Skip to main content
Log in

Monte carlo within simulated annealing for integral constrained optimizations

  • Original Research
  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

For years, Value-at-Risk and Expected Shortfall have been well established measures of market risk and the Basel Committee on Banking Supervision recommends their use when controlling risk. But their computations might be intractable if we do not rely on simplifying assumptions, in particular on distributions of returns. One of the difficulties is linked to the need for Integral Constrained Optimizations. In this article, two new stochastic optimization-based Simulated Annealing algorithms are proposed for addressing problems associated with the use of statistical methods that rely on extremizing a non-necessarily differentiable criterion function, therefore facing the problem of the computation of a non-analytically reducible integral constraint. We first provide an illustrative example when maximizing an integral constrained likelihood for the stress-strength reliability that confirms the effectiveness of the algorithms. Our results indicate no clear difference in convergence, but we favor the use of the problem approximation strategy styled algorithm as it is less expensive in terms of computing time. Second, we run a classical financial problem such as portfolio optimization, showing the potential of our proposed methods in financial 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

Similar content being viewed by others

Notes

  1. Without loss of generality, we only consider a minimization problem with equality constraints, since a maximization problem can be transformed into a minimization problem by negating the objective function while inequality constraints, say, \(g_{i}(x)\le 0\) can likewise be transformed into an equivalent equality constraint of the form \((g_{i}(x))^{+}=\max \{0,g_{i}(x)\}=0\).

  2. We denote the EE distribution with \(EE(\alpha ,\beta )\) where \(\alpha \) and \(\beta \) are, respectively, the shape and scale parameters. The respective distribution function, \(F_{X}(\cdot )\), may be written following the definition by Gupta and Kundu (2001) as:

    $$\begin{aligned} \begin{aligned} F_{X}(x;\alpha ,\beta )&= \left( 1-\exp (-\beta x) \right) ^{\alpha },~~\alpha>0, \beta>0, x>0. \end{aligned} \end{aligned}$$
  3. Smith et al. (2015) consider the change of variable \(z=\beta _{1}x\) to obtain:

    $$\begin{aligned} \mathcal {R} = \alpha _{1}\int _{0}^{\infty } \exp (-\beta _{1}z) \left( 1- \exp {(-\beta _{1}z)} \right) ^{\alpha _{1}-1}\left( 1- \exp {(-(\beta _{2}/\beta _{1})z)} \right) ^{\alpha _{2}}dz, \end{aligned}$$

    while Nadarajah (2011) poses the transformation \(z=\exp (-\beta _{1} x)\) and obtains:

    $$\begin{aligned} \mathcal {R} = \alpha _{1}\int _{0}^{1} \left( 1- z \right) ^{\alpha _{1}-1}\left( 1- z^{\beta _{2}/\beta _{1}} \right) ^{\alpha _{2}}dz. \end{aligned}$$
  4. All results are obtained using MatLab (version R2019a) on a laptop with a Intel(R) Core(TM) i7-5500U CPU @ 2.40GHz and 8GB RAM. The codes used in this article are available upon request from the corresponding author.

  5. In our robustness checks we vary the value of main parameters of the algorithms, namely: a, N and \(T^{*}\), obtaining similar results in terms of efficiency and computational time (Cf. Appendix B).

  6. We thank the first referee for noticing that the choice of the randomly chosen assets has no importance (since it can be generalized for all triplets of assets, and, more generally, to all sets of assets).

  7. We follow Boucher et al. (2013) to build up our database: the “Equity” asset class is represented by a composite index of 95% of the MSCI Europe and 5% of the MSCI World; the “Bonds” asset class comes from the x-trackers II iBoxx Sovereigns Eurozone AAA UCITS 1C index (denoted as XBAT index); the “Money Market” is represented by the Euro Overnight Index Average (EONIA).

References

Download references

Acknowledgements

We thank Jean-Luc Prigent for positive comments and encouragement on the preliminary draft of this article, as well as participants to the FEM2021 conference (Paris, on-line, June 2021), as well as the Editor in charge and two anonymous referees who significantly contributed to the improvement of the article. Roberto Casarin and Anthony Osuntuyi gratefully acknowledge the funding received from the EURO Horizon 2020 project on Energy Efficient Mortgages Action Plan (EeMAP), grant agreement No 746205 and the VERA project on economic and financial cycles under the grant number No. 479/2018 Prot. 33287 -VII / 16 of 08.06.2018. The usual disclaimers apply.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Anthony Osuntuyi.

Additional information

Publisher's Note

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

Appendices

Appendix A proof to theorem 1: ergodicity of MCWMSA

The proof for Theorem 1 follows standard arguments for the convergence of Inhomogeneous Markov Chains (e.g., Isaacson and Madsen (1976) and Seneta (1981) for related results). More precisely, we first form a lower bound on the probability of reaching any solution from any local minimum, and then show that this bound does not approach zero too quickly. Let G be the generation probability that satisfies 19.

  1. a.

    Let:

    $$\begin{aligned} \displaystyle \varDelta _{G} = \min _{\begin{array}{c} (x',\lambda ')\in \mathcal {N}({x,\lambda }),\\ {x,\lambda }\in \mathcal {S} \end{array}} G(x,\lambda ;x',\lambda ') \end{aligned}$$

    and:

    $$\begin{aligned} \displaystyle \varDelta _{L^{M}} = 2\max _{\begin{array}{c} (x',\lambda ')\in \mathcal {N}(x,\lambda ),\\ (x'\lambda ')\in \mathcal {S},y,y'\in \mathcal {Y}^M \end{array}}\{|L^{M}(x',\lambda ') -L^{M}(x,\lambda )|\}, \end{aligned}$$

    be, respectively, the value of the smallest one-step solution generation probabilities and the size of the largest potential jump in the value of the auxiliary penalty function between any pair of trial solutions. \(\mathcal {S}=\mathcal {X}\times \mathbb {R}_{+}^{r}\) denotes the search space while \(\mathcal {N}(x,\lambda )\) represents the neighborhood of \((x,\lambda )\). If \((x,\lambda )\ne (x',\lambda ')\), then for all \((x,\lambda )\in \mathcal {S}\) and \((x',\lambda ')\in \mathcal {N}(x,\lambda )\), we have:

    $$\begin{aligned} \begin{aligned}&P_{MCWMSA}(x,\lambda ;x',\lambda ') = \int _{\mathcal {Y}^{M}}\int _{\mathcal {Y}^{M}} G(x,\lambda ;x',\lambda ')A^{M}_{T_{k}}(x,\lambda ;x',\lambda ')q^{M}_{x}(y)q^{M}_{x'}(y')dydy' \\&\quad = G(x,\lambda ;x',\lambda ')\int _{\mathcal {Y}^{M}}\int _{\mathcal {Y}^{M}} A^{M}_{T_{k}}(x,\lambda ;x',\lambda ')q^{M}_{x}(y)q^{M}_{x'}(y')dydy' \\&\quad = {\left\{ \begin{array}{ll} G(x,\lambda ;x',\lambda ')\int _{\mathcal {Y}^{M}}\int _{\mathcal {Y}^{M}} \min \left\{ \exp \left( -\frac{(L^{M}(x',\lambda ')-L^{M}(x,\lambda ))}{T}\right) ,1\right\} q^{M}_{x}(y)q^{M}_{x'}(y')dydy', &{} {\text {if }u=1 }\\ G(x,\lambda ;x',\lambda ')\int _{\mathcal {Y}^{M}}\int _{\mathcal {Y}^{M}} \min \left\{ \exp \left( -\frac{(L^{M}(x,\lambda )-L^{M}(x',\lambda '))}{T}\right) ,1\right\} q^{M}_{x}(y)q^{M}_{x'}(y')dydy' , &{} {\text {if }u=0 } \end{array}\right. }\\&\quad \ge G(x,\lambda ;x',\lambda ')\exp \left( -\frac{\varDelta _{L^{M}}}{T_{k}}\right) \int _{\mathcal {Y}^{M}}\int _{\mathcal {Y}^{M}} q^{M}_{x}(y)q^{M}_{x'}(y')dydy'\\ \\&\quad \ge \varDelta _{G}\exp \left( -\frac{\varDelta _{L^{M}}}{T_{k}}\right) . \end{aligned} \end{aligned}$$
    (A.1)
  2. b.

    For any \(\lambda \), let:

    $$\begin{aligned} \displaystyle \varDelta _{min}^{M} = \min _{\begin{array}{c} (x',\lambda ')\in \mathcal {N}(x,\lambda ),\\ A^{M}_{T_{k}}<1, y,y'\in \mathcal {Y}^{M} \end{array}} \left( L^{M}(x',\lambda ') - L^{M}(x,\lambda ) \right) \ge 0, \end{aligned}$$

    be the size of the smallest positive jump in the value of the objective function between any solution pair. If \((x',\lambda ')=(x,\lambda )\), we have:

    $$\begin{aligned}&P_{MCWMSA}(x,\lambda ;x',\lambda ') \nonumber \\&\quad = 1-\int _{\mathcal {N}(x,\lambda )}\int _{\mathcal {Y}^{M}}\int _{\mathcal {Y}^{M}} A_{T_{k}}^{M}(x,\lambda ;x',\lambda ') q^{M}_{x}(y)q^{M}_{x'}(y')dydy'G(x,\lambda ;dx',d\lambda ') \nonumber \\&\quad = \int _{\begin{array}{c} \mathcal {N}(x,\lambda ),A^{M}_{T_{k}}<1 \end{array}}\int _{\mathcal {Y}^{M}}\int _{\mathcal {Y}^{M}} \left( 1- A_{T_{k}}^{M}(x,\lambda ;x',\lambda ')\right) q^{M}_{x}(y)q^{M}_{x'}(y')dydy'G(x,\lambda ;dx',d\lambda ') \nonumber \\&\quad \ge \int _{\begin{array}{c} \mathcal {N}(x,\lambda ),A^{M}_{T_{k}}<1 \end{array}}\int _{\mathcal {Y}^{M}}\int _{\mathcal {Y}^{M}} \varDelta _{G}(1 - \exp (-\varDelta _{min}^{M}/T_{k}))q^{M}_{x}(y)q^{M}_{x'}(y')dydy'dx'd\lambda '. \end{aligned}$$
    (A.2)

    Since, \(T_{k}\) is a decreasing sequence, it is always possible to find a \(k_{0}>0\) so that for all \(k\ge k_{0}\), \(\displaystyle (1-\exp (-\varDelta _{min}^{M}/T_{k}))\ge \exp (-\varDelta _{L^{M}}/T_{k})\).

    Thus:

    $$\begin{aligned} P_{MCWMSA}(x,\lambda ;x,\lambda )\ge \exp (-\varDelta _{L^{M}}/T_{k}). \end{aligned}$$
  3. c.

    Based on the result obtained above, for all \((x,\lambda ),(x',\lambda ')\in \mathcal {S}\) and \(k\ge k_{0}\), the N-step transition probability from \((x,\lambda )=(x_{0},\lambda _{0})\) to \((x',\lambda ')=(x_{N},\lambda _{N})\) satisfies the following:

    $$\begin{aligned} P_{MCWMSA}^{N}(x,\lambda ;x',\lambda ')= \prod _{i=0}^{N-1}P_{MCWMSA}(x_{i},\lambda _{i};x_{i+1},\lambda _{i+1}) \ge \left( \varDelta _{G}\exp (-\varDelta _{L^{M}}/T_{k}) \right) ^{N}. \end{aligned}$$

    Let \(\displaystyle {\tau _{1}\left( P\right) }\) denote the coefficient of ergodicity for the transition probability matrix P, defined by:

    $$\begin{aligned} \begin{aligned}&\tau _{1}\left( P_{MCWMSA}^{N}\right) = 1 - \min _{(x,\lambda ),(x',\lambda ')\in \mathcal {S}}\sum _{(x'',\lambda '')\in \mathcal {S}} d(x,\lambda ,x',\lambda ',x'',\lambda '') \end{aligned} \end{aligned}$$

    where \(d(x,\lambda ,x',\lambda ',x'',\lambda '')\!=\!\min \left\{ P_{MCWMSA}^{N}(x,\lambda ; x'',\lambda ''), P_{MCWMSA}^{N}(x',\lambda '; x'',\lambda '' ) \right\} \) An Inhomogeneous Markov Chain is weakly ergodic if and only if there is a strictly increasing sequence of positive numbers so that:

    $$\begin{aligned} \sum _{k=0}\left( 1-\tau _{1}(P_{MCWMSA}^{N})\right) = \infty , \end{aligned}$$
    (A.3)

    using the bound on \(P_{MCWMSA}^{N}(x,\lambda ;x',\lambda ')\), we find that:

    $$\begin{aligned}&1- \tau _{1}(P_{MCWMSA}^{N})=\min _{(x,\lambda ),(x',\lambda ')\in \mathcal {S}}\sum _{(x'',\lambda '')\in \mathcal {S}}d(x,\lambda ,x',\lambda ',x'',\lambda '') \nonumber \\&\quad \ge \min _{(x,\lambda ),(x',\lambda ')\in \mathcal {S}}\min _{(x'',\lambda '')\in \mathcal {S}}\left\{ P_{MCWMSA}^{N}(x,\lambda ; x'',\lambda ''), P_{MCWMSA}^{N}(x',\lambda '; x'',\lambda '' ) \right\} \nonumber \\&\quad \ge \left( \varDelta _{G}\exp (-\varDelta _{L^{M}}/T_{k}) \right) ^{N} = \varDelta _{G}^{N}\exp (-N\varDelta _{L^{M}}/T_{k}). \end{aligned}$$
    (A.4)

    Substituting \(\displaystyle {T_{k}\ge \frac{N\varDelta _{L^{M}}}{\log {(k+1)}}, ~~k\ge 0 }\) in Eq. A.3 gives:

    $$\begin{aligned} \begin{aligned} \sum _{k=0}\left( 1-\tau _{1}(P_{MCWMSA}^{N})\right)&\ge \sum _{k=k_{0}}\varDelta _{G}^{N}\exp (-N\varDelta _{L^{M}}/T_{k})\\&\ge \varDelta _{G}^{N}\sum _{k=k_{0}}\dfrac{1}{k+1} = \infty . \end{aligned} \end{aligned}$$
    (A.5)

    This implies that the Markov Chain, \((x,\lambda ,x',\lambda ')\), is weakly ergodic.

  4. d.

    Furthermore, since the acceptance probability function \(A_{T_{K}}^{M}\) is an exponential rational function in \(1/T_{k}\) and belongs to the closed class of asymptotically monotone functions (Anily and Federgruen (1987), Proposition 1) then the sequence of \((x_{n},\lambda _{n})\) converges in distribution (Anily and Federgruen (1987), Corollary 1). This implies that the Markov Chain is strongly ergodic.\(\square \)

Appendix B robustness checks

This section provides the robustness checks of the results in Tables 2 and 3 with respect the main parameters of the algorithms, namely a, N and \(T^{*}\) (boldface values), showing similar results in terms of efficiency, computational time and GIMHSA dominance.

Table 7 Parameter Estimates for Dataset#1, comparing CSA of Smith et al. (2015), MCWMSA and GIMHSA Algorithms when solving the Integral Constrained Log-likelihood in Eq. 23 using Dataset#1 as reported in Table 1, with \(a=\mathbf {0.80}\), \(N=\mathbf {250}\), \(T_{0}=10\) and \(T^{*}= 10^{-6}\)
Table 8 Reported Characteristics of the Methods in terms of Performances of the Algorithms on the Dataset#1 of Smith et al. (2015) with \(a=\mathbf {0.80}\), \(N=\mathbf {250}\), \(T_{0}=10\) and \(T^{*}= 10^{-6}\)
Table 9 Parameter Estimates for Dataset#1, comparing CSA of Smith et al. (2015), MCWMSA and GIMHSA Algorithms when solving the Integral Constrained Log-likelihood in Eq. 23 using Dataset#1 as reported in Table 1, with \(a=0.95\), \(N=\mathbf {250}\), \(T_{0}=10\) and \(T^{*}= \mathbf {10^{-4}}\)
Table 10 Reported Characteristics of the Methods in terms of Performances of the Algorithms on the Dataset#1 of Smith et al. (2015) with \(a=0.95\), \(N=\mathbf {500}\), \(T_{0}=10\) and \(T^{*}= \mathbf {10^{-4}}\)
Table 11 Parameter Estimates for Dataset#1, comparing CSA of Smith et al. (2015), MCWMSA and GIMHSA Algorithms when solving the Integral Constrained Log-likelihood in Eq. 23 using Dataset#1 as reported in Table 1, with \(a=0.95\), \(N=\mathbf {250}\), \(T_{0}=10\) and \(T^{*}= \mathbf {10^{-4}}\)
Table 12 Reported Characteristics of the Methods in terms of Performances of the Algorithms on the Dataset#1 of Smith et al. (2015) for \(a=0.95\), \(N=\mathbf {250}\), \(T_{0}=10\) and \(T^{*}= \mathbf {10^{-4}}\)

Appendix C Pseudo-codes

figure a
figure b
figure c
figure d
figure e

Rights and permissions

Springer Nature or its licensor 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

Casarin, R., Maillet, B.B. & Osuntuyi, A. Monte carlo within simulated annealing for integral constrained optimizations. Ann Oper Res 334, 205–240 (2024). https://doi.org/10.1007/s10479-022-04994-9

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-022-04994-9

Keywords

Navigation