Skip to main content
Log in

An efficient decoupled method for time-variant reliability-based design optimization

  • Research Paper
  • Published:
Structural and Multidisciplinary Optimization Aims and scope Submit manuscript

Abstract

Time-variant reliability-based design optimization (tRBDO) can rationally consider the time-variant uncertainties in engineering structures and find the optimal design that can keep reliable throughout its whole life cycle. However, solving the tRBDO involves a nested double-loop procedure and requires excessive computational cost. In this paper, a novel decoupled method called sequential approximate time-variant reliability analysis and optimization (SATO) is proposed to improve the efficiency of tRBDO. First, a two-step method is proposed to transform the original tRBDO problem into an equivalent deterministic optimization problem according to the results of time-variant reliability analysis (TRA). Second, a novel approximate TRA (ATRA) method based the least-square method is proposed to reduce the computational cost of TRA. Finally, the proposed SATO method decouples the original double-loop procedure in tRBDO into a sequential process of ATRA and deterministic optimization. Test results of a complicated welded beam problem verify that the proposed method can achieve similar accuracy and much higher efficiency than the compared methods. A rocket inter-stage structure problem demonstrates the capability of the proposed method in practical engineering applications.

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

Similar content being viewed by others

References

Download references

Acknowledgements

The present work was partially supported by the National Natural Science Foundation of China (Grant No. 11502209), the National Defense Fundamental Research Funds of China (Grant No. JCKY2016204B102, JCKY2016208C001).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chunlin Gong.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Replication of results

The core source code of the proposed method and the detailed results are provided in the supplementary material.

Additional information

Responsible Editor: Tae Hee Lee

Publisher's note

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

Supplementary information

ESM 1

(TAR 3094 kb)

Appendices

Appendix 1

1.1 AMPPT method for TRA and identifying critical time instants

The AMPPT method is based on the concept of MPP trajectory (Zhang et al. 2021). It can not only accurately calculate the time-variant reliability, but also identify the critical time instants within the time interval [0, T] via an adaptive sampling process.

For an arbitrary time instant ta ∈ [0, T] , denote the MPP of the instantaneous performance function g(d, X, Y(ta), ta) as uMPP(ta). When ta varies from 0 to T, the MPP uMPP(ta) will move from uMPP(0) to uMPP(T). If we connect all these MPPs, a curve uMPP(t)(t ∈ [0, T]) in u-space can be obtained, which is defined as the MPP trajectory. Figure 14 shows a schematic diagram of the MPP trajectory, where the solid curves represent the limit-state boundaries at the critical time instants.

Fig. 14
figure 14

Schematic diagram of the MPP trajectory and critical time instants

For a given TRA problem, the AMPPT method first approximates its MPP trajectory with the adaptive Kriging model \( {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t) \), and then calculates the time-variant reliability based on \( {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t) \), The solution process of AMPPT consists of three steps, which are briefly described as follows.

First, discretize [0, T] into Ninit equidistant time instants, and perform MPP searches at these time instants to obtain the initial time-MPP samples. {(ti, uMPP(ti))| i = 1, 2, .., Ninit} Then, build a rough Kriging model \( {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t) \) with these samples. Afterwards, perform an adaptive sampling process to iteratively identify the critical time instants t, at which the Kriging model \( {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t) \) has the largest prediction variance,

$$ {t}^{\ast }=\arg {\displaystyle \begin{array}{c}\max \\ {}t\in \left[{t}_s,{t}_e\right]\end{array}}\frac{\sum_{j=1}^{n+m}{\sigma}_j^2(t)}{n+m} $$
(34)

where \( {\sigma}_j^2(t)\left(j=1,2,\cdots n+m\right) \) is the prediction variance of the j-th component of \( {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t) \). Then, perform MPP-search at t to obtain a new sample (t, uMPP(t)) , and update \( {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t) \) accordingly. Repeat this process until the Kriging model \( {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t) \) is accurate enough.

Second, according to \( {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t) \), transform the response of the time-variant performance function g(d, X, Y, (t), t) into an equivalent Gaussian process H(t) . The mean, standard deviation, and autocorrelation coefficient functions of H(t) are derived as:

$$ {\displaystyle \begin{array}{c}\mu (t)=-\beta (t)=-\left\Vert {\hat{\mathbf{u}}}_{\mathrm{MPP}}(t)\right\Vert \\ {}\begin{array}{c}\sigma (t)=1\\ {}\rho \left({t}_1,{t}_2\right)=\frac{\mathrm{Cov}\left(H\left({t}_1\right),H\left({t}_2\right)\right)}{\sigma \left({t}_1\right)\sigma \left({t}_2\right)}\\ {}=\mathrm{Cov}\left(-\beta \left({t}_2\right)+{\left({\boldsymbol{a}}_{\mathrm{v}}\left({t}_1\right)\right)}^T\mathbf{V}+{\left({\boldsymbol{a}}_{\mathrm{w}}\left({t}_1\right)\right)}^T{\mathbf{W}}_{1,}-\beta \left({t}_2\right)+{\boldsymbol{a}}_{\mathrm{v}}\left({t}_2\right)\left){}^T\mathbf{V}+{\boldsymbol{a}}_{\mathrm{w}}\left({t}_1\right)\right){}^T{\mathbf{W}}_2\right)\end{array}\\ {}\begin{array}{c}=\mathrm{Cov}{\left(\Big({\boldsymbol{a}}_{\mathrm{v}}\left({t}_1\right)\right)}^T\mathbf{V}+{\left({\boldsymbol{a}}_{\mathrm{w}}\left({t}_1\right)\right)}^T{\mathbf{W}}_1,{\left({\boldsymbol{a}}_{\mathrm{v}}\left({t}_2\right)\right)}^T\mathbf{V}+{\left({\boldsymbol{a}}_{\mathrm{w}}\left({t}_2\right)\right)}^T{\mathrm{W}}_2\Big)\\ {}=\mathrm{Cov}\left(\Big({\boldsymbol{a}}_{\mathrm{v}}{\left({t}_1\right)}^T\mathbf{V},{\left({\boldsymbol{a}}_{\mathrm{v}}\Big({t}_2\right)}^T\mathbf{V}\right)+\mathrm{Cov}\left({\left({\boldsymbol{a}}_{\mathrm{w}}\left({t}_1\right)\right)}^T{\mathrm{W}}_1,{\left({\boldsymbol{a}}_w\left({t}_2\right)\right)}^T{\mathrm{W}}_2\right)\\ {}={\left({\boldsymbol{a}}_{\mathrm{v}}\left({t}_1\right)\right)}^T{\boldsymbol{a}}_{\mathrm{v}}\left({t}_2\right)+{\left({\boldsymbol{a}}_{\mathrm{w}}\left({t}_1\right)\right)}^T\mathbf{C}\left({t}_{1,}{t}_2\right){\boldsymbol{a}}_w\left({t}_2\right)\end{array}\end{array}} $$
(35)

where C(t1, t2) is a nY × nY correlation coefficient matrix. V and W are independent standard normal random variables transformed from X and Y(t), respectively. av(t) and aW(t) are calculated by

$$ {\displaystyle \begin{array}{c}{\boldsymbol{\upalpha}}_{\mathrm{V}}(t)={\hat{\mathbf{u}}}_{\mathrm{MPP},\mathrm{V}}(t)/\beta (t)\\ {}{\boldsymbol{\upalpha}}_{\mathrm{W}}(t)={\hat{\mathbf{u}}}_{\mathrm{MPP},\mathrm{W}}(t)/\beta (t)\\ {}{\hat{\mathbf{u}}}_{\mathrm{MPP}}(t)=\left[{\hat{\mathbf{u}}}_{\mathrm{MPP},\mathrm{V}}(t),{\hat{\mathbf{u}}}_{\mathrm{MPP},\mathrm{W}}(t)\right]\end{array}} $$
(36)

Finally, calculate the time-variant reliability based on spectral decomposition (Sudret and Der Kiureghian 2002) and Monte Carlo Simulation (MCS). Discretize the time interval [0, T] into s equidistant time instants ti(i = 1, 2, …, s), and construct a covariance matrix as

$$ \boldsymbol{\Sigma} ={\left[\begin{array}{cccc}\mathrm{Cov}\left({t}_1,{t}_1\right)& \mathrm{Cov}\left({t}_2,{t}_1\right)& \cdots & \mathrm{Cov}\left({t}_s,{t}_1\right)\\ {}\mathrm{Cov}\left({t}_1,{t}_2\right)& \mathrm{Cov}\left({t}_2,{t}_2\right)& \cdots & \mathrm{Cov}\left({t}_s,{t}_2\right)\\ {}\vdots & \vdots & \ddots & \vdots \\ {}\mathrm{Cov}\left({t}_1,{t}_s\right)& \mathrm{Cov}\left({t}_2,{t}_s\right)& \cdots & \mathrm{Cov}\left({t}_s,{t}_s\right)\end{array}\right]}_{s\times s} $$
(37)

where, Cov(ti, tj) = σ(ti)σ(tj)ρ(ti, tj), for i, j = 1, 2, ⋯, s. Then, H(t) can be decomposed as (Sudret and Der Kiureghian 2002)

$$ H(t)\approx \mu (t)+\sigma (t)\sum \limits_{k=1}^p\frac{\xi_k}{\sqrt{\lambda_k}}{\mathbf{Q}}_k^T\boldsymbol{\uprho} (t) $$
(38)

where ξk(k = 1, 2, …, p) are independent standard normal random variables;λk and Qk are the eigenvalues and eigenvectors of ∑, respectively; ρ(t) = [Cov(t1, t), Cov(t2, t), .., Cov(tp, t)]T is a covariance vector.

Use (38) to generate NMCS samples \( {H}^{(j)}=\left[{h}_1^{(j)},{h}_2^{(j)},\cdots, {h}_s^{(j)}\right] \), (j = 1, 2, ⋯, NMCS) of H(t), and the time-variant reliability index βcur can be estimated by

$$ {\beta}_{\mathrm{cur}}={\Phi}^{-1}\left(P(T)\right)={\Phi}^{-1}\left(\frac{\sum \limits_{j=1}^{N_{\mathrm{MCS}}}I\left({H}^{(i)}\right)}{N_{\mathrm{MCS}}}\right) $$
(39)

where I(H(j)) is an indicator function. If \( {\max}_{i=l}^s\left({h}_i^{(j)}\right)<0 \) , I(H(j)) = 1; otherwise, I(H(j)) = 0.

Appendix 2

1.1 Kriging model

In both the AMPPT method described above and the ATRA method proposed in Section 3.2, the Kriging model (Lophaven et al. 2002; Gano et al. 2006) is selected to approximate the MPP trajectory due to its advantage in providing the prediction variance and its successful applications in field of reliability analysis (Hawchar et al. 2018; Zhang et al. 2019; Li et al. 2020).

The Kriging model approximates the jth(j = 1, 2, …, n + m) component μMPP, j(t) of the MPP trajectory uMPP(t) (see Fig. 14) as

$$ {\hat{u}}_{\mathrm{MPP},j}(t)=f(t)+s(t) $$
(40)

where f(t) is a polynomial term of t and s(t) is a Gaussian process with zero mean and covariance Cov[s(tp), s(tq)] In this paper,f(t) is treated as a constant μ. The covariance Cov[s(tp), s(tq)] of s(t) is calculated by

$$ \mathrm{Cov}\left[s\left({t}_p\right),s\left({t}_q\right)\right]={\sigma}^2R\left({t}_p,{t}_q\right)={\sigma}^2\exp {\left[-\theta \left({t}_p-{t}_q\right)\right]}^2 $$
(41)

where σ2 is the variance of s(t), R(tp, tq) is the correlation coefficient, and θ is a parameter that can be determined by the maximum likelihood estimation (Giunta and Watson 1998).

Assume the number of “time-MPP” pairs {(ti, uMPP(ti))⌊i = 1, 2, ⋯n} is n. Denote y = {uMPP, j(ti)}. The natural logarithm of the likelihood function is defined as

$$ L\left(\theta |\mathbf{y}\right)=-\frac{1}{2}\left[n\ln \left(2\pi \right)+n\ln {\sigma}^2+\ln \left|\mathbf{R}\right|+\frac{{\left(\mathbf{y}-\mathbf{A}\mu \right)}^T{\mathbf{R}}^{-1}\left(\mathbf{y}-\mathbf{A}\mu \right)}{2{\sigma}^2}\right] $$
(42)

where R = [R(tp, tq)]n × n is a n × n correlation matrix and A is a n × 1 unit vector. By setting the derivatives of (42) with respect to μ and σ2 to zero, μ and σ2can be estimated as

$$ \hat{\mu}=\frac{{\mathbf{A}}^T{\mathbf{R}}^{-1}\mathbf{y}}{{\mathrm{A}}^T{\mathbf{R}}^{-1}\mathbf{y}}\ {\hat{\sigma}}^2=\frac{{\left(\boldsymbol{y}-\mathbf{A}\mu \right)}^T{\mathbf{R}}^{-1}\left(\mathbf{y}-\mathbf{A}\mu \right)}{n} $$
(43)

Substituting (43) into (42),θ can be determined by maximizing the likelihood function

$$ \theta ={\displaystyle \begin{array}{c}\mathrm{argmax}\\ {}\theta \end{array}}\left(-\frac{nl\mathrm{n}{\hat{\sigma}}^2+\ln \mid \mathbf{R}\mid }{2}\right) $$
(44)

Once all hyper parameters are obtained, the Kriging model \( {\hat{u}}_{\mathrm{MPP},j}(t) \) can be used to predict the jth (j = 1, 2, …, n + m) component of the MPP at an arbitrary time instant \( \overset{\acute{\mkern6mu}}{t} \):

$$ {\hat{u}}_{\mathrm{MPP},j}\left(\overset{\acute{\mkern6mu}}{t}\right)=\hat{\mu}+{\mathbf{r}}^T{\mathbf{R}}^{-1}\left(\mathbf{y}-\mathbf{A}\hat{\mu}\right) $$
(45)

where r is a correlation vector defined by \( \mathbf{r}={\left[R\left(\overset{\acute{\mkern6mu}}{t},{t}_1\right),R\left(\overset{\acute{\mkern6mu}}{t},{t}_2\right),\dots, R\left(\overset{\acute{\mkern6mu}}{t},{t}_n\right)\right]}^T \). The variance of the prediction in (45) is given by

$$ {\sigma}_j^2\left(\overset{\acute{\mkern6mu}}{t}\right)={\overset{\acute{\mkern6mu}}{\sigma}}^2\left[1-{\boldsymbol{r}}^T{\mathbf{R}}^{-1}\mathbf{r}+\frac{{\left(1-{\mathbf{A}}^T{\mathbf{R}}^{-1}\mathbf{r}\right)}^2}{{\mathbf{A}}^T{\mathbf{R}}^{-1}\mathbf{A}}\right] $$
(46)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, Y., Gong, C., Li, C. et al. An efficient decoupled method for time-variant reliability-based design optimization. Struct Multidisc Optim 64, 2449–2464 (2021). https://doi.org/10.1007/s00158-021-02999-9

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-021-02999-9

Keywords

Navigation