Velocity curve; (<b>c</b>) Acceleration curve; (<b>d</b>) Jerk curve.</p>
Full article ">Figure 12
<p>Comparison of iterative process before and after PSO algorithm improvement.</p>
Full article ">Figure 13
<p>Joint motions after optimization: (<b>a</b>) Position curve; (<b>b</b>) Velocity curve; (<b>c</b>) Acceleration curve; (<b>d</b>) Jerk curve.</p>
Full article ">Figure 14
<p>Time-optimal trajectory planning by the improved PSO: (<b>a</b>) section AB <span class="html-italic">t</span><sub>1</sub> trajectory; (<b>b</b>) section AB <span class="html-italic">t</span><sub>2</sub> trajectory; (<b>c</b>) section AB <span class="html-italic">t</span><sub>3</sub> trajectory.</p>
Full article ">Figure 15
<p>Experimental platform of posture adjustment.</p>
Full article ">Figure 16
<p>Flexible assembly experiment.</p>
Full article ">
26 pages, 2207 KiB  
Article
Multi-Fidelity Low-Rank Approximations for Uncertainty Quantification of a Supersonic Aircraft Design
by Sihmehmet Yildiz, Hayriye Pehlivan Solak and Melike Nikbay
Algorithms 2022, 15(7), 250; https://doi.org/10.3390/a15070250 - 19 Jul 2022
Cited by 2 | Viewed by 2572
Abstract
Uncertainty quantification has proven to be an indispensable study for enhancing reliability and robustness of engineering systems in the early design phase. Single and multi-fidelity surrogate modelling methods have been used to replace the expensive high fidelity analyses which must be repeated many [...] Read more.
Uncertainty quantification has proven to be an indispensable study for enhancing reliability and robustness of engineering systems in the early design phase. Single and multi-fidelity surrogate modelling methods have been used to replace the expensive high fidelity analyses which must be repeated many times for uncertainty quantification. However, since the number of analyses required to build an accurate surrogate model increases exponentially with the number of random input variables, most surrogate modelling methods suffer from the curse of dimensionality. As an alternative approach, the Low-Rank Approximation method can be applied to high-dimensional uncertainty quantification studies with a low computational cost, where the number of coefficients for building the surrogate model increases only linearly with the number of random input variables. In this study, the Low-Rank Approximation method is implemented for multi-fidelity applications with additive and multiplicative correction approaches to make the high-dimensional uncertainty quantification analysis more efficient and accurate. The developed uncertainty quantification methodology is tested on supersonic aircraft design problems and its predictions are compared with the results of single- and multi-fidelity Polynomial Chaos Expansion and Monte Carlo methods. For the same computational cost, the Low-Rank Approximation method outperformed both in surrogate modeling and uncertainty quantification cases for all the benchmarks and real-world engineering problems addressed in the present study. Full article
Show Figures

Figure 1

Figure 1
<p>Comparing number of unknown coefficients between the PCE and LRA methods for high dimensional problems.</p>
Full article ">Figure 2
<p>Comparison of sampling methods [<a href="#B47-algorithms-15-00250" class="html-bibr">47</a>].</p>
Full article ">Figure 3
<p>Surrogate model error metrics for Park (1991) function (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>35</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 4
<p>Comparison of probability density function (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>h</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>20</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>35</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 5
<p>Surrogate model error metrics for Borehole function (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>200</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 6
<p>Comparison of probability density function (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>h</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>50</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>200</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 7
<p>Surrogate model error metrics for Sobol’-Levitan function (<math display="inline"><semantics> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> </semantics></math> = 10,000).</p>
Full article ">Figure 8
<p>Comparison of probability density function (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>h</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>2500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> </semantics></math> = 10,000).</p>
Full article ">Figure 9
<p>Aeroacoustic propagation in sonic boom prediction. Reproduced with permission from Ref. [<a href="#B51-algorithms-15-00250" class="html-bibr">51</a>].</p>
Full article ">Figure 10
<p>The JAXA Wing-Body geometry.</p>
Full article ">Figure 11
<p>Pressure coefficient distribution comparison (<b>a</b>) on the upper surface (<b>b</b>) on the lower surface.</p>
Full article ">Figure 12
<p>Pressure signature comparison: (<b>a</b>) Near-field pressure signature (<b>b</b>); ground signature for JWB.</p>
Full article ">Figure 13
<p>Surrogate model error metrics for sonic boom problem 1 (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>70</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 14
<p>Comparison of probability density function (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>70</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>h</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>40</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 15
<p>Surrogate model error metrics for sonic boom problem 2 (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math>).</p>
Full article ">Figure 16
<p>Comparison of probability density function (<b>a</b>) Results of all surrogate models (<b>b</b>) Results of LRA and MF-LRA (<math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>h</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>500</mn> </mrow> </semantics></math>, <math display="inline"><semantics> <mrow> <msub> <mi>N</mi> <mrow> <mi>l</mi> <mi>f</mi> </mrow> </msub> <mo>=</mo> <mn>1000</mn> </mrow> </semantics></math>).</p>
Full article ">
35 pages, 6328 KiB  
Article
Multi-Fidelity Gradient-Based Optimization for High-Dimensional Aeroelastic Configurations
by Andrew S. Thelen, Dean E. Bryson, Bret K. Stanford and Philip S. Beran
Algorithms 2022, 15(4), 131; https://doi.org/10.3390/a15040131 - 16 Apr 2022
Cited by 10 | Viewed by 3487
Abstract
The simultaneous optimization of aircraft shape and internal structural size for transonic flight is excessively costly. The analysis of the governing physics is expensive, in particular for highly flexible aircraft, and the search for optima using analysis samples can scale poorly with design [...] Read more.
The simultaneous optimization of aircraft shape and internal structural size for transonic flight is excessively costly. The analysis of the governing physics is expensive, in particular for highly flexible aircraft, and the search for optima using analysis samples can scale poorly with design space size. This paper has a two-fold purpose targeting the scalable reduction of analysis sampling. First, a new algorithm is explored for computing design derivatives by analytically linking objective definition, geometry differentiation, mesh construction, and analysis. The analytic computation of design derivatives enables the accurate use of more efficient gradient-based optimization methods. Second, the scalability of a multi-fidelity algorithm is assessed for optimization in high dimensions. This method leverages a multi-fidelity model during the optimization line search for further reduction of sampling costs. The multi-fidelity optimization is demonstrated for cases of aerodynamic and aeroelastic design considering both shape and structural sizing separately and in combination with design spaces ranging from 17 to 321 variables, which would be infeasible using typical, surrogate-based methods. The multi-fidelity optimization consistently led to a reduction in high-fidelity evaluations compared to single-fidelity optimization for the aerodynamic shape problems, but frequently resulted in a cost penalty for cases involving structural sizing. While the multi-fidelity optimizer was successfully applied to problems with hundreds of variables, the results underscore the importance of accurately computing gradients and motivate the extension of the approach to constrained optimization methods. Full article
Show Figures

Figure 1

Figure 1
<p>Baseline ESP representation of the uCRM geometry, with grey and red showing outer mold line (OML) and inner mold line (IML) geometries, respectively.</p>
Full article ">Figure 2
<p>ESP parameterization of the uCRM. (<b>a</b>) ESP geometry used for LF analyses showing designable airfoil shapes at the wing Yehudi and tip. (<b>b</b>) Faces of IML geometry components with varying thickness (color variations only used to denote different components).</p>
Full article ">Figure 3
<p>Extended Design Structure Matrix (XDSM) [<a href="#B25-algorithms-15-00131" class="html-bibr">25</a>] depicting couplings of meshing and analysis for multi-fidelity, multidisciplinary analysis. Multidisciplinary elements are shown in green, structures in yellow, and aerodynamics in blue. Low-fidelity aerodynamics’ pre-processing is distinguished from high-fidelity using dark blue with faded inputs and outputs.</p>
Full article ">Figure 4
<p>Pressure coefficient and stress (normalized by yield stress) at the same statically deflected design. The aluminum skin, rib, and spar faces have arbitrary thicknesses of 2, 3, and 4 cm, respectively. (<b>a</b>) Pressure coefficient, Euler. (<b>b</b>) Normalized von Mises stress, Euler. (<b>c</b>) Differential pressure coefficient, VLM. (<b>d</b>) Normalized von Mises stress, VLM.</p>
Full article ">Figure 5
<p>Outlines of the deflected geometries, showing the Euler model’s slightly larger deflection (black: undeflected, blue: deflected Euler, red: deflected VLM).</p>
Full article ">Figure 6
<p>HF adjoint, geometric sensitivities, and their dot product around the baseline aeroelastic solution. (<b>a</b>) Log-scale lift coefficient sensitivity. (<b>b</b>) Log-scale Yehudi twist sensitivity.</p>
Full article ">Figure 7
<p>Mean case preparation and run times (wall-clock) as a function of shape design variables. The adjoint-based analyses (Subplots (<b>b</b>–<b>d</b>)) include one forward and three adjoint solves. (<b>a</b>) Low-fidelity rigid; (<b>b</b>) High-fidelity rigid; (<b>c</b>) Low-fidelity aeroelastic; (<b>d</b>) High-fidelity aeroelastic.</p>
Full article ">Figure 8
<p>Extended design structure matrix diagram of the multi-fidelity BFGS process. Blue elements represent iterative processes (i.e., optimization convergence or line searching); red represents the fidelity switching decision; green represents the calculation of values involved in the optimization process.</p>
Full article ">Figure 9
<p>Convergence histories of aeroelastic problem P1a using SLSQP. (<b>a</b>) Objective; (<b>b</b>) Feasibility.</p>
Full article ">Figure 10
<p>Baseline and optimized geometries. For reference, the low- and high-fidelity angles of attack are 1.339148 and 0.8001988 degrees, respectively. (<b>a</b>) Yehudi; (<b>b</b>) Tip.</p>
Full article ">Figure 11
<p>Optimization results for airfoil design cases. In (<b>a</b>–<b>c</b>), solid and dotted lines represent single- and multi-fidelity results, respectively. (<b>a</b>) Penalty-constrained objective; (<b>b</b>) Feasibility; (<b>c</b>) Percent reduction, all; (<b>d</b>) Percent reduction, difference.</p>
Full article ">Figure 12
<p>Optimization results for structural sizing cases, fixed airfoils. (<b>a</b>) Penalty-constrained objective; (<b>b</b>) Feasibility; (<b>c</b>) Percent reduction, all; (<b>d</b>) Percent reduction, difference.</p>
Full article ">Figure 13
<p>Optimization results for structural sizing cases, variable airfoils. (<b>a</b>) Penalty-constrained objective; (<b>b</b>) Feasibility; (<b>c</b>) Percent reduction, all; (<b>d</b>) Percent reduction, difference.</p>
Full article ">Figure 14
<p>Baseline and optimized airfoil shapes, where aeroelastic cases show the undeflected jig shapes. (<b>a</b>) Yehudi, rigid; (<b>b</b>) Yehudi, aeroelastic; (<b>c</b>) Yehudi, aeroelastic with sizing; (<b>d</b>) Tip, rigid; (<b>e</b>) Tip, aeroelastic; (<b>f</b>) Tip, aeroelastic with sizing.</p>
Full article ">Figure 15
<p>Baseline and optimized jig shapes and deflected geometries in 1g cruise and 2.5 g pull-up conditions (7-DoF airfoil parameterization).</p>
Full article ">Figure 16
<p>Stress and thickness contours of the best shape+sizing optimization solution (P3a). (<b>a</b>) Structural thicknesses; (<b>b</b>) von Mises over yield stress for a +2.5 g maneuver.</p>
Full article ">Figure 17
<p>Performance of multi-fidelity model within airfoil design. (<b>a</b>) Rigid 7-DoF Kulfan; (<b>b</b>) Aeroelastic 7-DoF Kulfan.</p>
Full article ">Figure 18
<p>Performance of multi-fidelity model within simultaneous shape and structural design. (<b>a</b>) 7 DoF Kulfan; (<b>b</b>) 17 DoF Kulfan.</p>
Full article ">
26 pages, 622 KiB  
Article
Combinatorial Integral Approximation Decompositions for Mixed-Integer Optimal Control
by Clemens Zeile, Tobias Weber and Sebastian Sager
Algorithms 2022, 15(4), 121; https://doi.org/10.3390/a15040121 - 31 Mar 2022
Cited by 3 | Viewed by 2473
Abstract
Solving mixed-integer nonlinear programs (MINLPs) is hard from both a theoretical and practical perspective. Decomposing the nonlinear and the integer part is promising from a computational point of view. In general, however, no bounds on the objective value gap can be established and [...] Read more.
Solving mixed-integer nonlinear programs (MINLPs) is hard from both a theoretical and practical perspective. Decomposing the nonlinear and the integer part is promising from a computational point of view. In general, however, no bounds on the objective value gap can be established and iterative procedures with potentially many subproblems are necessary. The situation is different for mixed-integer optimal control problems with binary variables that switch over time. Here, a priori bounds were derived for a decomposition into one continuous nonlinear control problem and one mixed-integer linear program, the combinatorial integral approximation (CIA) problem. In this article, we generalize and extend the decomposition idea. First, we derive different decompositions and analyze the implied a priori bounds. Second, we propose several strategies to recombine promising candidate solutions for the binary control functions in the original problem. We present the extensions for ordinary differential equations-constrained problems. These extensions are transferable in a straightforward way, though, to recently suggested variants for certain partial differential equations, for algebraic equations, for additional combinatorial constraints, and for discrete time problems. We implemented all algorithms and subproblems in AMPL for a proof-of-concept study. Numerical results show the improvement compared to the standard CIA decomposition with respect to objective function value and compared to general-purpose MINLP solvers with respect to runtime. Full article
Show Figures

Figure 1

Figure 1
<p>Example visualization of the GreedyTime algorithm. We use two candidate control solutions, here from CIA and SCIA, to construct new candidates. We perform an enumeration between 0 and 1 at all times <math display="inline"><semantics> <msub> <mi>t</mi> <mi>j</mi> </msub> </semantics></math> when the input vectors differ. Then, the two candidate solutions <math display="inline"><semantics> <mi mathvariant="bold">w</mi> </semantics></math> are fixed and we evaluate (MIOCP) for both vectors. We compare the resulting objective function values with their previous values. Moreover, the binary <math display="inline"><semantics> <msub> <mi mathvariant="bold">w</mi> <mi>j</mi> </msub> </semantics></math> values with the lower objective values are fixed in the candidate solutions. This procedure is repeated on the next grid point with unequal candidate solutions.</p>
Full article ">Figure 2
<p>Visualization of the singular arc block recombination heuristic for two MILP control vectors (which we name CIA and SCIA) with three singular arcs. Every possible variation from the singular arcs and candidate controls is generated and we evaluate (MIOCP) for each of the constructed variation. The minimal objective value of all variations represents the heuristic’s result.</p>
Full article ">Figure 3
<p>Performance profile comparing the deviation of differential states based on SCIA and CIA solutions. Relaxed solutions are shown in maximum norm and log-scale. The results are based on the instances “Double tank (Multimode)” and “Lotka–Volterra (absolute fishing variant)” from the mintoc.de benchmark library. Using (SCIA1) or (SCIAmax) can improve the performance of the CIA decomposition significantly.</p>
Full article ">Figure 4
<p>Performance profile comparing objective deviation from the relaxed solution in percentage and log-scale of <math display="inline"><semantics> <mi>λ</mi> </semantics></math>-CIA and CIA solutions. The results are based on the instance “Quadrotor (binary variant)” from the mintoc.de benchmark library. (<math display="inline"><semantics> <mi>λ</mi> </semantics></math>CIA) appeared to provide no clear improvement compared with the (CIA) solutions.</p>
Full article ">Figure 5
<p>Performance profile comparing objective deviation from the relaxed solution in percentage and log-scale of (CIA) and its backward variant solutions. The results are based on the instance “Lotka–Volterra (terminal constraint violation)” from the mintoc.de benchmark library. Using (CIA1B) or (CIAmaxB) can improve the performance of the CIA decomposition significantly.</p>
Full article ">Figure 6
<p>Box plot comparing objective deviation from the relaxed solution in percentage and log-scale of several MILP (marked in blue) and recombination heuristic (marked in red) solutions. The results are based on instances from the mintoc.de benchmark library. The box borders are 1/4 and 3/4-quantiles, whereas the whiskers represent 1/20 and 19/20-quantiles. We visualize the median values by black lines in the box and additionally display them numerically above the box. We represent the average values of the respective algorithms by red asterisks and the outliers by black crosses. The boxes of recombination strategies are shifted towards lower objective values compared with (CIA) algorithms and, thus, can improve the CIA decomposition performance significantly.</p>
Full article ">Figure 7
<p>Log plot of run time and objective value deviation from the relaxed solution of the constructed solutions for different approaches and for the Lotka–Volterra multimode problem, with differential state discretization <span class="html-italic">N</span> = 12,000. The numbers in the plot indicate the applied corresponding number of control grid points <span class="html-italic">M</span>. We illustrate the outcomes of the solutions constructed by (CIAmax), <math display="inline"><semantics> <mrow> <mi>Opt</mi> <mo>(</mo> <msup> <mi>S</mi> <mi>CIA</mi> </msup> <mo>)</mo> </mrow> </semantics></math>, the GreedyTime recombination heuristic, and the MINLP solver Bonmin. By <math display="inline"><semantics> <mrow> <mi>Opt</mi> <mo>(</mo> <msup> <mi>S</mi> <mi>CIA</mi> </msup> <mo>)</mo> </mrow> </semantics></math>, we denote the best objective value outcome of over all MILP solutions. For each control discretization, we connect the outcomes of the four approaches with lines in order to compare the behavior for different discretizations. One observes the convergence of all approaches towards the lower bound provided by the relaxed solution and the closure of the gap between (CIAmax) and Bonmin solutions for a fixed discretization. GreedyTime is roughly two orders of magnitude slower than (CIAmax), but is faster than Bonmin.</p>
Full article ">
13 pages, 3712 KiB  
Article
Dynamic Line Scan Thermography Parameter Design via Gaussian Process Emulation
by Simon Verspeek, Ivan De Boi, Xavier Maldague, Rudi Penne and Gunther Steenackers
Algorithms 2022, 15(4), 102; https://doi.org/10.3390/a15040102 - 22 Mar 2022
Cited by 1 | Viewed by 2296
Abstract
We address the challenge of determining a valid set of parameters for a dynamic line scan thermography setup. Traditionally, this optimization process is labor- and time-intensive work, even for an expert skilled in the art. Nowadays, simulations in software can reduce some of [...] Read more.
We address the challenge of determining a valid set of parameters for a dynamic line scan thermography setup. Traditionally, this optimization process is labor- and time-intensive work, even for an expert skilled in the art. Nowadays, simulations in software can reduce some of that burden. However, when faced with many parameters to optimize, all of which cover a large range of values, this is still a time-consuming endeavor. A large number of simulations are needed to adequately capture the underlying physical reality. We propose to emulate the simulator by means of a Gaussian process. This statistical model serves as a surrogate for the simulations. To some extent, this can be thought of as a “model of the model”. Once trained on a relative low amount of data points, this surrogate model can be queried to answer various engineering design questions. Moreover, the underlying model, a Gaussian process, is stochastic in nature. This allows for uncertainty quantification in the outcomes of the queried model, which plays an important role in decision making or risk assessment. We provide several real-world examples that demonstrate the usefulness of this method. Full article
Show Figures

Figure 1

Figure 1
<p>Visualization of the finite element simulation consisting of a flat bottom hole plate (blue) and a line heater (yellow). The line heater moves above the sample in a linear motion. For a more detailed figure and explanation, we refer the reader to [<a href="#B2-algorithms-15-00102" class="html-bibr">2</a>].</p>
Full article ">Figure 2
<p>Response surface as generated in [<a href="#B2-algorithms-15-00102" class="html-bibr">2</a>]. The surface is generated from 1000 finite element simulations, using eight input parameters. The simplified response surface has all input parameters fixed, except for the heat load and the source velocity. The fixed parameters are: <math display="inline"><semantics> <msub> <mi>d</mi> <mrow> <mi>h</mi> <mi>e</mi> <mi>a</mi> <mi>t</mi> <mo>−</mo> <mi>c</mi> <mi>a</mi> <mi>m</mi> </mrow> </msub> </semantics></math> = 425 mm, <math display="inline"><semantics> <msub> <mi>d</mi> <mrow> <mi>s</mi> <mi>t</mi> <mi>a</mi> <mi>r</mi> <mi>t</mi> </mrow> </msub> </semantics></math> = 5.8 mm, <math display="inline"><semantics> <msub> <mi>D</mi> <mrow> <mi>h</mi> <mi>o</mi> <mi>l</mi> <mi>e</mi> </mrow> </msub> </semantics></math> = 9 mm, <math display="inline"><semantics> <msub> <mi>d</mi> <mrow> <mi>h</mi> <mi>e</mi> <mi>i</mi> <mi>g</mi> <mi>h</mi> <mi>t</mi> </mrow> </msub> </semantics></math> = 430 mm, <math display="inline"><semantics> <msub> <mi>T</mi> <mrow> <mi>a</mi> <mi>m</mi> <mi>b</mi> <mi>i</mi> <mi>e</mi> <mi>n</mi> <mi>t</mi> </mrow> </msub> </semantics></math> = 48 <math display="inline"><semantics> <msup> <mrow/> <mo>°</mo> </msup> </semantics></math>C. Using this response surface, one can find the best temperature difference as a valley or top.</p>
Full article ">Figure 3
<p>Graphical visualization of the learning process of the Gaussian process for 5 runs of 500 iterations. (<b>a</b>) represents the root mean square error (RMSE) of the learned surrogate compared to the response surface created in [<a href="#B2-algorithms-15-00102" class="html-bibr">2</a>]. (<b>b</b>) shows the average standard deviation of the Gaussian process posterior prediction.</p>
Full article ">Figure 4
<p>Visualization of the temperature difference for six different defect diameters. (<b>a</b>) 12 mm, (<b>b</b>) 14 mm, (<b>c</b>) 16 mm, (<b>d</b>) 18 mm, (<b>e</b>) 20 mm and (<b>f</b>) 22 mm. Red indicates temperature differences that might result in damaging the sample under inspection. These plots serve as a warning when designing a setup.</p>
Full article ">Figure 4 Cont.
<p>Visualization of the temperature difference for six different defect diameters. (<b>a</b>) 12 mm, (<b>b</b>) 14 mm, (<b>c</b>) 16 mm, (<b>d</b>) 18 mm, (<b>e</b>) 20 mm and (<b>f</b>) 22 mm. Red indicates temperature differences that might result in damaging the sample under inspection. These plots serve as a warning when designing a setup.</p>
Full article ">Figure 5
<p>Visualization of the temperature difference for six different defect diameters. (<b>a</b>) 12 mm, (<b>b</b>) 14 mm, (<b>c</b>) 16 mm, (<b>d</b>) 18 mm, (<b>e</b>) 20 mm and (<b>f</b>) 22 mm. Red are temperature differences below 5 °C, yellow above 25 °C and green in between. Only the green regions are of practical value in real-world applications.</p>
Full article ">
19 pages, 1080 KiB  
Article
Multi-Fidelity Sparse Polynomial Chaos and Kriging Surrogate Models Applied to Analytical Benchmark Problems
by Markus P. Rumpfkeil, Dean Bryson and Phil Beran
Algorithms 2022, 15(3), 101; https://doi.org/10.3390/a15030101 - 21 Mar 2022
Cited by 4 | Viewed by 2762
Abstract
In this article, multi-fidelity kriging and sparse polynomial chaos expansion (SPCE) surrogate models are constructed. In addition, a novel combination of the two surrogate approaches into a multi-fidelity SPCE-Kriging model will be presented. Accurate surrogate models, once obtained, can be employed for evaluating [...] Read more.
In this article, multi-fidelity kriging and sparse polynomial chaos expansion (SPCE) surrogate models are constructed. In addition, a novel combination of the two surrogate approaches into a multi-fidelity SPCE-Kriging model will be presented. Accurate surrogate models, once obtained, can be employed for evaluating a large number of designs for uncertainty quantification, optimization, or design space exploration. Analytical benchmark problems are used to show that accurate multi-fidelity surrogate models can be obtained at lower computational cost than high-fidelity models. The benchmarks include non-polynomial and polynomial functions of various input dimensions, lower dimensional heterogeneous non-polynomial functions, as well as a coupled spring-mass-system. Overall, multi-fidelity models are more accurate than high-fidelity ones for the same cost, especially when only a few high-fidelity training points are employed. Full-order PCEs tend to be a factor of two or so worse than SPCES in terms of overall accuracy. The combination of the two approaches into the SPCE-Kriging model leads to a more accurate and flexible method overall. Full article
Show Figures

Figure 1

Figure 1
<p>MF kriging algorithm flowchart.</p>
Full article ">Figure 2
<p>All four fidelity levels of Forrester function.</p>
Full article ">Figure 3
<p>Rosenbrock function (from left to right, high-, medium-, and low-fidelity).</p>
Full article ">Figure 4
<p>Shifted-rotated Rastrigin function (from left to right, high-, medium-, and low-fidelity).</p>
Full article ">Figure 5
<p>Plot of one- (<b>left</b>) and two-dimensional (<b>right</b>) HF and LF ALOS functions with initial HF training points.</p>
Full article ">Figure 6
<p>Springs connecting three masses.</p>
Full article ">Figure 7
<p>Analytical solution for the spring-mass problem for <math display="inline"><semantics> <mrow> <mn>0</mn> <mo>≤</mo> <mi>t</mi> <mo>≤</mo> <mn>30</mn> </mrow> </semantics></math>.</p>
Full article ">Figure 8
<p>Numerical solution for <math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>)</mo> </mrow> </mrow> </semantics></math> of the example problem for <math display="inline"><semantics> <mrow> <mn>0</mn> <mo>≤</mo> <mi>t</mi> <mo>≤</mo> <mn>6</mn> </mrow> </semantics></math> using <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>t</mi> <mo>=</mo> <mn>0.01</mn> </mrow> </semantics></math> labeled as high-fidelity (HF) and <math display="inline"><semantics> <mrow> <mo>Δ</mo> <mi>t</mi> <mo>=</mo> <mn>0.6</mn> </mrow> </semantics></math> labeled as low-fidelity (LF).</p>
Full article ">Figure 9
<p><math display="inline"><semantics> <mrow> <msub> <mi>x</mi> <mn>1</mn> </msub> <mrow> <mo>(</mo> <mi>t</mi> <mo>=</mo> <mn>6</mn> <mo>)</mo> </mrow> </mrow> </semantics></math> when <math display="inline"><semantics> <mrow> <mn>1</mn> <mo>≤</mo> <msub> <mi>k</mi> <mn>1</mn> </msub> <mo>,</mo> <msub> <mi>k</mi> <mn>2</mn> </msub> <mo>≤</mo> <mn>4</mn> </mrow> </semantics></math> with HF and LF results shown in red and blue, respectively.</p>
Full article ">Figure 10
<p>Forrester function RMSE (from left to right, PCE and kriging).</p>
Full article ">Figure 11
<p>PCE: Rosenbrock function RMSE (from left to right, two, five and ten dimensions). PCE and SPCE in dashed and solid lines, respectively.</p>
Full article ">Figure 12
<p>Kriging: Rosenbrock function RMSE (from left to right, two, five and ten dimensions). Ordinary and SPCE-Kriging in dashed and solid lines, respectively.</p>
Full article ">Figure 13
<p>PCE: Rastrigin function RMSE (from left to right, two, five and ten dimensions). PCE and SPCE in dashed and solid lines, respectively.</p>
Full article ">Figure 14
<p>Kriging: Rastrigin function RMSE (from left to right, two, five and ten dimensions). Ordinary and SPCE-Kriging in dashed and solid lines, respectively.</p>
Full article ">Figure 15
<p>PCE: ALOS function RMSE (from left to right, one, two and three dimensions). PCE and SPCE in dashed and solid lines, respectively.</p>
Full article ">Figure 16
<p>Kriging: ALOS function RMSE (from left to right, one, two and three dimensions). Ordinary and SPCE-Kriging in dashed and solid lines, respectively.</p>
Full article ">Figure 17
<p>Spring-mass system (only springs) RMSE (from left to right, PCE and Kriging).</p>
Full article ">Figure 18
<p>Spring-Mass System RMSE (from left to right, PCE and Kriging).</p>
Full article ">
16 pages, 4336 KiB  
Article
Design of Selective Laser Melting (SLM) Structures: Consideration of Different Material Properties in Multiple Surface Layers Resulting from the Manufacturing in a Topology Optimization
by Jan Holoch, Sven Lenhardt, Sven Revfi and Albert Albers
Algorithms 2022, 15(3), 99; https://doi.org/10.3390/a15030099 - 19 Mar 2022
Cited by 4 | Viewed by 3106
Abstract
Topology optimization offers a possibility to derive load-compliant structures. These structures tend to be complex, and conventional manufacturing offers only limited possibilities for their production. Additive manufacturing provides a remedy due to its high design freedom. However, this type of manufacturing can cause [...] Read more.
Topology optimization offers a possibility to derive load-compliant structures. These structures tend to be complex, and conventional manufacturing offers only limited possibilities for their production. Additive manufacturing provides a remedy due to its high design freedom. However, this type of manufacturing can cause areas of different material properties in the final part. For example, in selective laser melting, three areas of different porosity can occur depending on the process parameters, the geometry of the part and the print direction, resulting in a direct interrelation between manufacturing and design. In order to address this interrelation in design finding, this contribution presents an optimization method in which the three porous areas are identified and the associated material properties are considered iteratively in a topology optimization. For this purpose, the topology optimization is interrupted in each iteration. Afterwards, the three areas as well as the material properties are determined and transferred back to the topology optimization, whereby those properties are used for the calculation of the next iteration. By using the optimization method, a design with increased volume-specific stiffness compared to a design of a standard topology optimization can be created and will be used in the future as a basis for the extension by a global strength constraint to maintain the maximum permissible stress and the minimum wall thickness. Full article
Show Figures

Figure 1

Figure 1
<p>Exemplary representation (cut through <span class="html-italic">x</span>–<span class="html-italic">y</span> plane) of the three porous areas (<b>left</b>) and CT scan to visualize the porosity distribution in one print layer (<b>right</b>) adapted from [<a href="#B17-algorithms-15-00099" class="html-bibr">17</a>].</p>
Full article ">Figure 2
<p>Workflow of the developed optimization method.</p>
Full article ">Figure 3
<p>Surface meshes of the interim result (<b>top</b>) as well as offset 1 (<b>center</b>) and offset 2 (<b>bottom</b>).</p>
Full article ">Figure 4
<p>Surface meshes (<b>top</b>) as well as meshed areas including assigned material properties of the interim result (<b>bottom</b>).</p>
Full article ">Figure 5
<p>Porous areas including assigned material properties of the interim result considering the print direction (<b>top</b>) and material properties transferred to the initial topology optimization mesh (<b>bottom</b>).</p>
Full article ">Figure 6
<p>Three-point bending beam with dimensions, clamping and loading.</p>
Full article ">Figure 7
<p>Result of the developed optimization method: overall design (<b>left</b>) and sectional view (<b>right</b>).</p>
Full article ">Figure 8
<p>Result of the standard topology optimization: overall design (<b>left</b>) and sectional view (<b>right</b>).</p>
Full article ">Figure 9
<p>Comparison of the optimization history of the standard topology optimization and the developed optimization method.</p>
Full article ">Figure 10
<p>Resulting design of the developed optimization method including porous areas (<b>left</b>) and result of the static FE analysis (<b>right</b>).</p>
Full article ">Figure 11
<p>Resulting design of the standard topology optimization including porous areas (<b>left</b>) and result of the static FE analysis (<b>right</b>).</p>
Full article ">Figure 12
<p>Section in the layer plane through the design from the developed optimization method: front view (<b>top</b>) and plan view (<b>bottom</b>).</p>
Full article ">
Back to TopTop