Skip to main content
Log in

Portfolio optimization and marginal contribution to risk on multivariate normal tempered stable model

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

Abstract

This paper proposes a market model with returns assumed to follow a multivariate normal tempered stable distribution defined by a mixture of the multivariate normal distribution and the tempered stable subordinator. This distribution can capture two stylized facts: fat-tails and asymmetry, that have been empirically observed for asset return distributions. We discuss a new portfolio optimization method on the new market model, which is an extension of Markowitz’s mean-variance optimization. The new optimization method considers not only reward and dispersion but also asymmetry in tails. The efficient frontier is extended to a curved surface on three-dimensional space of reward, dispersion, and asymmetry in tails. We also propose a new performance measure, which is an extension of the Sharpe ratio. Moreover, we derive closed-form solutions for portfolio managers’ two important measures in portfolio construction: the marginal value-at-risk (VaR) and the marginal conditional VaR (CVaR). We illustrate the proposed model using stocks comprising the Dow Jones Industrial Average. First, perform the new portfolio optimization and then demonstrating how the marginal VaR and marginal CVaR can be used for portfolio optimization under the model. Based on this paper’s empirical evidence, our framework offers realistic portfolio optimization and tractable methods for portfolio risk management.

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

Similar content being viewed by others

Notes

  1. See Carr et al. (2002), Koponen (1995), and Boyarchenko and Levendorskiĭ (2000)

  2. The tempered subordinator is defined by the characteristic function (12) in the “Appendix”

  3. In this paper, we consider the long only portfolio.

  4. Kim et al. (2012) and Anand et al. (2016) used the DJIA index, and Anand et al. (2017) used the S&P 1200 financial index (SGFS) as the index to estimate the parameters (\(\alpha \), \(\theta \)).

  5. CDF of stdNTS distribution is obtained by the fast Fourier transform method by Gils-Pelaez (1951).

  6. We select the sample size 752 (3 years) to see at least two outliers (left and right tails). If the data is normal, we expect more than 2 outliers which are not included \(3\sigma \) range covers 99.7% out of 752 samples.

  7. More precisely, it has the semi-fat-tails having the exponential decaying tails

  8. See Stoyanov et al. (2010) and Hitaj et al. (2019).

  9. See Proposition 2 in Kim et al. (2010) and (2012).

References

  • Aas, K., Hobæk Haff, I., & Dimakos, X. K. (2006). Risk estimation using the multivariate normal inverse Gaussian distribution. Journal of Risk, 8(2), 39–60.

    Google Scholar 

  • Adcock, C. (2010). Asset pricing and portfolio selection based on the multivariate extended skew-Student-\(t\) distribution. Annals of Operations Research, 176, 221–234.

    Google Scholar 

  • Anand, A., Li, T., Kurosaki, T., & Kim, Y. S. (2016). Foster–Hart optimal portfolios. Journal of Banking and Finance, 68, 117–130.

    Google Scholar 

  • Anand, A., Li, T., Kurosaki, T., & Kim, Y. S. (2017). The equity risk posed by the too-big-to-fail banks: A foster-hart estimation. Annals of Operations Research, 253(1), 21–41.

    Google Scholar 

  • Barndorff-Nielsen, O. E., & Levendorskii, S. (2001). Feller processes of normal inverse Gaussian type. Quantitative Finance, 1, 318–331.

    Google Scholar 

  • Barndorff-Nielsen, O. E., & Shephard, N. (2001). Normal modified stable processes. Economics Series Working Papers from University of Oxford, Department of Economics 72.

  • Bianchi, M. L., & Tassinari, G. L. (2020). Forward-looking portfolio selection with multivariate non-Gaussian models. Quantitative Finance, 20(10), 1645–1661.

    Google Scholar 

  • Bianchi, M. L., Tassinari, G. L., & Fabozzi, F. J. (2016). Riding with the four horsemen and the multivariate normal tempered stable model. International Journal of Theoretical and Applied Finance, 19(4), 1650027.

    Google Scholar 

  • Boyarchenko, S. I., & Levendorskiĭ, S. Z. (2000). Option pricing for truncated Lévy processes. International Journal of Theoretical and Applied Finance, 3, 549–552.

    Google Scholar 

  • Carr, P., Geman, H., Madan, D., & Yor, M. (2002). The fine structure of asset returns: An empirical investigation. Journal of Business, 75(2), 305–332.

    Google Scholar 

  • Carr, P., Geman, H., Madan, D., & Yor, M. (2003). Stochastic volatility for Lévy processes. Mathematical Finance, 13, 345–382.

    Google Scholar 

  • Dahlquist, M., Farago, A., & Tédongap, R. (2017). Asymmetries and portfolio choice. The Review of Financial Studies, 30(2), 667–702.

    Google Scholar 

  • Eberlein, E., & Glau, K. (2014). Variational solutions of the pricing pides for European options in Lévy models. Applied Mathematical Finance, 21(5), 417–450.

    Google Scholar 

  • Eberlein, E., Gehrig, T., & Madan, D. (2012). Pricing to acceptability: With applications to valuation of one’s own credit risk. Journal of Risk, 15, 91–120. https://doi.org/10.21314/JOR.2012.252

    Article  Google Scholar 

  • Eberlein, E., & Madan, D. B. (2010). On correlating Lévy processes. Journal of Risk, 13(1), 3–16.

    Google Scholar 

  • Eberlein, E., & Özkan. (2005). The Lévy LIBOR model. Finance and Stochastics, 9, 327–348.

    Google Scholar 

  • Fallahgoul, H. A., Kim, Y. S., Fabozzi, F. J., & Park, J. (2019). Quanto option pricing with Lévy models. Computational Economics, 53(3), 1279–1308.

    Google Scholar 

  • Fama, E. (1963). Mandelbrot and the stable Paretian hypothesis. Journal of Business, 36, 420–429.

    Google Scholar 

  • Gil-Pelaez, J. (1951). Note on the inversion theorem. Biometrika, 38, 3–4.

    Google Scholar 

  • Gourieroux, C., Laurent, J., & Scaillet, O. (2000). Sensitivity analysis of values at risk. Journal of Empirical Finance, 7, 225–245.

    Google Scholar 

  • Harvey, C. R., Liechty, J. C., Liechty, M. W., & Müller, P. (2010). Portfolio selection with higher moments. Quantitative Finance, 10(5), 469–485. https://doi.org/10.1080/14697681003756877

    Article  Google Scholar 

  • Hitaj, A., Mercuri, L., & Rroji, E. (2019). Sensitivity analysis of mixed tempered stable parameters with implications in portfolio optimization. Computational Management Science, 16, 71–95. https://doi.org/10.1080/00949655.2021.1962878

    Article  Google Scholar 

  • Kim, S. I., & Kim, Y. S. (2018). Normal tempered stable structural model. Review of Derivatives Research, 21(1), 119–148.

    Google Scholar 

  • Kim, Y., Lee, J., Mittnik, S., & Park, J. (2015). Quanto option pricing in the presence of fat tails and asymmetric dependence. Journal of Econometrics, 187(2), 512–520.

    Google Scholar 

  • Kim, Y. S., Giacometti, R., Rachev, S. T., Fabozzi, F. J., & Mignacca, D. (2012). Measuring financial risk and portfolio optimization with a non-Gaussian multivariate model. Annals of Operations Research, 201(1), 325–343.

    Google Scholar 

  • Kim, Y. S., Rachev, S. T., Bianchi, M. L., & Fabozzi, F. J. (2010). Computing VaR and AVaR in infinitely divisible distributions. Probability and Mathematical Statistics, 30(2), 223–245.

    Google Scholar 

  • Kim, Y. S., Roh, K. H., & Douady, R. (2021). Tempered stable processes with time-varying exponential tails. Quantitative Finance.https://doi.org/10.1080/14697688.2021.1962958, to appear.

  • Kim, Y. S., & Volkmann, D. (2013). NTS copula and finance. Applied Mathematics Letters, 26, 676–680.

    Google Scholar 

  • King, A. J. (1993). Asymmetric risk measures and tracking models for portfolio optimization under uncertainty. Annals of Operations Research, 45, 165–177.

    Google Scholar 

  • Konno, H., Shirakawa, H., & Yamazaki, H. (1993). A mean-absolute deviation-skewness portfolio optimization model. Annals of Operations Research, 45, 205–220.

    Google Scholar 

  • Koponen, I. (1995). Analytic approach to the problem of convergence of truncated Lévy flights towards the Gaussian stochastic process. Physical Review E, 52, 1197–1199.

    Google Scholar 

  • Kraus, A., & Litzenberger, R. (1976). Skewness preference and the valuation of risk assets. Journal of Finance, 31, 1085–1100.

    Google Scholar 

  • Kurosaki, T., & Kim, Y. S. (2018). Foster–Hart optimization for currency portfolio. Studies in Nonlinear Dynamics & Econometrics, 23(2):Published Online. https://doi.org/10.1515/snde-2017-0119.

  • Liu, Y., Djurić, P. M., Kim, Y. S., Rachev, S. T., & Glimm, J. (2021). Systemic risk modeling with lévy copulas. Journal of Risk and Financial Management. https://doi.org/10.3390/jrfm14060251.

  • Mandelbrot, B. B. (1963a). New methods in statistical economics. Journal of Political Economy, 71, 421–440.

    Google Scholar 

  • Mandelbrot, B. B. (1963b). The variation of certain speculative prices. Journal of Business, 36, 394–419.

    Google Scholar 

  • Mansini, R., Ogryczak, W., & Speranza, M. G. (2007). Conditional value at risk and related linear programming models for portfolio optimization. Annals of Operations Research, 152, 227–256.

    Google Scholar 

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

    Google Scholar 

  • Øigård, T. A., Hanssen, A., Hansen, R. E., & Godtliebsen, F. (2005). EM-estimation and modeling of heavy-tailed processes with the multivariate normal inverse Gaussian distribution. Signal Processing, 85, 1655–1673.

    Google Scholar 

  • Pflug, G. (2000). Some remarks on the value-at-risk and the conditional value-at-risk. In S. Uryasev (Ed.), Probabilistic constrained optimization: Methodology and applications (pp. 272–281). The Netherlands: Kluwer Academic Publishers.

    Google Scholar 

  • Rachev, S. T., & Mittnik, S. (2000). Stable Paretian models in finance. New York: Wiley.

    Google Scholar 

  • Rachev, S. T., Stoyanov, S., & Fabozzi, F. J. N. J. (2007). Advanced stochastic models, risk assessment, and portfolio optimization: The ideal risk, uncertainty, and performance measures. Hoboken: Wiley.

    Google Scholar 

  • Rachev, S. T., Kim, Y. S., Bianchi, M. L., & Fabozzi, F. J. (2011). Financial models with Lévy processes and volatility clustering. New York: Wiley.

    Google Scholar 

  • Rockafellar, R. T., & Uryasev, S. (2000). Optimization of conditional value-at-risk. Journal of Risk, 2(3), 21–41.

    Google Scholar 

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

    Google Scholar 

  • Rubinstein, M. (1973). The fundamental theorem of parameter-preference security valuation. Journal of Financial and Quantitative Analysis, 8, 61–69.

    Google Scholar 

  • Sharpe, W. F. (1966). Mutual funds performance. Journal of Business, 36(1), 119–138.

    Google Scholar 

  • Sharpe, W. F. (1994). The sharpe ratio. Journal of Portfolio Management, 21(1), 45–58.

    Google Scholar 

  • Stoyanov, S. V., Racheva-Iotova, B., Rachev, S. T., & Fabozzi, F. J. (2010). Stochastic models for risk estimation in volatile markets: A survey. Annals of Operations Research, 176(1), 293–309.

    Google Scholar 

  • Stoyanov, S. V., Rachev, S. T., & Fabozzi, F. J. (2013). Sensitivity of portfolio VaR and CVaR to portfolio return characteristics. Annals of Operations Research, 205, 169–187.

    Google Scholar 

  • Xia, Y., & Grabchak, M. (2021). Estimation and simulation for multivariate tempered stable distributions. Journal of Statistical Computation and Simulation. https://doi.org/10.1080/00949655.2021.1962878

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Young Shin Kim.

Additional information

Publisher's Note

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

The author gratefully acknowledges the support of GlimmAnalytics LLC and Juro Instruments Co., Ltd. The author is grateful to Minseob Kim, who reviewed this paper and corrected editorial errors. Also, all remaining errors are entirely my own.

Appendix: Multivariate normal tempered stable distribution

Appendix: Multivariate normal tempered stable distribution

Let \(\alpha \in (0,2)\) and \(\theta >0\), and let \({{{\mathcal {T}}}}\) be a positive random variable whose characteristic function \(\phi _{{{{\mathcal {T}}}}}\) is equal to

$$\begin{aligned} \phi _{{{{\mathcal {T}}}}}(u) = \exp \left( -\frac{2\theta ^{1-\frac{\alpha }{2}}}{\alpha }\left( (\theta -iu)^{\frac{\alpha }{2}}-\theta ^{\frac{\alpha }{2}}\right) \right) . \end{aligned}$$
(12)

The random variable \({{{\mathcal {T}}}}\) is referred to as Tempered Stable Subordinator. Let \(X=(X_1, X_2, \cdots , X_N)^{{\texttt {T}}}\) be a multivariate random variable given by

$$\begin{aligned} X = \mu + \beta ({{{\mathcal {T}}}}-1) + diag (\gamma ) \varepsilon \sqrt{{{{\mathcal {T}}}}} , \end{aligned}$$

where

  • \(\mu = (\mu _1, \mu _2, \cdots , \mu _N)^{{\texttt {T}}}\in {{\mathbb {R}}}^N\)

  • \(\beta = (\beta _1, \beta _2, \cdots , \beta _N)^{{\texttt {T}}}\in {{\mathbb {R}}}^N\)

  • \(\gamma = (\gamma _1, \gamma _2, \cdots , \gamma _N)^{{\texttt {T}}}\in {{\mathbb {R}}}_+^N\) with \({{\mathbb {R}}}_+=[0,\infty )\)

  • \(\varepsilon = (\varepsilon _1, \varepsilon _2, \cdots , \varepsilon _N)^{{\texttt {T}}}\) is a N-dimensional standard normal distribution with a covariance matrix \(\varSigma \). That is, \(\varepsilon _n\sim \varPhi (0,1)\) for \(n\in \{1,2,\ldots , N\}\) and (kl)th element of \(\varSigma \) is given by \(\rho _{k,l}=\mathrm{cov}(\varepsilon _k,\varepsilon _l)\) for \(k,l\in \{1,2,\ldots ,N\}\).

  • \({{{\mathcal {T}}}}\) is the Tempered Stable Subordinator with parameters \((\alpha ,\theta )\), and is independent of \(\varepsilon _n\) for all \(n=1,2,\ldots , N\).

Then X is referred to as the N-dimensional NTS random variable with parameters \((\alpha \), \(\theta \), \(\beta \), \(\gamma \), \(\mu \), \(\varSigma )\) which we denote by \(X\sim NTS _N(\alpha \), \(\theta \), \(\beta \), \(\gamma \), \(\mu \), \(\varSigma )\). The NTS distribution has the following properties:

  1. 1.

    The mean of X are equal to \(E[X] = \mu \).

  2. 2.

    The covariance between \(X_k\) and \(X_l\) is given by

    $$\begin{aligned} \mathrm{cov}(X_k,X_l)=\rho _{k,l}\gamma _k\gamma _l+\beta _k\beta _l\left( \frac{2-\alpha }{2\theta }\right) \end{aligned}$$
    (13)

    for \(k,l\in \{1,2,\ldots ,N\}\).

  3. 3.

    The variance of \(X_n\) is

    $$\begin{aligned} \mathrm{var}(X_n)=\gamma _n^2+\beta _n^2\left( \frac{2-\alpha }{2\theta }\right) \text { for } n\in \{1,2,\ldots , N\}. \end{aligned}$$
  4. 4.

    Characteristic function of \(X_n\) is

    $$\begin{aligned} \phi _{X_n}(u) = \exp \left( (\mu -\beta )ui-\frac{2\theta ^{1-\frac{\alpha }{2}}}{\alpha } \left( \left( \theta -i\beta u+\frac{\gamma ^2u^2}{2}\right) ^{\frac{\alpha }{2}}-\theta ^{\frac{\alpha }{2}}\right) \right) \end{aligned}$$

Providing \(\mu _n=0\) and \(\gamma _n = \sqrt{1-\beta _n^2 \left( \frac{2-\alpha }{2\theta }\right) }\) with \(|\beta _n|<\sqrt{ \frac{2\theta }{2-\alpha }}\) for n \(\in \) \(\{ 1\),2, \(\cdots \),\(N\}\), the N-dimensional NTS random variable X has \(E[X] = (0,0,\ldots ,0)^{{\texttt {T}}}\) and \(\mathrm{var}(X)\) \(=\) (1,1,\(\cdots \),\(1)^{{\texttt {T}}}\). In this case, X is referred to as the N-dimensional standard NTS random variable with parameters \((\alpha \), \(\theta \), \(\beta \), \(\varSigma )\) and we denote it by \(X\sim stdNTS _N(\alpha \), \(\theta \), \(\beta \), \(\varSigma )\).

For one dimensional NTS distribution, \(\varSigma = 1\), we can prove the following Lemma which is changing parameterization.

Lemma 1

Let \(X\sim NTS_1(\alpha \), \(\theta \), \(\beta \), \(\gamma \), \(\mu \), 1) and \(\xi \sim stdNTS _1({{{\bar{\alpha }}}}\), \({{{\bar{\theta }}}}\), \({{{\bar{\beta }}}}, 1)\). Suppose \({{{\bar{\alpha }}}} = \alpha \), \({{{\bar{\theta }}}} = \theta \), and \( {{{\bar{\beta }}}} = \beta /\sigma \), where \(\sigma =\sqrt{\gamma ^2+\beta ^2\left( \frac{2-\alpha }{2\theta }\right) }\). Then we have \(X = \mu + \sigma \xi \).

Proof

The Ch.F of X is given by

$$\begin{aligned} \phi _{X}(u)&=\exp \left( (\mu -\beta )iu-\frac{2\theta ^{1-\frac{\alpha }{2}}}{\alpha } \left( \left( \theta -i\beta u+\frac{\gamma ^2 u^2}{2}\right) ^{\frac{\alpha }{2}}-\theta ^{\frac{\alpha }{2}}\right) \right) \end{aligned}$$
(14)

By the definition of stdNTS distribution, the Ch.F of \(\mu +\sigma \xi \) is equal to

$$\begin{aligned} \phi _{\mu +\sigma \xi }(u)&= E[e^{(\mu +\sigma \xi )ui}]=e^{\mu ui}E[e^{iu\sigma \xi }] \nonumber \\&=\exp \left( (\mu -{{{\bar{\beta }}}} \sigma ) iu-\frac{2{{{\bar{\theta }}}}^{1-\frac{{{{\bar{\alpha }}}}}{2}}}{{{{\bar{\alpha }}}}} \left( \left( {{{\bar{\theta }}}}-i{{{\bar{\beta }}}} \sigma u+\left( 1-{{{\bar{\beta }}}}^2\left( \frac{2-{{{\bar{\alpha }}}}}{2{{{\bar{\theta }}}}}\right) \right) \frac{\sigma ^2 u^2}{2}\right) ^{\frac{{{{\bar{\alpha }}}}}{2}}-{{{\bar{\theta }}}}^{\frac{{{{\bar{\alpha }}}}}{2}}\right) \right) . \end{aligned}$$
(15)

Hence (14)=(15) if \({{{\bar{\alpha }}}} = \alpha \), \({{{\bar{\theta }}}} = \theta \), \({{{\bar{\beta }}}} \sigma = \beta \), and \(\gamma ^2 =\sigma ^2\left( 1-{{{\bar{\beta }}}}^2\left( \frac{2-\alpha }{2\theta }\right) \right) \). Since \(\sigma = \beta /{{{\bar{\beta }}}}\), we have

$$\begin{aligned} \gamma ^2 = \frac{\beta ^2}{{{{\bar{\beta }}}}^2}\left( 1-{{{\bar{\beta }}}}^2\left( \frac{2-\alpha }{2\theta }\right) \right) , \end{aligned}$$

or

$$\begin{aligned} \gamma ^2{{{\bar{\beta }}}}^2 = \beta ^2\left( 1-{{{\bar{\beta }}}}^2\left( \frac{2-\alpha }{2\theta }\right) \right) = \beta ^2-\beta ^2{{{\bar{\beta }}}}^2\left( \frac{2-\alpha }{2\theta }\right) , \end{aligned}$$

and hence

$$\begin{aligned} {{{\bar{\beta }}}}^2\left( \gamma ^2+\beta ^2\left( \frac{2-\alpha }{2\theta }\right) \right) =\beta ^2. \end{aligned}$$

Therefore, we have

$$\begin{aligned} {{{\bar{\beta }}}} = \frac{\beta }{\sigma }, \end{aligned}$$

where

$$\begin{aligned} \sigma = \sqrt{\gamma ^2+\beta ^2\left( \frac{2-\alpha }{2\theta }\right) }. \end{aligned}$$

\(\square \)

The linear combination of NTS member variables of the NTS vector is again NTS distributed as the following proposition.

Lemma 2

Let \(w = (w_1, w_2, \cdots , w_N)^{\texttt {T}}\in {{\mathbb {R}}}^N\) and \(X\sim NTS _N(\alpha \), \(\theta \), \(\beta \), \(\gamma \), \(\mu \), \(\varSigma )\). Then \(w^{{\texttt {T}}} X \sim NTS _1(\alpha ,\theta ,{{{\bar{\beta }}}},{{{\bar{\gamma }}}},{{{\bar{\mu }}}},1)\), where

$$\begin{aligned} {{{\bar{\mu }}}}=w^{{\texttt {T}}} \mu , ~~~{{{\bar{\beta }}}}=w^{{\texttt {T}}} \beta ~~~ \text { and } ~~~ {{{\bar{\gamma }}}}=\sqrt{w^{\texttt {T}}\mathrm{diag}(\gamma )\varSigma \mathrm{diag}(\gamma ) w }. \end{aligned}$$

Proof

Since we have

$$\begin{aligned} w^{\texttt {T}}X = w^{\texttt {T}}\mu + w^{\texttt {T}}\beta ({{{\mathcal {T}}}}-1) + w^{\texttt {T}}\mathrm{diag}(\gamma )\epsilon \sqrt{{{{\mathcal {T}}}}}, \end{aligned}$$

and \(w^{\texttt {T}}\mathrm{diag}(\gamma )\epsilon {\mathop {=}\limits ^{\mathrm {d}}}\sqrt{w^{\texttt {T}}\mathrm{diag}(\gamma )\varSigma \mathrm{diag}(\gamma ) w } \epsilon _0\) with \(\epsilon _0\sim \varPhi (0,1)\), it is trivial. \(\square \)

Remark 1

In Lemma 2, the first four moments of \(Y=w^{\texttt {T}}X\) are as follows [See Rachev et al. (2011) and Kim et al. (2021)]

  • Mean: \(\displaystyle E[Y]={{{\bar{\mu }}}}\)

  • Variance: \(\displaystyle \mathrm{var}(Y)={{{\bar{\gamma }}}}^2+{{{\bar{\beta }}}}^2\left( \frac{2-\alpha }{2\theta }\right) \)

  • skewness: \(\displaystyle S (Y)= \frac{{{{\bar{\beta }}}}\,\left( 2-\alpha \right) \,\left( 6\,{{{{\bar{\gamma }}}}}^2\,\theta -\alpha {{{\bar{\beta }}}}^2+4{{{\bar{\beta }}}}^2\right) }{\sqrt{2\theta }\,{\left( 2\,{{{{\bar{\gamma }}}}}^2\,\theta -\alpha {{{\bar{\beta }}}}^2+2{{{\bar{\beta }}}}^2\right) }^{3/2}} \)

  • Excess kurtosis:

    \(\displaystyle K (Y){=}\frac{\left( 2{-}\alpha \right) \,\left( {\alpha }^2{{{\bar{\beta }}}}^4{-}10\,\alpha {{{\bar{\beta }}}}^4{-}12\,\alpha {{{\bar{\beta }}}}^2\,{{{{\bar{\gamma }}}}}^2\,\theta {+}24\beta ^4+48{{{\bar{\beta }}}}^2\,{{{{\bar{\gamma }}}}}^2\,\theta {+}12\,{{{{\bar{\gamma }}}}}^4\,{\theta }^2\right) }{2\,\theta \,{\left( 2\,{{{{\bar{\gamma }}}}}^2\,\theta {-}\alpha {{{\bar{\beta }}}}^2+2{{{\bar{\beta }}}}^2\right) }^2} \)

In general, higher moments can be obtained by the cumulants \(c_n(Y)\):

$$\begin{aligned} c_n(Y) = \frac{1}{i^n}\left[ \frac{d^n}{du^n}\log (\phi _Y(u))\right] _{u=0}, n = 1,2,\cdots . \end{aligned}$$

Finally we can provide the proof of Proposition 1.

1.1 Proof of Proposition 1

Proof of Proposition 1

By (1) and (2), we have

$$\begin{aligned} R&= \mu + \mathrm{diag}(\sigma )(\beta ({{{\mathcal {T}}}}-1)+\mathrm{diag}(\gamma )\epsilon \sqrt{{{{\mathcal {T}}}}})\\&= \mu + \mathrm{diag}(\sigma )\beta ({{{\mathcal {T}}}}-1)+\mathrm{diag}(\sigma )\mathrm{diag}(\gamma )\epsilon \sqrt{{{{\mathcal {T}}}}} \\ {}&\sim NTS _N (\alpha , \theta , \mathrm{diag}(\sigma ) \beta , \mathrm{diag}(\sigma ) \gamma , \mu , \varSigma ), \end{aligned}$$

where \(\gamma =(\gamma _1, \gamma _2, \cdots , \gamma _N)^{{\texttt {T}}}\) with \(\gamma _n = \sqrt{1-\beta _n^2 \left( \frac{2-\alpha }{2\theta }\right) }\) and \({{{\mathcal {T}}}}\) is the tempered stable subordinator with parameter \((\alpha ,\theta )\). By Lemma 2,

$$\begin{aligned} R_P(w) = w^{\texttt {T}}R \sim \mathrm{NTS}_1(\alpha , \theta , w^{\texttt {T}}\mathrm{diag}(\sigma )\beta , \sqrt{w^{\texttt {T}}\mathrm{diag}(\sigma )\mathrm{diag}(\gamma ) \varSigma \mathrm{diag}(\gamma )\mathrm{diag}(\sigma )w}, w^{\texttt {T}}\mu , 1). \end{aligned}$$

By Lemma 1, we have

$$\begin{aligned} R_P(w) = w^{\texttt {T}}\mu + {{{\bar{\sigma }}}}(w) \varXi \end{aligned}$$

where

$$\begin{aligned} {{{\bar{\sigma }}}}(w) = \sqrt{w^{\texttt {T}}\mathrm{diag}(\sigma )\mathrm{diag}(\gamma ) \varSigma \mathrm{diag}(\gamma )\mathrm{diag}(\sigma )w+(w^{\texttt {T}}\mathrm{diag}(\sigma )\beta )^2\left( \frac{2-\alpha }{2\theta }\right) } \end{aligned}$$

and

$$\begin{aligned} \varXi \sim \mathrm{stdNTS}_1(\alpha , \theta , w^{\texttt {T}}\mathrm{diag}(\sigma )\beta /{{{\bar{\sigma }}}}(w), 1). \end{aligned}$$

Also, by (13), we have

$$\begin{aligned} \left( {{{\bar{\sigma }}}}(w)\right) ^2&= \sum _{k=1}^N\sum _{l=1}^N w_k w_l \gamma _k\gamma _l\sigma _k\sigma _l \rho _{k,l} + \left( \sum _{k=1}^N w_k\sigma _k\beta _k\right) ^2\left( \frac{2-\alpha }{2\theta }\right) \\&= \sum _{k=1}^N\sum _{l=1}^N w_k w_l \sigma _k\sigma _l \left( \gamma _k\gamma _l\rho _{k,l}+\beta _k\beta _l\left( \frac{2-\alpha }{2\theta }\right) \right) \\&= \sum _{k=1}^N\sum _{l=1}^N w_k w_l \sigma _k\sigma _l\mathrm{cov}(X_k,X_l). \end{aligned}$$

Hence, we have

$$\begin{aligned} \left( {{{\bar{\sigma }}}}(w)\right) ^2 = \sum _{k=1}^N\sum _{l=1}^N w_k w_l \mathrm{cov}(R_k, R_l) = w^{\texttt {T}}\varSigma _R w \end{aligned}$$

where \(\varSigma _R\) is the covariance matrix of R. \(\square \)

1.2 Proof of Proposition 2

Proof of Proposition 2

By (2), we have

$$\begin{aligned} R_P(w) = {{{\bar{\mu }}}}(w) + {{{\bar{\sigma }}}}(w) \varXi ~~~\text { for }~~~ \varXi \sim stdNTS _1(\alpha , \theta , {{{\bar{\beta }}}}(w), 1), \end{aligned}$$

where

$$\begin{aligned} {{{\bar{\mu }}}}(w)&= w^{\texttt {T}}\mu = \sum _{k=1}^N w_k\mu _k, \\ {{{\bar{\beta }}}}(w)&= \frac{w^{\texttt {T}}\mathrm{diag}(\sigma )\beta }{{{{\bar{\sigma }}}}(w)}= \frac{\sum _{k=1}^N w_k\sigma _k\beta _k}{{{{\bar{\sigma }}}}(w)},\\ {{{\bar{\sigma }}}}(w)&=\sqrt{w^{\texttt {T}}\varSigma _R w}=\sqrt{\sum _{k=1}^N\sum _{l=1}^N w_k w_l \mathrm{cov}(R_k, R_l)}. \end{aligned}$$

Hence, the first derivative of \({{{\bar{\beta }}}}(w)\) and \({{{\bar{\sigma }}}}(w)\) are obtained as follows:

$$\begin{aligned} \frac{\partial {{{\bar{\beta }}}}(w)}{\partial w_n}&= \frac{\partial }{\partial w_n} \left( \frac{\sum _{k=1}^N w_k \sigma _k \beta _k}{{{{\bar{\sigma }}}}(w)}\right) \nonumber \\&=\frac{\sigma _n\beta _n{{{\bar{\sigma }}}}(w) - \left( \sum _{k=1}^N w_n \sigma _k \beta _k\right) \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n}}{\left( {{{\bar{\sigma }}}}(w)\right) ^2} \nonumber \\&=\frac{\sigma _n\beta _n}{{{{\bar{\sigma }}}}(w)} - \frac{{{{\bar{\beta }}}}(w)}{{{{\bar{\sigma }}}}(w)} \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n} \end{aligned}$$
(16)

and

$$\begin{aligned} \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n}&= \frac{1}{2{{{\bar{\sigma }}}}(w)} \frac{\partial }{\partial w_n}\sum _{k=1}^N\sum _{l=1}^N w_k w_l \mathrm{cov}(R_k, R_l) \\&= \frac{\sum _{k=1}^Nw_k\mathrm{cov}(R_n, R_k)}{{{{\bar{\sigma }}}}(w)}. \end{aligned}$$

Since we have

$$\begin{aligned} \mathrm{VaR}_\eta (R_p(w)) = -{{{\bar{\mu }}}}(w) - {{{\bar{\sigma }}}}(w) F_\mathrm{stdNTS}^{-1}(\eta , \alpha , \theta , {{{\bar{\beta }}}}(w)) \end{aligned}$$

we obtain

$$\begin{aligned}&\frac{\partial \mathrm{VaR}_\eta (R_p(w))}{\partial w_n} \nonumber \\&\quad = -\frac{\partial {{{\bar{\mu }}}}(w)}{\partial w_n} - \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n} F_\mathrm{stdNTS}^{-1}(\eta , \alpha , \theta , {{{\bar{\beta }}}}(w)) - {{{\bar{\sigma }}}}(w)\frac{\partial F_\mathrm{stdNTS}^{-1}(\eta , \alpha , \theta , {{{\bar{\beta }}}}(w))}{\partial w_n} \nonumber \\&\quad =-\mu _n - \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n}F_\mathrm{stdNTS}^{-1}(\eta , \alpha , \theta , {{{\bar{\beta }}}}(w)) + {{{\bar{\sigma }}}}(w)\frac{\partial F_\mathrm{stdNTS}^{-1}(\eta , \alpha , \theta , \beta )}{\partial \beta }\Big |_{\beta = {{{\bar{\beta }}}}(w)}\frac{\partial {{{\bar{\beta }}}}(w)}{\partial w_n} \end{aligned}$$
(17)

By substituting (16) into (17), we obtain (4).

Since there is \(\delta >0\) such that \(|\phi _\varXi (-u+i\delta )|<\infty \) for all \(u\in {{\mathbb {R}}}\), we have

$$\begin{aligned} \frac{\partial \mathrm{CVaR}_\eta (\varXi )}{\partial \beta }\Big |_{\beta ={{{\bar{\beta }}}}(w)}=\frac{\partial }{\partial \beta }\mathrm{CVaR}_\mathrm{stdNTS}(\eta , \alpha , \theta , \beta )\Big |_{\beta ={{{\bar{\beta }}}}(w)}. \end{aligned}$$

By (3), we have

$$\begin{aligned}&\frac{\partial }{\partial \beta }\mathrm{CVaR}_\mathrm{stdNTS}(\eta , \alpha , \theta , \beta ) \\&\quad =-\frac{\partial }{\partial \beta }F_\mathrm{stdNTS}^{-1}(u,\alpha ,\theta , \beta ) \\&\qquad -\frac{1}{\pi \eta }\mathrm{Re} \int _0^\infty e^{(iu+\delta )F_\mathrm{stdNTS}^{-1}(u,\alpha ,\theta , \beta )}\frac{\phi _\mathrm{stdNTS}(-u+i\delta , \alpha , \theta , \beta )}{(-u+i\delta )^2}\\&\times \left( (\delta +iu)\frac{\partial }{\partial \beta }F_\mathrm{stdNTS}^{-1}(u,\alpha ,\theta , \beta ) +\frac{\partial }{\partial \beta }\log \phi _\mathrm{stdNTS}(-u+i\delta , \alpha , \theta , \beta )\right) du. \end{aligned}$$

By setting \(\psi (z, \alpha , \theta , \beta )=\frac{\partial }{\partial \beta }\log \phi _\mathrm{stdNTS}(z, \alpha , \theta , \beta )\), we can simplify

$$\begin{aligned} \frac{\partial }{\partial \beta }\mathrm{CVaR}_\mathrm{stdNTS}(\eta , \alpha , \theta , \beta )&=-\frac{\partial }{\partial \beta }F_\mathrm{stdNTS}^{-1}(u,\alpha ,\theta , \beta ) \\&-\frac{1}{\pi \eta }\mathrm{Re} \int _0^\infty e^{(iu+\delta )F_\mathrm{stdNTS}^{-1}(u,\alpha ,\theta , \beta )}\frac{\phi _\mathrm{stdNTS}(-u+i\delta , \alpha , \theta , \beta )}{(-u+i\delta )^2}\\&\times \left( (\delta +iu)\frac{\partial }{\partial \beta }F_\mathrm{stdNTS}^{-1}(u,\alpha ,\theta , \beta ) +\psi (-u+i\delta , \alpha , \theta , \beta )\right) du. \end{aligned}$$

The characteristic function \(\phi _\mathrm{stdNTS}(u, \alpha , \theta , \beta )\) is equal to

$$\begin{aligned}&\phi _\mathrm{stdNTS}(u, \alpha , \theta , \beta ) \\&\quad =\exp \left( -\beta ui-\frac{2\theta ^{1-\frac{\alpha }{2}}}{\alpha } \left( \left( \theta -i\beta u+\left( 1-\frac{\beta ^2(2-\alpha )}{2\theta }\right) \frac{u^2}{2}\right) ^{\frac{\alpha }{2}}-\theta ^{\frac{\alpha }{2}}\right) \right) , \end{aligned}$$

hence we have

$$\begin{aligned}&\psi (z, \alpha , \theta , \beta ) \\&\quad =\frac{\partial }{\partial \beta } \left( -\beta zi-\frac{2\theta ^{1-\frac{\alpha }{2}}}{\alpha } \left( \left( \theta -i\beta z+\left( 1-\frac{\beta ^2(2-\alpha )}{2\theta }\right) \frac{z^2}{2}\right) ^{\frac{\alpha }{2}}-\theta ^{\frac{\alpha }{2}}\right) \right) \\&\quad =-zi +\left( 1-\frac{iz\beta }{\theta } +\left( 1-\frac{\beta ^2(2-\alpha )}{2\theta } \right) \frac{z^2}{2\theta } \right) ^{\frac{\alpha }{2}-1} \left( zi +\frac{\beta (2-\alpha )}{2\theta }z^2 \right) . \end{aligned}$$

As VaR case, CVaR for \(R_P(w)\) is calculated using \(\mathrm{CVaR}\eta (\varXi )\) that

$$\begin{aligned} \mathrm{CVaR}_\eta (R_p(w)) = -{{{\bar{\mu }}}}(w) +{{{\bar{\sigma }}}}(w) \mathrm{CVaR}_\eta (\varXi ). \end{aligned}$$

Therefore, we have

$$\begin{aligned} \frac{\partial \mathrm{CVaR}_\eta (R_p(w))}{\partial w_n}&= -\mu _n + \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n} \mathrm{CVaR}_\eta (\varXi ) + {{{\bar{\sigma }}}}(w)\frac{\partial \mathrm{CVaR}_\eta (\varXi )}{\partial w_n} \\&= -\mu _n + \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n} \mathrm{CVaR}_\eta (\varXi ) + {{{\bar{\sigma }}}}(w)\frac{\partial \mathrm{CVaR}_\eta (\varXi )}{\partial \beta }\Big |_{\beta ={{{\bar{\beta }}}}(w)}\frac{\partial {{{\bar{\beta }}}}(w)}{\partial w_n}, \end{aligned}$$

and, by (16), we obtain

$$\begin{aligned}&\frac{\partial \mathrm{CVaR}_\eta (R_p(w))}{\partial w_n} \nonumber \\&\quad = -\mu _n + \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n} \mathrm{CVaR}_\eta (\varXi ) + \left( \sigma _n\beta _n - {{{\bar{\beta }}}}(w) \frac{\partial {{{\bar{\sigma }}}}(w)}{\partial w_n}\right) \frac{\partial \mathrm{CVaR}_\eta (\varXi )}{\partial \beta }\Big |_{\beta = {{{\bar{\beta }}}}(w)} \end{aligned}$$
(18)

By substituting \(\mathrm{CVaR}_\eta (\varXi )=\mathrm{CVaR}_\mathrm{stdNTS}(\eta ,\alpha ,\theta ,{{{\bar{\beta }}}}(w))\) into (18), we obtain (5). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kim, Y.S. Portfolio optimization and marginal contribution to risk on multivariate normal tempered stable model. Ann Oper Res 312, 853–881 (2022). https://doi.org/10.1007/s10479-022-04613-7

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-022-04613-7

Keywords

Navigation