Skip to main content
Log in

KDE distributionally robust portfolio optimization with higher moment coherent risk

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

Abstract

In this paper, distributionally robust mean-HMCR (higher moment coherent risk) portfolio optimization model based on kernel density estimation (KDE) and \(\phi \)-divergence is proposed. In order to overcome the so-called “curse of dimensionality”, we consider the one-dimensional probability distribution of the portfolio return, rather than the joint probability distribution of the assets return vector. The two issues of “the distribution dependent on the decision variables” and “the metric-based distributional uncertainty set for the continuous distribution” are effectively addressed by using the finite dimensional KDE based probability distribution. Under the mild conditions of the kernel function and \(\phi \)-divergence function, the tractable reformulation of the corresponding distributionally robust optimization model is derived by Fenchel’s Duality Theorem. Moreover, the convergence of optimal value and solution set of the KDE mean-HMCR distributionally robust portfolio optimization problem to those of the corresponding stochastic optimization model with the real distribution is proved. We conduct some empirical tests with the rolling horizon approach and compare the performance of the optimal portfolio strategy obtained by the proposed model to other three strategies by four performance criteria and their cumulative wealth curves. Empirical test results show that the quality of the portfolio strategy obtained by the proposed model is better at most cases. We also conduct empirically sensitivity analysis of model parameters.

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

Similar content being viewed by others

References

  • Artzner, P., Delbaen, F., Eber, J. M., & Heath, D. (1999). Coherent measures of risk. Mathematical Finance, 9(3), 203–228.

    Article  Google Scholar 

  • Bayraksan, G., & Love, D. (2015) Data-driven stochastic programming using phi-divergences. TutOrials in Operations Research

  • Ben-Tal, A., & Nemirovski, A. (1998). Robust convex optimization. Mathematics of Operations Research, 23(4), 769–805.

    Article  Google Scholar 

  • Ben-Tal, A., El Ghaoui, L., & Nemirovski, A. (2009). Robust optimization, (Vol. 28). Princeton University Press.

  • Ben-Tal, A., Den Hertog, D., De Waegenaere, A., Melenberg, B., & Rennen, G. (2013). Robust solutions of optimization problems affected by uncertain probabilities. Management Science, 59(2), 341–357.

    Article  Google Scholar 

  • Ben-Tal, A., Den Hertog, D., & Vial, J. P. (2015). Deriving robust counterparts of nonlinear uncertain inequalities. Mathematical Programming, 149(1–2), 265–299.

    Article  Google Scholar 

  • Bertsimas, D., & Popescu, I. (2005). Optimal inequalities in probability theory: A convex optimization approach. SIAM Journal on Optimization, 15(3), 780–804.

    Article  Google Scholar 

  • Bertsimas, D., & Sim, M. (2004). The price of robustness. Operations Research, 52(1), 35–53.

    Article  Google Scholar 

  • Bertsimas, D., Gupta, V., & Kallus, N. (2018). Robust sample average approximation. Mathematical Programming, 171(1), 217–282.

    Article  Google Scholar 

  • Boyd, S., Kim, S., Vandenberghe, L., & Hassibi, A. (2007). A tutorial on geometric programming. Optimization and Engineering, 8(1), 67–127.

    Article  Google Scholar 

  • Calafiore, G. C. (2007). Ambiguous risk measures and optimal robust portfolios. SIAM Journal on Optimization, 18(3), 853–877.

    Article  Google Scholar 

  • Chekhlov, A., Uryasev, S., & Zabarankin, M. (2005). Drawdown measure in portfolio optimization. International Journal of Theoretical and Applied Finance, 8(1), 13–58.

    Article  Google Scholar 

  • Cui, X., Sun, X., Zhu, S., Jiang, R., & Li, D. (2018). Portfolio optimization with nonparametric value at risk: A block coordinate descent method. INFORMS Journal on Computing, 30(3), 454–471.

    Article  Google Scholar 

  • Delage, E., & Ye, Y. (2010). Distributionally robust optimization under moment uncertainty with application to data-driven problems. Operations Research, 58(3), 595–612.

    Article  Google Scholar 

  • DeMiguel, V., Garlappi, L., & Uppal, R. (2009). Optimal versus naive diversification: How inefficient is the 1/N portfolio strategy? Review of Financial Studies, 22(5), 1915–1953.

    Article  Google Scholar 

  • El Ghaoui, L., Oks, M., & Oustry, F. (2003). Worst-case value-at-risk and robust portfolio optimization: A conic programming approach. Operations Research, 51(4), 543–556.

    Article  Google Scholar 

  • Erdoğan, E., & Iyengar, G. (2006). Ambiguous chance constrained problems and robust optimization. Mathematical Programming, 107(1), 37–61.

    Article  Google Scholar 

  • Fabozzi, F. J., Huang, D., & Zhou, G. (2010). Robust portfolios: Contributions from operations research and finance. Annals of Operations Research, 176(1), 191–220.

    Article  Google Scholar 

  • Fishburn, P. C. (1977). Mean-risk analysis with risk associated with below-target returns. American Economic Review, 67(2), 116–126.

    Google Scholar 

  • Gao R, Kleywegt AJ (2016) Distributionally robust stochastic optimization with Wasserstein distance. arXiv:1604.02199v2.

  • Goh, J., & Sim, M. (2010). Distributionally robust optimization and its tractable approximations. Operations Research, 58(4), 902–917.

    Article  Google Scholar 

  • Grant, M., & Boyd, S. (2013) CVX: Matlab software for disciplined convex programming, version 2.0 beta. http://cvxr.com/cvx

  • Gupta, V. (2019). Near-optimal Bayesian ambiguity sets for distributionally robust optimization. Management Science, 65(9), 4242–4260.

    Article  Google Scholar 

  • Hanasusanto, GA., & Kuhn, D. (2013) Robust data-driven dynamic programming. In: Burges CJC, Bottou L, Welling M, Ghahramani Z, Weinberger KQ (eds) Advances in Neural Information Processing Systems 26, Curran Associates, Inc., pp 827–835, http://papers.nips.cc/paper/5123-robust-data-driven-dynamic-programming.pdf

  • Hanasusanto, G. A., Kuhn, D., Wallace, S. W., & Zymler, S. (2015). Distributionally robust multi-item newsvendor problems with multimodal demand distributions. Mathematical Programming, 152(1), 1–32.

    Article  Google Scholar 

  • Hu, Z., & Hong, LJ. (2012) Kullback-Leibler divergence constrained distributionally robust optimization. Optimization Online http://www.optimization-online.org/DB_HTML/2012/11/3677.html

  • Izenman, A. J. (1991). Recent developments in nonparametric density estimation. Journal of the American Statistical Association, 86(413), 205–224.

    Google Scholar 

  • Jiang, R., & Guan, Y. (2016). Data-driven chance constrained stochastic program. Mathematical Programming, 158(1), 291–327.

    Article  Google Scholar 

  • Jones, M. C., Marron, J. S., & Sheather, S. J. (1996). A brief survey of bandwidth selection for density estimation. Journal of the American Statistical Association, 91(433), 401–407.

    Article  Google Scholar 

  • Morgan, J. P. (1996). \(RiskMetrics^{{{{\rm TM}}}}\) (4th ed.). Technical Document, J.P: Morgan Company, New York.

  • Konno, H., & Yamazaki, H. (1991). Mean-absolute deviation portfolio optimization model and its applications to tokyo stock market. Management Science, 37(5), 519–531.

    Article  Google Scholar 

  • Krokhmal, P. A. (2007). Higher moment coherent risk measures. Quantitative Finance, 7(4), 373–387.

    Article  Google Scholar 

  • Krokhmal, P. A., Zabarankin, M., & Uryasev, S. (2011). Modeling and optimization of risk. Surveys in Operations Research and Management Science, 16(2), 49–66.

    Article  Google Scholar 

  • Li, Q., & Racine, J. S. (2007). Nonparametric econometrics: theory and practice. Princeton New Jersey: Princeton University Press.

  • Lobo, M. S., Vandenberghe, L., Boyd, S., & Lebret, H. (1998). Applications of second-order cone programming. Linear Algebra and its Applications, 284, 193–228.

    Article  Google Scholar 

  • Markowitz, H. (1952). Portfolio selection. Journal of Finance, 7(1), 77–91.

    Google Scholar 

  • Markowitz, H. (1959). Portfolio selection: efficient diversification of investments. New Jersey: Wiley.

  • Mohajerin Esfahani, P., & Kuhn, D. (2018). Data-driven distributionally robust optimization using the Wasserstein metric: Performance guarantees and tractable reformulations. Mathematical Programming, 171(1), 115–166.

    Article  Google Scholar 

  • Nemirovski, A., & Shapiro, A. (2006). Convex approximations of chance constrained programs. SIAM Journal on Optimization, 17(4), 969–996.

    Article  Google Scholar 

  • Ogryczak, W., & Ruszczyński, A. (2001). On consistency of stochastic dominance and mean-semideviation models. Mathematical Programming, 89(2), 217–232.

    Article  Google Scholar 

  • Popescu, I. (2007). Robust mean-covariance solutions for stochastic optimization. Operations Research, 55(1), 98–112.

    Article  Google Scholar 

  • Postek, K., den Hertog, D., & Melenberg, B. (2016). Computationally tractable counterparts of distributionally robust constraints on risk measures. SIAM Review, 58(4), 603–650.

    Article  Google Scholar 

  • Rockafellar, R. T. (1997). Convex Analysis. Princeton New Jersey: Princeton University Press.

  • Rockafellar, R. T., & Uryasev, S. (2000). Optimization of conditional value-at-risk. Journal of Risk, 2(3), 21–42.

    Article  Google Scholar 

  • Rockafellar, R. T., & Uryasev, S. (2002). Conditional value-at-risk for general loss distributions. Journal of Banking and Finance, 26(7), 1443–1471.

    Article  Google Scholar 

  • Rockafellar, R. T., Uryasev, S., & Zabarankin, M. (2006). Generalized deviations in risk analysis. Finance and Stochastics, 10(1), 51–74.

    Article  Google Scholar 

  • Scarf, H. (1958). A min-max solution of an inventory problem. Stanford University Press. In H. Scarf, K. Arrow, & S. Karlin (Eds.), Studies in the mathematical theory of inventory and production (Vol. 10, pp. 201–209). CA: Stanford.

  • Shapiro, A. (2017). Distributionally robust stochastic programming. SIAM Journal on Optimization, 27(4), 2258–2275.

    Article  Google Scholar 

  • Shapiro, A., Dentcheva, D., & Ruszczyński, A. (2014). Lectures on stochastic programming: Modeling and theory (2nd ed.). Philadelphia, USA: SIAM.

  • Silverman, B. W. (1986). Density estimation for statistics and data analysis (2nd ed.). Chapman and Hall, New York, NY: MPS-SIAM series on optimization.

  • Wang, Z., Glynn, P. W., & Ye, Y. (2016). Likelihood robust optimization for data-driven problems. Computational Management Science, 13(2), 241–261.

    Article  Google Scholar 

  • Wiesemann, W., Kuhn, D., & Sim, M. (2014). Distributionally robust convex optimization. Operations Research, 62(6), 1358–1376.

    Article  Google Scholar 

  • Wozabal, D. (2014). Robustifying convex risk measures for linear portfolios: A nonparametric approach. Operations Research, 62(6), 1302–1315.

    Article  Google Scholar 

  • Yao, H., Li, Z., & Lai, Y. (2013). Mean-CVaR portfolio selection: A nonparametric estimation framework. Computers and Operations Research, 40(4), 1014–1022.

    Article  Google Scholar 

  • Zhao, C., & Guan, Y. (2018). Data-driven risk-averse stochastic optimization with Wasserstein metric. Operations Research Letters, 46(2), 262–267.

    Article  Google Scholar 

  • Zhu, S., & Fukushima, M. (2009). Worst-case conditional value-at-risk with application to robust portfolio management. Operational Research, 57(5), 1155–1168.

    Article  Google Scholar 

  • Zymler, S., Kuhn, D., & Rustem, B. (2013). Distributionally robust joint chance constraints with second-order moment information. Mathematical Programming, 137(1), 167–198.

    Article  Google Scholar 

Download references

Acknowledgements

We wish to thank two anonymous reviewers for very constructive comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Bo Yu.

Additional information

Publisher's Note

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

This work was supported by the National Natural Science Foundation of China (11971092, 11571061, 11301050, 11401075) and the Fundamental Research Funds for the Central Universities (DUT15RC(3)037, DUT18RC(4)067).

Appendix

Appendix

1.1 Proof of Proposition 1

Proof

(1) It follows from Assumption 1 and \(\psi _p'(x)=(G_p(x))^{\frac{1}{p}-1}G_{p-1}(x)\ge 0\) that \(\frac{\partial }{\partial x}\varPsi _p(x,h)=\psi _p'(\frac{x}{h})\ge 0\).

$$\begin{aligned} \frac{\partial }{\partial h}\varPsi _p(x,h)= & {} \left( G_p\left( \frac{x}{h}\right) \right) ^{\frac{1}{p}} -\frac{x}{h}\left( G_p\left( \frac{x}{h}\right) \right) ^{\frac{1}{p}-1} G_{p-1}\left( \frac{x}{h}\right) \\= & {} \left( G_p\left( \frac{x}{h}\right) \right) ^{\frac{1}{p}-1} \left( G_{p}\left( \frac{x}{h}\right) -\frac{x}{h}G_{p-1}\left( \frac{x}{h}\right) \right) \\= & {} \left( G_p\left( \frac{x}{h}\right) \right) ^{\frac{1}{p}-1} \left( \int _{-\infty }^{x/h} \left( \frac{x}{h}-t\right) ^{p-1}(-t)k(t)dt\right) \\\ge & {} \left( G_p\left( \frac{x}{h}\right) \right) ^{\frac{1}{p}-1} \left( \int _{-x/h}^{0} \left( \frac{x}{h}-t\right) ^{p-1}(-t)k(t)dt\right. \\&+ \left. \int _{0}^{x/h}\left( \frac{x}{h}-t\right) ^{p-1}(-t)k(t)dt\right) \\= & {} \left( G_p\left( \frac{x}{h}\right) \right) ^{\frac{1}{p}-1} \left( \int _{0}^{x/h} \left( \left( \frac{x}{h}+t\right) ^{p-1} -\left( \frac{x}{h}-t\right) ^{p-1}\right) tk(t)dt\right) \ge 0, \end{aligned}$$

where the last equality holds from Assumption 2 and the last inequality holds from the monotonicity of function \(f(x)=x^{p-1},x\ge 0\) for \(p\ge 1\). Thus \(\nabla \varPsi _p(x,h)\ge 0\)

(2) By \(\psi _p''(x)=(p-1)\left( G_{p}(x)\right) ^{\frac{1}{p}-2}\left[ G_{p}(x)G_{p-2}(x)-(G_{p-1}(x))^2\right] \ge 0, \) where the inequality holds from Cauchy-Schwarz inequality, and the convexity-preserving property of the operation of right multiplication (see Rockafellar 1997, p. 35), we know that \(\varPsi _p(x,h)\) is a jointly convex function.

(3) For \(x<0\), we have \( \lim \limits _{h\rightarrow 0^+}(\varPsi _p(x,h))^p =x^p\lim \limits _{h\rightarrow 0^+}\int _{-\infty }^{x/h}\left( 1-\frac{th}{x}\right) ^pk(t)dt=0. \) For \(x=0\), we have \( \lim \limits _{h\rightarrow 0^+}(\varPsi _p(0,h))^p=\lim \limits _{h\rightarrow 0^+}h^p\int _{-\infty }^{0}(-t)^pk(t)dt=0. \) For \(x>0\), we have

$$\begin{aligned} \lim _{h\rightarrow 0^+}(\varPsi _p(x,h))^p= & {} x^p\lim _{h\rightarrow 0^+}\int _{-\infty }^{x/h}\left( 1-\frac{th}{x}\right) ^pk(t)dt \\= & {} x^p\lim _{h\rightarrow 0^+}\left( \int _{-\infty }^{-x/h}\left( 1-\frac{th}{x}\right) ^pk(t)dt+ \int _{-x/h}^{-x/h}\left( 1-\frac{th}{x}\right) ^pk(t)dt\right) \\= & {} x^p\lim _{h\rightarrow 0^+}\Big (\int _{-\infty }^{-x/h}\left( 1-\frac{th}{x}\right) ^pk(t)dt\\&+ \int _{-x/h}^{-x/h}\left[ \sum _{n=0}^\infty \frac{p(p-1)\cdots (p-n+1)}{n!} \left( \frac{th}{x}\right) ^n\right] k(t)dt\Big )\\= & {} x^p\lim _{h\rightarrow 0^+}\sum _{n=0}^\infty \frac{p(p-1)\cdots (p-n+1)h^n}{n!x^n}\int _{-x/h}^{-x/h}t^nk(t)dt\\= & {} x^p+x^p\lim _{h\rightarrow 0^+}\sum _{n=1}^\infty \frac{p(p-1)\cdots (p-n+1)h^n}{n!x^n}\int _{-x/h}^{-x/h}t^nk(t)dt=x^p, \end{aligned}$$

where the third equality holds from Maclaurin series of \(f(x)=(1+x)^p,x\in (-1,1),~p\ge 1\) and the last equality holds from Assumption 2. Thus \(\lim \limits _{h\rightarrow 0^+}\varPsi _p(x,h)=\lim \limits _{h\rightarrow 0^+}h\psi _p(h^{-1}x)=[x]^+\). It follows from the definition of recession function (see Rockafellar 1997, p. 66-67) that \([x]^+\) is the recession function of \(\psi _p(x)\). Moreover, for \(x\ne 0\), by using L’Hospital’s rule and Assumption 2, we have

$$\begin{aligned}&\lim _{h\rightarrow 0^+}\frac{\varPsi _p(x,h)-[x]^+}{h}\\= & {} \lim _{h\rightarrow 0^+}\frac{x\left[ \int _{-\infty }^{x/h}\left( 1-\frac{th}{x}\right) ^pk(t)dt\right] ^{1/p}-[x]^+}{h} \\= & {} \lim _{h\rightarrow 0^+}\left[ \int _{-\infty }^{x/h}\left( 1-\frac{th}{x}\right) ^pk(t)dt\right] ^{1/p-1} \int _{-\infty }^{x/h}\left( 1-\frac{th}{x}\right) ^{p-1}tk(t)dt=0. \end{aligned}$$

For \(x=0\), it follows from Assumption 2 that

$$\begin{aligned} \lim _{h\rightarrow 0^+}\frac{\varPsi _p(0,h)}{h} = \lim _{h\rightarrow 0^+}\frac{\left[ h^p\int _{-\infty }^{0}(-t)^pk(t)dt\right] ^{1/p}}{h} = \left( \int _{-\infty }^{0}(-t)^pk(t)dt\right) ^{1/p}\ne 0. \end{aligned}$$

\(\square \)

1.2 Proof of Theorem 1

Proof

Due to the fact that \(f(\varvec{x})=\Vert \varvec{x}\Vert _p, \varvec{x}\in \mathbb {R}_+^n\) is nondecreasing in \(x_i\) for \(i=1,\ldots ,n\), problem (6) is equivalent to

$$\begin{aligned} \begin{array}{cl} \min &{}\rho \alpha -\sum _{i=1}^T\lambda _i\varvec{r}_i^\mathsf {T}\varvec{w}+ \frac{\rho \nu }{1-\gamma }\\ \mathrm {s.t.} &{}\Vert \varvec{y}\Vert _p\le \nu ,\\ &{} \lambda _i^{1/p}\varPsi _p(-\varvec{r}_i^\mathsf {T}\varvec{w}-\alpha ,h)\le y_i,~~i=1,\ldots ,T,\\ &{} \varvec{w}\in \mathcal {W},\alpha \in \mathbb {R},\nu \in \mathbb {R},\varvec{y}\in \mathbb {R}_+^T,\\ \end{array} \end{aligned}$$

where the first constraint \(\Vert \varvec{y}\Vert _p\le \nu \) can be reformulated as \(\left( \varvec{y},\nu \right) \in C_p^T\) and \(C_p^T=\{(\varvec{x},t)\in \mathbb {R}^T_+\times \mathbb {R}_+:\Vert \varvec{x}\Vert _p\le t\}\) is the T-dimensional p-order cone within the nonnegative orthant (see Krokhmal 2007). Since \(f(\varvec{x})=\Vert \varvec{x}\Vert _p, \varvec{x}\in \mathbb {R}_+^n\) is a convex function, we know that the first constraint is a convex constraint. By Proposition 1, (1)-(2) and Lemma 1 (\(m=2\)), we know \(\varPsi \left( u_i(\varvec{w}, \alpha ),h\right) \) is convex in \(\varvec{w}\) and \(\alpha \) for \(i=1,\ldots , T\), where \(u_i(\varvec{w},\alpha ):= -\varvec{r}_i^\mathsf {T}\varvec{w}-\alpha \) is linear in \(\varvec{w}\) and \(\alpha \) for \(i=1,\ldots ,T\), thus the conclusion holds. \(\square \)

1.3 Proof of Corollary 1

Proof

The proof of this result is quite similar to that of Theorem 1 due to the fact that \(h=c\Vert MR\varvec{w}\Vert \) is convex and so is omitted. \(\square \)

1.4 Proof of Lemma 3

Proof

By \(\frac{\partial ^2 f}{\partial x^2}=p(p-1)y^{-q}x^{p-2}\), \(\frac{\partial ^2 f}{\partial y^2}=q(q+1)y^{-q-2}x^{p}\) and \(\frac{\partial ^2 f}{\partial x\partial x} =-pqy^{-q-1}x^{p-1}\), we have the Hessian of f(xy)

$$\begin{aligned} \left( \begin{array}{cc} p(p-1)y^{-q}x^{p-2} &{} -pqy^{-q-1}x^{p-1} \\ -pqy^{-q-1}x^{p-1} &{} q(q+1)y^{-q-2}x^{p} \\ \end{array} \right) \succeq 0 \end{aligned}$$

and \(\frac{\partial f}{\partial x}=py^{-q}x^{p-1}\ge 0\), thus the conclusion holds. \(\square \)

1.5 Proof of Corollary 2

Proof

For \(p=\frac{t}{s}\), the first group of constraints in problem (15) can be reformulated as

$$\begin{aligned} \left\{ \begin{aligned}&z_i^t\le y^{t-s}(v_i+\varvec{r}_i^\mathsf {T}\varvec{w})^s,~~ i=1,\ldots , T,\\&v_i+\varvec{r}_i^\mathsf {T}\varvec{w}\ge 0,i=1,\ldots , T,~~i=1,\cdots ,T,\\&\varPsi _p\left( -\varvec{r}_i^\mathsf {T}\varvec{w}-\alpha ,h\right) \le z_i,~~i=1,\ldots , T,\\&\varvec{w}\in \mathcal {W},\alpha \in \mathbb {R}, u\in \mathbb {R}_+,\eta \in \mathbb {R},y\in \mathbb {R}_+,\varvec{v}\in \mathbb {R}^T,\varvec{z}\in \mathbb {R}^T. \end{aligned} \right. \end{aligned}$$

Thus the conclusion holds. \(\square \)

1.6 Proof of Proposition 2

Proof

We only prove problem (12) is equivalent to \(\min \limits _{( w,\alpha ) \in \mathcal {W}\times \mathcal {D}}\max \limits _{\lambda \in \varLambda _\phi ^\tau } \hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda })\), for some bounded and closed interval \(\mathcal {D}\subset \mathbb {R}\), and omit the proof of the similar and simpler result associated with problem (3). From Assumption 3 and Proposition 1, (3), we have

$$\begin{aligned} \hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda })= \rho \alpha -\sum _{i=1}^T \lambda _i\varvec{r}_i^\mathsf {T}\varvec{w} +\frac{\rho }{1-\gamma }\left( \sum _{i=1}^T \lambda _i\left( [-\varvec{r}_i^\mathsf {T}\varvec{w}-\alpha ]^+\right) ^p\right) ^{1/p}+O(h). \end{aligned}$$

It follows from the boundedness of \(\varOmega \) and \(\mathcal {W}\) that, for \(i=1,\ldots ,T\), \(\varvec{r}_i^\mathsf {T}\varvec{w}\) is bounded, and there exists a positive number B such that \(\varvec{r}_i^\mathsf {T}\varvec{w}\le B\). Thus, we obtain

$$\begin{aligned} \lim _{\alpha \rightarrow +\infty }\max _{\lambda \in \varLambda _\phi ^\tau } \hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda })= & {} \lim _{\alpha \rightarrow +\infty }\max _{\lambda \in \varLambda _\phi ^\tau }\left\{ \rho \alpha -\sum _{i=1}^T\lambda _i \varvec{r}_i^\mathsf {T}\varvec{w}+O(h)\right\} \\\ge & {} \lim _{\alpha \rightarrow +\infty }\left\{ \rho \alpha -B+O(h)\right\} = +\infty \end{aligned}$$

and

$$\begin{aligned} \lim _{\alpha \rightarrow -\infty }\max _{\lambda \in \varLambda _\phi ^\tau } \hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda })= & {} \lim _{\alpha \rightarrow -\infty }\max _{\lambda \in \varLambda _\phi ^\tau }\left\{ -\frac{\rho \gamma \alpha }{1-\gamma }-\sum _{i=1}^T\lambda _i \varvec{r}_i^\mathsf {T}\varvec{w}+O(h)\right\} \\\ge & {} \lim _{\alpha \rightarrow -\infty }\left\{ -\frac{\rho \alpha \gamma }{1-\gamma }-B+O(h)\right\} = +\infty . \end{aligned}$$

Therefore there exists a bounded and closed interval \(\mathcal {D}\subset \mathbb {R}\) such that the minimizer of problem (12) falls into the compact set \(\mathcal {W}\times \mathcal {D}\), that is, problem (12) is equivalent to \(\min \limits _{( w,\alpha ) \in \mathcal {W}\times \mathcal {D}}\max \limits _{\lambda \in \varLambda _\phi ^\tau } \hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda })\). \(\square \)

1.7 Proof of Proposition 3

Proof

Denote the objective function of problem (14) by \(H^{\gamma ,p,\rho }_{T}(\varvec{w}, \alpha ; u,\eta ,z)\) and using Theorem 2, we have

$$\begin{aligned} \max _{\lambda \in \varLambda _\phi ^\tau } \hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda }) =\min _{u\ge 0,z\ge 0,\eta }H^{\gamma ,p,\rho }_{T}(\varvec{w}, \alpha ; u,\eta ,z). \end{aligned}$$

Therefore, if the following two inequalities can be proved, the conclusion holds.

$$\begin{aligned} \lim _{\tau \rightarrow 0}\max _{\varvec{\lambda }\in \varLambda _{\phi }^\tau } \hat{F}^{\gamma ,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda })\ge F^{p,\rho }_{\gamma }(\varvec{w},\alpha ) \end{aligned}$$
(34)

and

$$\begin{aligned} \lim _{\tau \rightarrow 0}\min _{u\ge 0,z\ge 0,\eta } H^{\gamma ,p,\rho }_{T}(\varvec{w}, \alpha ; u,\eta ,z) \le F^{p,\rho }_{\gamma }(\varvec{w},\alpha ). \end{aligned}$$
(35)

It is easy to show that

$$\begin{aligned} \lim _{T\rightarrow \infty }\max _{\varvec{\lambda }\in \varLambda _{\phi }^{\tau }} \hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda }) \ge \lim _{T\rightarrow \infty }\hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda }^0) =\lim _{T\rightarrow \infty }\tilde{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ) \overset{\mathrm {w.p.1}}{=}F^{p,\rho }_{\gamma }(\varvec{w},\alpha ), \end{aligned}$$

where the first equality holds from Assumption 3 and Proposition 1, (3) and the last equality holds from Shapiro et al. (2014, Theorem 5.3). On the other hand,

$$\begin{aligned}&\lim _{T\rightarrow \infty }\min _{u\ge 0,z\ge 0,\eta } H^{\gamma ,p,\rho }_{T}(\varvec{w}, \alpha ; u,\eta ,z) \nonumber \\&\quad \le \min _{u\ge 0,z\ge 0, \eta }\limsup _{T\rightarrow \infty } H^{\gamma ,p,\rho }_{T}(\varvec{w}, \alpha ; u,\eta ,z) \nonumber \\&\quad \overset{\mathrm {w.p.1}}{=} \min _{u\ge 0,z\ge 0,\eta } \Bigg \{\eta + \mathbb {E}_P\left[ u\phi ^*\left( \frac{ z\left( \left[ -\varvec{\xi }^\mathsf {T}\varvec{w}-\alpha \right] ^+\right) ^p-\varvec{\xi }^\mathsf {T}\varvec{w} -\eta }{u}\right) \right] +\rho \alpha \nonumber \\&\qquad -\left( \frac{\rho }{p(1-\gamma )}\right) ^{\frac{p}{p-1}}(1-p)z^{\frac{1}{1-p}} \Bigg \} \end{aligned}$$
(36)
$$\begin{aligned}&\le \min _{z\ge 0} \Bigg \{ \rho \alpha -\frac{1}{T}\sum _{i=1}^T\varvec{r}_i^\mathsf {T}\varvec{w}+z\mathbb {E}_P\left[ \left( \left[ -\varvec{\xi }^\mathsf {T}\varvec{w}-\alpha \right] ^+\right) ^p\right] \nonumber \\&\qquad -\left( \frac{\rho }{p(1-\gamma )}\right) ^{\frac{p}{p-1}}(1-p)z^{\frac{1}{1-p}}\Bigg \} \end{aligned}$$
(37)
$$\begin{aligned}&\quad = \rho \alpha -\mathbb {E}_P\left[ \varvec{\xi }^\mathsf {T}\varvec{w}\right] +\frac{\rho }{(1-\gamma )}\mathbb {E}_P\left[ \left( \left( \left[ -\varvec{\xi }^\mathsf {T}\varvec{w}-\alpha \right] ^+\right) ^p\right) ^{1/p}\right] \nonumber \\&\quad = F^{p,\rho }_{\gamma }(\varvec{w},\alpha ), \end{aligned}$$
(38)

where (36) holds from Assumptions 34, Proposition 1, (3) and Shapiro et al. (2014, Theorem 5.3), and (37) holds from Assumption 5. Moreover, (38) holds from the fact that the concave conjugate function of \(f(x)=\frac{x^p}{p}(x\ge 0,~0<p<1)\) is \(f_*(y)=\frac{y^q}{q}(y\ge 0)\), where \(1/p+1/q=1\) (see Ben-Tal et al. 2015). Combining (34) and (35), for given \(\varvec{w}\) and \(\alpha \), we have \(\lim _{T\rightarrow \infty }\max _{\varvec{\lambda }\in \varLambda _{\phi }^\tau } \hat{F}^{\gamma ,p,\rho }_{T}(\varvec{w},\alpha ;\varvec{\lambda })\overset{\mathrm {w.p.1}}{=} F^{p,\rho }_{\gamma }(\varvec{w},\alpha )\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Liu, W., Yang, L. & Yu, B. KDE distributionally robust portfolio optimization with higher moment coherent risk. Ann Oper Res 307, 363–397 (2021). https://doi.org/10.1007/s10479-021-04171-4

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-021-04171-4

Keywords

Navigation