Skip to main content
Log in

Coupling and selecting constraints in Bayesian optimization under uncertainties

  • Research Article
  • Published:
Optimization and Engineering Aims and scope Submit manuscript

Abstract

We consider Reliability-based Robust Design Optimization (RRDO) where it is sought to optimize the mean of an objective function while satisfying constraints in probability. The high computational cost of the simulations underlying the objective and constraints strongly limits the number of evaluations and makes this type of problems particularly challenging. The numerical cost issue and the parametric uncertainties have been addressed with Bayesian optimization algorithms which leverage Gaussian processes of the objective and constraint functions. Current Bayesian optimization algorithms call the objective and constraint functions simultaneously at each iteration. This is often not necessary and overlooks calculation savings opportunity. This article proposes a new efficient RRDO Bayesian optimization algorithm that optimally selects for evaluation, not only the usual design variables, but also one or several constraints along with the uncertain parameters. The algorithm relies on a multi-output Gaussian model of the constraints. The coupling of constraints and their separated selection are gradually implemented in three algorithm variants which are compared to a reference Bayesian approach. The results are promising in terms of convergence speed, accuracy and stability as observed on a two, a four and a 27-dimensional problem.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

Notes

  1. The Kriging Believer assumption states that the value of a function at an unobserved point is equal to the kriging prediction at that point. For our constraints, it means \(g_p(\tilde{\textbf{x}}\tilde{\textbf{u}}) = {{\textbf{m}}_{\textbf{G}}^{(t)}}_p(\tilde{\textbf{x}},\tilde{\textbf{u}}) ~,~p=1,\ldots ,l\).

References

  • Alvarez MA, Rosasco L, Lawrence ND (2011) Kernels for vector-valued functions: A review. arXiv preprint arXiv:1106.6251

  • Balesdent M, Brevault L, Morio J, Chocat R (2020) Overview of problem formulations and optimization algorithms in the presence of uncertainty. Springer, Cham, pp 147–183. https://doi.org/10.1007/978-3-030-39126-3_5

  • Bect J, Ginsbourger D, Li L, Picheny V, Vazquez E (2012) Sequential design of computer experiments for the estimation of a probability of failure. Stat Comput 22(3):773–793

    Article  MathSciNet  Google Scholar 

  • Beland JJ, Nair PB (2017) Bayesian optimization under uncertainty. In: NIPS BayesOpt 2017 workshop

  • Brooks CJ, Forrester A, Keane A, Shahpar S (2011) Multi-fidelity design optimisation of a transonic compressor rotor

  • Cakmak S, Astudillo R, Frazier P, Zhou E (2020) Bayesian optimization of risk measures. arXiv preprint arXiv:2007.05554

  • Chen Z, Peng S, Li X, Qiu H, Xiong H, Gao L, Li P (2015) An important boundary sampling method for reliability-based design optimization using kriging model. Struct Multidiscip Optim 52(1):55–70

    Article  MathSciNet  Google Scholar 

  • Cousin A, Garnier J, Guiton M, Zuniga M (2020) Chance constraint optimization of a complex system: Application to the fatigue design of a floating offshore wind turbine mooring system. In: WCCM-ECCOMAS2020. https://www.scipedia.com/public/Cousin_et_al_2021a

  • Deville Y, Ginsbourger D, Contributors OR, Durrande N, Deville MY, Rcpp D, DiceKriging S, Imports M, Rcpp L (2015) Package ‘kergp’

  • Dubourg V, Sudret B, Bourinet J-M (2011) Reliability-based design optimization using kriging surrogates and subset simulation. Struct Multidiscip Optim 44(5):673–690

    Article  Google Scholar 

  • Dunham J (1998) CFD validation for propulsion system components. Technical report, Advisory Group for Aerospace Research and Development, Neuilly-sur-Seine (France)

    Google Scholar 

  • El Amri R, Le Riche R, Helbert C, Blanchet-Scalliet C, Da Veiga S (2021) A sampling criterion for constrained Bayesian optimization with uncertainties. https://arxiv.org/abs/2103.05706

  • Fernández-Godino MG, Park C, Kim N-H, Haftka RT (2016) Review of multi-fidelity models. arXiv preprint arXiv:1609.07196

  • Garland N, Le Riche R, Richet Y, Durrande N (2020) Multi-fidelity for MDO using gaussian processes. In: Aerospace system analysis and optimization in uncertainty. Springer, pp 295–320

  • Genz A, Bretz F (2009) Computation of Multivariate Normal and T Probabilities, vol 195. Springer

  • Ghoreishi SF, Allaire D (2019) Multi-information source constrained Bayesian optimization. Struct Multidiscip Optim 59(3):977–991

    Article  MathSciNet  Google Scholar 

  • Goulard M, Voltz M (1992) Linear coregionalization model: tools for estimation and choice of cross-variogram matrix. Math Geol 24(3):269–286

    Article  Google Scholar 

  • Hernández-Lobato JM, Gelbart MA, Adams RP, Hoffman MW, Ghahramani Z (2016) A general framework for constrained Bayesian optimization using information-based search. J Mach Learn Res 17:1–53

    MathSciNet  Google Scholar 

  • Janusevskis J, Le Riche R (2012) Simultaneous kriging-based estimation and optimization of mean response. J Glob Optim. https://doi.org/10.1007/s10898-011-9836-5

    Article  Google Scholar 

  • Jones DR, Schonlau M, Welch WJ (1998) Efficient global optimization of expensive black-box functions. J Global Optim 13(4):455–492

    Article  MathSciNet  Google Scholar 

  • Kennedy MC, O’Hagan A (2000) Predicting the output from a complex computer code when fast approximations are available. Biometrika 87(1):1–13

    Article  MathSciNet  Google Scholar 

  • Lacaze S, Missoum S (2013) Reliability-based design optimization using kriging and support vector machines. In: Proceedings of the 11th international conference on structural safety & reliability, New York

  • Le Gratiet L, Garnier J (2014) Recursive co-kriging model for design of computer experiments with multiple levels of fidelity. Int J Uncertain Quantific 4(5):1

    Article  MathSciNet  Google Scholar 

  • Li X, Qiu H, Chen Z, Gao L, Shao X (2016) A local kriging approximation method using mpp for reliability-based design optimization. Comput Struct 162:102–115

    Article  Google Scholar 

  • Mattrand C, Beaurepaire P, Gayton N (2021) Adaptive kriging-based methods for failure probability evaluation: Focus on ak methods. Mechanical Engineering in Uncertainties From Classical Approaches to Some Recent Developments 205:1

    Google Scholar 

  • Meliani M, Bartoli N, Lefebvre T, Bouhlel, M-A, Martins JR, Morlier J (2019) Multi-fidelity efficient global optimization: methodology and application to airfoil shape design. In: AIAA Aviation 2019 Forum, p 3236

  • Menz M, Gogu C, Dubreuil S, Bartoli N, Morio J (2020) Adaptive coupling of reduced basis modeling and kriging based active learning methods for reliability analyses. Reliabil Eng Syst Saf 196:106771

    Article  Google Scholar 

  • Mockus J (2012) Bayesian approach to global optimization: theory and applications, vol 37. Springer, Berlin

    Google Scholar 

  • Moustapha M, Sudret B, Bourinet J-M, Guillaume B (2016) Quantile-based optimization under uncertainties using adaptive kriging surrogate models. Struct Multidiscip Optim 54(6):1403–1421

    Article  MathSciNet  Google Scholar 

  • Nemirovski A (2012) On safe tractable approximations of chance constraints. Eur J Oper Res 219(3):707–718

    Article  MathSciNet  Google Scholar 

  • Peherstorfer B, Willcox K, Gunzburger M (2018) Survey of multifidelity methods in uncertainty propagation, inference, and optimization. SIAM Rev 60(3):550–591

    Article  MathSciNet  Google Scholar 

  • Pelamatti J, Brevault L, Balesdent M, Talbi E-G, Guerin Y (2020) Overview and comparison of gaussian process-based surrogate models for mixed continuous and discrete variables: application on aerospace design problems. In: High-performance simulation-based optimization. Springer , pp 189–224

  • Perdikaris P, Raissi M, Damianou A, Lawrence ND, Karniadakis GE (2017) Nonlinear information fusion algorithms for data-efficient multi-fidelity modelling. Proc R Soc A: Math Phys Eng Sci 473(2198):20160751

    Article  CAS  ADS  Google Scholar 

  • Picheny V, Wagner T, Ginsbourger D (2013) A benchmark of kriging-based infill criteria for noisy optimization. Struct Multidiscip Optim 48(3):607–626

    Article  Google Scholar 

  • Powell MJ (1994) A direct search optimization method that models the objective and constraint functions by linear interpolation. In: Advances in optimization and numerical analysis. Springer, pp 51–67

  • Qian PZG, Wu H, Wu CJ (2008) Gaussian process models for computer experiments with qualitative and quantitative factors. Technometrics 50(3):383–396

    Article  MathSciNet  Google Scholar 

  • Rasmussen CE (2003) Gaussian processes in machine learning. In: Summer school on machine learning. Springer, pp 63–71

  • Reid L, Moore RD (1978) Design and overall performance of four highly loaded, high speed inlet stages for an advanced high-pressure-ratio core compressor. Technical report, NASA (1978). report NASA-TP-1337

  • Ribaud M, Blanchet-Scalliet C, Helbert C, Gillot F (2020) Robust optimization: a kriging-based multi-objective optimization approach. Reliab Eng Syst Saf 200:106913. https://doi.org/10.1016/j.ress.2020.106913

    Article  Google Scholar 

  • Roustant O, Ginsbourger D, Deville Y (2012) Dicekriging, diceoptim: Two r packages for the analysis of computer experiments by kriging-based metamodeling and optimization

  • Schonlau M, Welch WJ, Jones DR (1998) Global versus local search in constrained optimization of computer models. Lecture Notes-Monograph Series, pp 11–25

  • Scott W, Frazier P, Powell W (2011) The correlated knowledge gradient for simulation optimization of continuous parameters using gaussian process regression. SIAM J Optim 21(3):996–1026

    Article  MathSciNet  Google Scholar 

  • Shah A, Ghahramani Z (2016) Pareto frontier learning with expensive correlated objectives. In: International conference on machine learning. PMLR, pp 1919–1927

  • Swiler LP, Hough PD, Qian P, Xu X, Storlie C, Lee H (2014) Surrogate models for mixed discrete-continuous variables. In: Constraint programming and decision making. Springer, pp 181–202

  • Tao S, Van Beek A, Apley DW, Chen W (2021) Multi-model Bayesian optimization for simulation-based design. J Mech Des 143(11):111701

    Article  Google Scholar 

  • Torossian L, Picheny V, Durrande N (2020) Bayesian quantile and expectile optimisation. arXiv preprint arXiv:2001.04833

  • Valdebenito MA, Schuëller GI (2010) A survey on approaches for reliability-based optimization. Struct Multidiscip Optim 42(5):645–663

    Article  MathSciNet  Google Scholar 

  • Williams C, Bonilla EV, Chai KM (2007) Multi-task gaussian process prediction. Adv Neural Inf Process Syst 1:153–160

    Google Scholar 

  • Yao W, Chen X, Luo W, Van Tooren M, Guo J (2011) Review of uncertainty-based multidisciplinary design optimization methods for aerospace vehicles. Prog Aerosp Sci 47(6):450–479

    Article  Google Scholar 

  • Zhang Y, Notz WI (2015) Computer experiments with qualitative and quantitative variables: a review and reexamination. Qual Eng 27(1):2–13

    Article  Google Scholar 

  • Zhang J, Taflanidis A, Medina J (2017) Sequential approximate optimization for design under uncertainty problems utilizing kriging metamodeling in augmented input space. Comput Methods Appl Mech Eng 315:369–395

    Article  MathSciNet  ADS  Google Scholar 

  • Zhang Y, Tao S, Chen W, Apley DW (2020) A latent variable approach to gaussian process modeling with qualitative and quantitative factors. Technometrics 62(3):291–302

    Article  MathSciNet  Google Scholar 

  • Zhou Q, Qian PZ, Zhou S (2011) A simple approach to emulation for computer models with qualitative and quantitative factors. Technometrics 53(3):266–273

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

This work was partly supported by the OQUAIDO research Chair in Applied Mathematics.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Julien Pelamatti.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Main notations and acronyms

figure b

Appendix B: Covariance kernel for a nominal input: the hypersphere decomposition.

The hypersphere decomposition (Zhou et al. 2011) is a possible choice to parameterize a discrete covariance kernel. The underlying idea is to map each of the l levels \( \{z_1,..,z_p,...,z_l\}\) of the considered discrete variable onto a distinct point on the surface of a l-dimensional hypersphere by:

$$\begin{aligned}{} & {} \phi : \{z_1,..,z_p,...,z_l\}\rightarrow \mathbb {R}^l \\{} & {} \phi (z_p) = \sigma _z [ b_{p,1}, \dots , b_{p,l}]^\top \nonumber \quad \text{ for } p=1,\ldots , l \end{aligned}$$

where \(b_{p,r}\) represents the r-th coordinate of the p-th discrete level mapping, and is calculated as follows:

$$\begin{aligned}&b_{1,1} = 1&\\&b_{p,r} = \cos \theta _{p,r} \prod _{k = 1}^{r-1} \sin \theta _{p,k}&\text{ for } r = 1,\dots ,p-1 \\&b_{p,p} =\prod _{k = 1}^{p-1} \sin \theta _{p,k}&\text{ for } p \ne 1 \\&b_{p,r} = 0&\text{ for } r \ge p \ne 1 \end{aligned}$$

with \(-\pi \le \theta _{p,r} \le \pi \). It can be noticed that in the equations above, some of the mapping coordinates are arbitrarily set to 0. This allows to avoid rotation indeterminacies (i.e., an infinite number of hyperparameter sets characterizing the same covariance matrix), while also reducing the number of parameters required to define the mapping. The resulting kernel is then computed as the Euclidean scalar product between the hypersphere mappings,

$$\begin{aligned} k(z,z') =\phi (z)^\top \phi (z')~. \end{aligned}$$

The discrete kernel can then be characterized as an \(l\times l\) symmetric positive definite matrix \(\textbf{T}\) containing the covariance values between the discrete variable levels computed as:

$$\begin{aligned} \textbf{T} = \textbf{L}\textbf{L}^\top \end{aligned}$$

where each element of \(\textbf{L}_{i,j}\) is computed as \(b_{i,j}\),

$$\begin{aligned} \textbf{L} = \sigma _z \left[ \begin{array}{c c c c c} 1 &{} 0 &{} \dots &{} \dots &{} 0 \\ \cos \theta _{2,1} &{} \sin \theta _{2,1} &{} 0 &{} \dots &{} \dots \\ \vdots &{} \vdots &{} \vdots &{} \vdots &{} \vdots \\ \cos \theta _{l,1} &{} \sin \theta _{l,1} \cos \theta _{l,2} &{} \dots &{} \cos \theta _{l,l-1} \prod _{r = 1}^{l-2} \sin \theta _{l,r} &{} \prod _{r = 1}^{l-1} \sin \theta _{l,r} \end{array} \right] ~. \end{aligned}$$

In the output-as-input model of the constraints, Equation (3), the matrix \(\textbf{T}\) contains the covariance terms related to the constraint index, \(k_p(i,j) = \textbf{T}_{ij}~,~i,j = \{1,\ldots ,l\}\).

Appendix C: Probability of feasibility with dependent constraints

The probability of satisfying the coupled constraints intervenes in Eq. (6) for the EFI acquisition criterion. It can be estimated with GPs as:

$$\begin{aligned} \mathbb {P}(C^{(t)}(\textbf{x}) \le 0) \approx \frac{1}{N} \sum \limits _{k=1}^{N} \mathbb {1}_{\big (1 - \alpha - \frac{1}{M} \sum \limits _{j=1}^{M}\mathbb {1}_{\big (\textbf{G}^{(t)}(\textbf{x},\mathbf {u_j},\omega _k) \le \textbf{0} \big )} \le 0\big )}. \end{aligned}$$
(17)

In practice, a set of M instances of the random variables \(\textbf{U}\) is sampled from \(\rho _U\). Subsequently, N independent multi-output trajectories of \(\textbf{G}(\textbf{x},\cdot )\) are simulated at the aforementioned sampled uncertain parameters. The probability of feasibility can finally be computed by simply counting the number of trajectories for which the ratio of samples associated to feasible constraints is larger than \(1-\alpha \). Note that a multi-output GP prediction of the constraint vector is defined as feasible when all of its components are below their specific threshold value (which is 0 in this work).

In (7), the improvement is computed with respect to the incumbent best feasible solution \(z_{min}^{feas}\). However, the objective function mean is not observed therefore \(z_{min}^{feas}\) cannot be read in the data set used to condition the GPs F and \(\textbf{G}\). For this reason, in El Amri et al. (2021) the incumbent best feasible solution is defined by taking into account the mean of the GP \(Z^{(t)}(\textbf{x})\) and the expected value of the process \(C^{(t)}(\textbf{x})\):

$$\begin{aligned} z_{min}^{feas} = \min _{\textbf{x}} m^{(t)}_Z(\textbf{x}) ~\text{ s.t. }~ \mathbb E[ C^{(t)}(\textbf{x}) ] \le 0~. \end{aligned}$$
(18)

Given that the Fubini condition holds since the value of \(C(\cdot )\) is bounded by definition, the expected value of the process C can be written

$$\begin{aligned} \mathbb E[C^{(t)}(\textbf{x})]= & {} 1 - \alpha - \mathbb E_\textbf{U}[\mathbb E[1_{\textbf{G} (\textbf{x}, \textbf{U}) \le \textbf{0} } ]] \nonumber \\= & {} 1 - \alpha - \int _{\mathbb {R}^m} \mathbb P(\textbf{G} (\textbf{x}, \textbf{u}) \le \textbf{0}))\rho _\textbf{U} (\textbf{u}) d\textbf{u}\nonumber \\= & {} 1 - \alpha - \int _{\mathbb {R}^m} \Phi \left( \textbf{0} -\textbf{m}_\textbf{G}^{(t)}(\textbf{x},\textbf{u}),\textbf{K}_\textbf{G}^{(t)}(\textbf{x},\textbf{u}) \right) \rho _\textbf{U} (\textbf{u}) d\textbf{u} \end{aligned}$$

where \(\Phi (\cdot )\) is the cumulative distribution function of a multivariate Gaussian distribution which, like the univariate version, is estimated numerically (Genz and Bretz 2009). The above Equation and Eq. (18) are used to compute the current feasible minimum. If Eq. (18) yields no solution, \(m^{(t)}_Z(\textbf{x})\) at the \(\textbf{x}\) providing the largest probability of feasibility is taken as the incumbent optimal solution.

Appendix D: Estimating the proxy of the one-step-ahead feasible improvement variance

During the optimization process, the value of the coordinates \(\textbf{u}^{t+1}\) of the point to be added to the training data set \(\mathcal {D}\) is computed by minimizing the proxy of the one-step-ahead variance of the EFI at \(\textbf{x}_{targ}\) (see Eq. (8)), which is defined as:

$$\begin{aligned}{} & {} S(\textbf{x}_{targ},\tilde{\textbf{u}}) = Var\big ( I^{(t+1)}(\textbf{x}_{targ})\big ) \int _{\mathbb {R}^m} Var\big ( 1_{\{\textbf{G}^{(t+1)}(\textbf{x}_{targ},\textbf{u}) \le \textbf{0} \}} \big ) \rho _{\textbf{U}}(\textbf{u}) d\textbf{u},\\{} & {} \quad =~Var(\big (z_{\min }^{\text{ feas }} - Z^{(t+1)}(\textbf{x}_{targ})\big )^+)\int _{\mathbb {R}^m} Var\big ( 1_{\{\textbf{G}^{(t+1)}(\textbf{x}_{targ},\textbf{u}) \le \textbf{0} \}} \big ) \rho _{\textbf{U}}(\textbf{u}) d\textbf{u}. \nonumber \end{aligned}$$
(19)

In this appendix, we recall some details about the computation of the first term in the previous equation. An expression of the improvement variance \(Var\left( I^{(t+1)}(\textbf{x}_{targ})\right) \) which bears some similarities to the one of expected improvement is given in El Amri et al. (2021) and can be expressed in terms of probability and density functions of Gaussian distributions:

$$\begin{aligned} EI^{(s)}(\textbf{x})= & {} \big (z_{\min }^\text {feas} - m_Z^{(s)}(\textbf{x}) \big ) \phi \bigg (\frac{z_{\min }^\text {feas} - m_Z^{(s)}(\textbf{x})}{\sigma _Z^{(s)}(\textbf{x})}\bigg ) + \sigma _Z^{(s)}(\textbf{x}) \Phi \bigg (\frac{z_{\min }^\text {feas} - m_Z^{(s)}(\textbf{x})}{\sigma _Z^{(s)}(\textbf{x})}\bigg ),\\= & {} \psi _{EI}(m_Z^{(s)}(\textbf{x}), \sigma _Z^{(s)}(\textbf{x})). \\ Var\left( I^{(s)}(\textbf{x})\right)= & {} EI^{(s)}(\textbf{x}) \big (z_{\min }^\text {feas} - m_Z^{(s)}(\textbf{x}) - EI^{(s)}(\textbf{x})\big ) \\{} & {} + (\sigma _Z^{(s)})^2(\textbf{x}) \Phi \bigg (\frac{z_{\min }^\text {feas} - m_Z^{(s)}(\textbf{x})}{\sigma _Z^{(s)}(\textbf{x})}\bigg ),\\= & {} \psi _{VI}(m_Z^{(s)}(\textbf{x}), \sigma _Z^{(s)}(\textbf{x})). \end{aligned}$$

At step \(t+1\), the training data set \(\mathcal {D}^{(t)}\) is enriched by \((\tilde{\textbf{x}}, \tilde{\textbf{u}})\) on which the output \(f(\tilde{\textbf{x}},\tilde{\textbf{u}})\) is unknown and represented by \(F(\tilde{\textbf{x}},\tilde{\textbf{u}})\).

As a consequence, \(Var\left( I^{(t+1)}(\textbf{x}_{targ})\right) \) cannot directly be computed because of the randomness of \(m_Z^{(t+1)}(\textbf{x}_{targ})\). Indeed, \(m_Z^{(t+1)}(\textbf{x}_{targ})\) is equal to

\(E\left( Z(\textbf{x}) \vert F(\mathcal {D}^{(t)})=f^{(t)}, F(\tilde{\textbf{x}}, \tilde{\textbf{u}}) \right) \) and follows:

$$\begin{aligned} m_Z^{(t+1)}(\textbf{x}_{targ}) \sim \mathcal {N}\Bigg ( m_Z^{(t)}(\textbf{x}_{targ}) , \bigg (\frac{\int _{\mathbb {R}^m} k_{*F}^{(t)}(\textbf{x}_{targ},\textbf{u};\tilde{\textbf{x}},\tilde{\textbf{u}}) \rho _\textbf{U}(\textbf{u}) d\textbf{u} }{\sqrt{k_{*F}^{(t)}(\tilde{\textbf{x}},\tilde{\textbf{u}};\tilde{\textbf{x}},\tilde{\textbf{u}})}}\bigg )^2 \Bigg ). \end{aligned}$$

We can note that \(\sigma _Z^{(t+1)}(\textbf{x}_{targ})\) is not random as it depends only on the location \((\tilde{\textbf{x}},\tilde{\textbf{u}})\) and not on the function evaluation at this point. By applying the law of total variance, it can be shown that:

$$\begin{aligned} Var \left( I^{t+1}(\textbf{x}_{targ})\right)= & {} \mathbb E\left[ Var\left( \big (z_{\min }^{\text {feas}} - Z(\textbf{x}_{targ})\big )^+ \vert m_Z^{(t+1)}(\textbf{x}_{targ}) \right) \right] \\{} & {} +Var\left[ \mathbb E\left( \big (z_{\min }^{\text {feas}} - Z(\textbf{x}_{targ})\big )^+ \vert m_Z^{(t+1)}(\textbf{x}_{targ}) \right) \right] . \\= & {} \mathbb E\left[ \psi _{VI}(m_Z^{(t+1)}(\textbf{x}), \sigma _Z^{(t+1)}(\textbf{x})) \right] \\{} & {} +Var\left[ \psi _{EI}(m_Z^{(t+1)}(\textbf{x}), \sigma _Z^{(t+1)}(\textbf{x})) \right] . \\ \end{aligned}$$

This calculation is performed numerically using samples of \(m_Z^{(t+1)}(\textbf{x}_{targ}) \). For the sake of clarity, the reader is referred to the previous work (El Amri et al. 2021) for details regarding the implementation.

Appendix E: Optimization bounds of the industrial test case

The bounds of the optimization problem detailed in Sect. 6.4 are provided in the table below.

 

\(x_1\)

\(x_2\)

\(x_3\)

\(x_4\)

\(x_5\)

\(x_6\)

\(x_7\)

\(x_8\)

\(x_9\)

\(x_{10}\)

Lower bound

0.05

0.05

0.05

0.05

0.09

0.09

0.09

0.09

0.45

0.45

Upper bound

0.06

0.06

0.06

0.06

0.11

0.11

0.11

0.11

0.55

0.55

 

\(x_{11}\)

\(x_{12}\)

\(x_{13}\)

\(x_{14}\)

\(x_{15}\)

\(x_{16}\)

\(x_{17}\)

\(x_{18}\)

\(x_{19}\)

\(x_{20}\)

Lower bound

0.45

0.45

30

39

47

56

0.2

0.2

0.2

0.2

Upper bound

0.55

0.55

42

51

59

68

8.0

8.0

8.0

8.0

 

\(u_1\)

\(u_2\)

\(u_3\)

\(u_4\)

\(u_5\)

\(u_6\)

\(u_7\)

   

Lower bound

0.05

0.0000004

0.98

-5.763

-0.5

19.78

16845

   

Upper bound

0.80

0.0000012

1.02

5.763

0.5

20.59

17532

   
figure c

REF algorithm

figure d

SMCS algorithm

figure e

MMCU algorithm

figure f

MMCS algorithm

Appendix F: Flow charts of the algorithms

This appendix contains the flowcharts of the four algorithms which are implemented and compared in the body of the article. The REF, SMCS, MMCU and MMCS methods are described in Algorithms 2, 3, 4 and 5, respectively. REF stands for reference and was initially proposed in El Amri et al. (2021). It is a method where the GPs of the constraints and objective function are independent, and the same pair \((\textbf{x}^{t}, \textbf{u}^{t})\) is added to every GP at each iteration. SMCS, which stands for Single Models of the constraints and Constraint Selection, has independent GPs, like the REF algorithm, but only one constraint is updated at each iteration. The random parameters of the objective function and the selected constraint, \(\textbf{u}_f\) and \(\textbf{u}_g\), are different. MMCU means Multiple Model of the constraints and Common \(\textbf{u}\). The MMCU algorithm has a joint model of all the constraints and the same iterate \((\textbf{x}^{t}, \textbf{u}^{t})\) enriches all GPs. Finally, MMCS is the acronym of Multiple Model of the constraints and Constraint Selection. The MMCS algorithm relies on a joint model of the constraints and identifies at each iteration a single constraint and the associated random sample, \(\textbf{u}_g\), to carry out the next evaluation.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pelamatti, J., Le Riche, R., Helbert, C. et al. Coupling and selecting constraints in Bayesian optimization under uncertainties. Optim Eng 25, 373–412 (2024). https://doi.org/10.1007/s11081-023-09807-x

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11081-023-09807-x

Keywords

Navigation