This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A discussion of stochastic dominance and mean-risk optimal portfolio problems based on mean-variance-mixture models

Hasanjan Sayit
Xi’an Jiaotong Liverpool University, Suzhou, China
(September 20, 2021)
Abstract

The classical Markowitz mean-variance model uses variance as a risk measure and calculates frontier portfolios in closed form by using standard optimization techniques. For general mean-risk models such closed form optimal portfolios are difficult to obtain. In this note, assuming returns follow the class of normal mean-variance mixture (NMVM) distributions, we obtain closed form expressions for frontier portfolios under mean-risk criteria when risk is modeled by the general class of law invariant convex risk measures. To achieve this goal, we first present a sufficient condition for the stochastic dominance relation on NMVM models and we apply this result to derive closed form solution for frontier portfolios. Our main result in this paper states that when return vectors follow the class of NMVM distributions the associated mean-risk frontier portfolios can be obtained by optimizing a Markowitz mean-variance model with an appropriately adjusted return vector.

Keywords: Frontier portfolios; Mean-variance mixtures; Risk measures; Stochastic dominance; Mean-risk criteria

JEL Classification: G11

1 Motivation

In risk management and portfolio optimization problems it is vital to model the underlying asset returns by proper distributions. The classical Markowitz portfolio optimization approach is based on the assumption that asset returns are normally distributed. However normality assumption of the asset returns is disputed extensively in the past literature. Numerous past studies on asset returns indicate that models based on normality assumption fail to fit into real-world data. As the seminal paper [7] demonstrates empirical densities of returns have more mass around the origin and in the tails and less in the flanks compared to the normal distributions. One of the noticeable feature of empirical asset returns is its asymmetry, a property of asset returns that is taken to mean skewness. Skewness of asset returns was recognized early on in modern finance theory and since then a great deal of effort was put on developing better models for asset returns with the hope that new models may lead to better financial decisions. It is readily apparent that for reliable financial decisions it is important to specify, as a starting point, a proper multivariate probability distribution for asset returns. Generalized hyperbolic distributions are flexible family of distributions that can capture most of empirical features of asset returns. As it was demonstrated in the paper [7] hyperbolic distributions give almost perfect fit for the empirical return densities. Hyperbolic distributions are subclass of the class of Normal mean-variance mixture (NMVM) models.

A dd-dimensional random vector XX has normal mean-variance mixture distribution with a mixing distribution ZZ if for any given Z=z+=:(0,+)Z=z\in\mathbb{R}_{+}=:(0,+\infty) the random vector XX satisfies X|Z=zWd(μ+zγ,zΣ)X|_{Z=z}\sim W_{d}(\mu+z\gamma,z\Sigma), where μ,γd\mu,\gamma\in\mathbb{R}^{d} are constant vectors, Σ\Sigma is a constant and positive-definite d×dd\times d matrix, and WdW_{d} is an nn-dimensional normal random vector with mean μ+zγ\mu+z\gamma and co-variance matrix zΣz\Sigma. It is well known that such random vector XX satisfies

X=𝑑μ+γZ+ZANd,X\overset{d}{=}\mu+\gamma Z+\sqrt{Z}AN_{d}, (1)

where A=Σ12A=\Sigma^{\frac{1}{2}} and NdN_{d} is an dd-dimensional normal random vector with mean equal to zero vector in d\mathbb{R}^{d} and with co-variance matrix equal to the identity matrix IdI_{d} in d×d\mathbb{R}^{d}\times\mathbb{R}^{d}. The mixing distribution ZZ in (1) is independent from NdN_{d}.

There are quite popular models within (1) in financial modeling. If ZZ follows a Generalized Inverse Gaussian (GIG) distribution, the distribution of XX follows a multi-dimensional generalized hyperbolic (mGH) distribution. This class of distributions were first introduced in the paper [2] and their numerous financial applications were discussed in the subsequent papers [1], [6], [7], [21], [22], [9], and [8]. For example, the paper [21] demonstrates that the variance-gamma model and its elliptical multivariate generalization can describe asset returns in unite time period exceptionally well. The paper [7] shows that hyperbolic distributions can give almost perfect fit to the empirical density functions of return data. The paper [1] finds that the Generalized Hyperbolic Skew Student’s tt distribution matches empirical financial data exceptionally well. In the multivariate financial data setting, the paper [24] calibrates the multivariate generalized hyperbolic (mGH) model to both multi-variate stock and multi-variate exchange rate returns and in their likelihood-ratio test the Gaussian model is always rejected against the general mGH models. The paper [1] applies the multivariate NIG (Normal Inverse Gaussian) distributions in risk management and demonstrates that they provide a much better fit to the empirical distribution of hedge fund returns than the Normal distribution. Generalized hyperbolic distributions were also applied in pricing interest rate derivatives and it was demonstrated that they allow for very accurate pricing of caplets and other interest rate derivatives, see for example [9] and [8].

In this paper, we consider a market with dd risky assets and we assume that the return vector XX of these dd risky assets satisfy (1). The portfolio set is given by d\mathbb{R}^{d} and for each portfolio ωd\omega\in\mathbb{R}^{d} the corresponding portfolio return is given by ωTX\omega^{T}X, where TT denotes a transpose. Our main goal of this paper is to study the following problem

minωρ(ωTX),s.t.E(ωTX)=r,ωTe=1,\begin{split}\min_{\omega}\;&\rho(-\omega^{T}X),\\ s.t.\;&E(-\omega^{T}X)=r,\\ &\omega^{T}e=1,\\ \end{split} (2)

for any given expected return level rr\in\mathbb{R}. In the optimization problem (2), ρ\rho is a risk measure and ede\in\mathbb{R}^{d} is the dd-dimensional vector of ones. In our note we will provide closed form solution to the problem (2) for risk measures ρ\rho that satisfy certain conditions that will be discussed in Section 3 below. As one of the conditions on ρ\rho we requite that ρ\rho is law-invariant, i.e., ρ(η1)=ρ(η2)\rho(\eta_{1})=\rho(\eta_{2}) as long as η1=𝑑η2\eta_{1}\overset{d}{=}\eta_{2}. Since

ωTX=𝑑ωTμ+ωTγZ+ZωTΣωN(0,1),\omega^{T}X\overset{d}{=}\omega^{T}\mu+\omega^{T}\gamma Z+\sqrt{Z}\omega^{T}\Sigma\omega N(0,1),

where N(0,1)N(0,1) is a standard normal random variable independent from ZZ, we have

ρ(ωTX)=ρ(ωTμωTγZZωTΣωN(0,1)).\rho(-\omega^{T}X)=\rho(-\omega^{T}\mu-\omega^{T}\gamma Z-\sqrt{Z}\omega^{T}\Sigma\omega N(0,1)).

We also require that ρ\rho is consistent with second order stochastic dominance relation. Therefore, as we shall see, the discussion of our problem (2) in fact reduces to studying stochastic dominance relations among one dimensional NMVM models.

We assume that all the random variables in this paper are defined in an atomless probability space (Ω,,P)(\Omega,\mathcal{F},P). We denote by L0=L0(Ω,,P)L^{0}=L^{0}(\Omega,\mathcal{F},P) the space of all finite valued random variables and by Lk=Lk(Ω,,P),1k<+,L^{k}=L^{k}(\Omega,\mathcal{F},P),1\leq k<+\infty, the space of random variables with finite kk moments. L(Ω,,P)L^{\infty}(\Omega,\mathcal{F},P) denotes the space of bounded random variables. For vectors x=(x1,x2,,xd)Tx=(x_{1},x_{2},\cdots,x_{d})^{T} and y=(y1,y2,,yd)Ty=(y_{1},y_{2},\cdots,y_{d})^{T} we denote by xy=xTy=i=1dxiyix\cdot y=x^{T}y=\sum_{i=1}^{d}x_{i}y_{i} the scalar product of them and by x=(x12++xd2)12||x||=(x_{1}^{2}+\cdots+x_{d}^{2})^{\frac{1}{2}} the Euclidean norm of xx. As in [10], for each positive integer dd we denote by Ldk=(Lk)dL_{d}^{k}=(L^{k})^{d} the space of all dd-dimensional random vectors with all the components belong to LkL^{k}. In this paper, unless otherwise stated we assume that the mixing distribution satisfies ZL1Z\in L^{1}. We also assume that ZZ has probability density function fZ(z)f_{Z}(z). This latter assumption on ZZ is in fact is not necessary for our results of this paper. However, we impose this assumption to keep our calculations simple. We use ϕ()\phi(\cdot) to denote the density function of standard normal random variable and by Φ()\Phi(\cdot) its cumulative distribution function throughout the paper.

Our initial motivation in studying the problem (2) was to discuss solutions for (2) under the widely used and popular risk measures VaRVaR and CVaRCVaR. The value-at-risk (VaR) is defined for any random variable in L0L^{0}. For any ηL0\eta\in L^{0}, denote by qα1(η)=inf{a:Fη(a)α}q_{\alpha}^{-1}(\eta)=\inf\{a\in\mathbb{R}:F_{\eta}(a)\geq\alpha\} the left-continuous inverse of the cumulative distribution function Fη()F_{\eta}(\cdot) of η\eta. For each α(0,1)\alpha\in(0,1) the VaRαVaR_{\alpha} is defined as VaRα(η)=qα1(η)VaR_{\alpha}(\eta)=q_{\alpha}^{-1}(\eta). The conditional value-at-risk (CVaR) is defined on L1L^{1} as CVaRα(η)=(1/(1α))α1qβ(η)𝑑βCVaR_{\alpha}(\eta)=(1/(1-\alpha))\int_{\alpha}^{1}q_{\beta}(\eta)d\beta for any ηL1\eta\in L^{1}. In our note we are able to provide closed form solution for (2) when ρ=CVaRα\rho=CVaR_{\alpha} while an analytical solution for (2) when ρ=VaRα\rho=VaR_{\alpha} seems difficult. The main reason for this is the risk measure CVaRαCVaR_{\alpha} is consistent with the second order stochastic dominance property. Another important property of CVaR that is essential for our discussions in this paper is its continuity in the LkL^{k} space for any k1k\geq 1, see Theorem 4.1 of [17] and page 11 of [10] and the references there for this. In fact, while we don’t use this property, it is also known that CVaR is Lipschitz continuous with respect to LkL^{k} convergence for 1k1\leq k\leq\infty, see Proposition 4.4 in [11] for this. As mentioned earlier, the risk measure VaR is well defined on L0L^{0} and, unlike CVaR, it is not a convex risk measure. It is consistent with the first order stochastic dominance property but it is not consistent with the second order stochastic dominance property. Due to these facts, optimization problems associated with it is challenging.

The paper is organized as follows. In Section 2 below we formulate a sufficient condition for the stochastic dominance relation for NMVM models. Then we use it to derive closed form solutions for frontier portfolios. In Section 3 we study high-order stochastic dominance relations among NMVM models.

2 Frontier portfolios

In this section we discuss solutions of the problem (2). For the simplicity of our discussions we first fix a positive integer k1k\geq 1 and we consider risk measures ρ:Lk{}\rho:L^{k}\rightarrow\mathbb{R}\cup\{\infty\} for the problem (2). Denoting by dom(ρ)={ηLk:ρ(η)<}dom(\rho)=\{\eta\in L^{k}:\rho(\eta)<\infty\} the domain of ρ\rho, we call ρ\rho proper if ρ(η)>\rho(\eta)>-\infty for all ηLk\eta\in L^{k} and dom(ρ)dom(\rho)\neq\emptyset. A risk measure on LkL^{k} is called convex risk measure if it is a proper monetary risk measure (i.e., it is monotone and cash-invariant) which is convex, i.e., ρ(λη1+(1λ)η2)λρ(η1)+(1λ)ρ(η2),λ[0,1],η1,η2Lk\rho(\lambda\eta_{1}+(1-\lambda)\eta_{2})\leq\lambda\rho(\eta_{1})+(1-\lambda)\rho(\eta_{2}),\forall\lambda\in[0,1],\forall\eta_{1},\eta_{2}\in L^{k}. A convex risk measure is called coherent if it is positive homogeneous, i.e., ρ(λη)=λρ(η),λ>0\rho(\lambda\eta)=\lambda\rho(\eta),\forall\lambda>0. From Corollary 2.3 of [17] any finite convex risk measure on LkL^{k} is continuous in LkL^{k}. The risk measure CVaRCVaR is a coherent risk measure on LkL^{k} and hence it is continuous on LkL^{k}. This property of CVaRCVaR is useful in the proof of the Proposition 2.14 below.

Another property of ρ\rho that is essential for our discussions in this section is its consistency with the second order stochastic dominance property. We say that ρ\rho is consistent with the second order stochastic dominance (SSD-consistent) property if for any two η1,η2Lk\eta_{1},\eta_{2}\in L^{k} the relation η1(2)η2\eta_{1}\succeq_{(2)}\eta_{2} implies that ρ(η1)ρ(η2)\rho(\eta_{1})\geq\rho(\eta_{2}). This property of ρ\rho is needed in the proof of our Theorem 2.4 below and an equivalent version of it is stated in the Lemma 2.13 below. We call ρ\rho is law-invariant if ρ(η1)=ρ(η2)\rho(\eta_{1})=\rho(\eta_{2}) whenever η1=𝑑η2\eta_{1}\overset{d}{=}\eta_{2}. Any finite valued, law-invariant, convex risk measure is SSD-consistent, see the paper [23] and the references there for this.

Remark 2.1.

We remark here that the solution for the problem (2) always exists when the risk measure ρ\rho is finite-valued, law-invariant, and convex risk measure. To see this, note first that the map ωρ(ωTξ)\omega\rightarrow\rho(-\omega^{T}\xi) is a continuous function as long as ξLdk\xi\in L_{d}^{k} as ρ\rho is continuous map on LkL^{k} as explained above. On the other hand, for any integrable random variable η\eta one has Eη(2)ηE\eta\succeq_{(2)}\eta and hence ρ(η)ρ(Eη)\rho(-\eta)\geq\rho(-E\eta), see Appendix B of [14] and Corollary 5.1 of [12] for similar arguments. Therefore on the set {ωd:E(ωTX)=r,ωTe=1}\{\omega\in\mathbb{R}^{d}:E(-\omega^{T}X)=r,\omega^{T}e=1\} we have r(2)ωTX-r\succeq_{(2)}\omega^{T}X which implies

ρ(ωTX)ρ(r)=r+ρ(0).\rho(-\omega^{T}X)\geq\rho(r)=r+\rho(0).

Before we state our main result in this section, we first give a short review for the Markovitz mean-variance portfolio optimization problem. In the Markowitz mean-variance portfolio optimization framework the variance is used as a risk measure. In this case, the corresponding optimization problem is

minωVar(ωTX),E(ωTX)=r,ωTe=1.\begin{split}\min_{\omega}Var(-\omega^{T}X),&\\ E(-\omega^{T}X)=r,&\\ \omega^{T}e=1.&\\ \end{split} (3)

The closed form solution of (3) is standard and can be found in any standard textbooks that discusses the capital asset pricing model, see page 64 of [16] for example. Here we write down the solution of (3) as it is relevant with our main results in this paper. To this end, let μ=EX\mu=EX denote the mean vector and let V=Cov(X)V=Cov(X) denote the co-variance matrix of XX. For each rr\in\mathbb{R}, the solution of (3) is given by

ωr=ωr(μ,V)=1d4[d2(V1e)d1(V1μ)]+rd4[d3(V1μ)d1(V1e)],\omega^{\star}_{r}=\omega^{\star}_{r}(\mu,V)=\frac{1}{d^{4}}[d^{2}(V^{-1}e)-d^{1}(V^{-1}\mu)]+\frac{r}{d^{4}}[d^{3}(V^{-1}\mu)-d^{1}(V^{-1}e)], (4)

where

d1=eTV1μ,d2=μTV1μ,d3=eTV1e,d4=d2d3(d1)2.d^{1}=e^{T}V^{-1}\mu,\;\;d^{2}=\mu^{T}V^{-1}\mu,\;\;d^{3}=e^{T}V^{-1}e,\;\;d^{4}=d^{2}d^{3}-(d^{1})^{2}.
Remark 2.2.

Note that since V1V^{-1} is positive definite we have (d1μd2e)TV1(d1μd2e)>0\large(d^{1}\mu-d^{2}e\large)^{T}V^{-1}\large(d^{1}\mu-d^{2}e\large)>0 and d2>0d^{2}>0 (we assume μ=EX0\mu=EX\neq 0 in this paper). Since

(d1μd2e)TV1(d1μd2e)=d2(d2d3(d1)2)=d2d4>0,\large(d^{1}\mu-d^{2}e\large)^{T}V^{-1}\large(d^{1}\mu-d^{2}e\large)=d^{2}(d^{2}d^{3}-(d^{1})^{2})=d^{2}d^{4}>0,

we have d4>0d^{4}>0 in (4).

In this section, we will show that the optimal portfolio in our mean-risk portfolio optimization problem (2) also takes a similar form as in (4). Therefore, we first introduce the following notations. For any random vector θ\theta with mean vector μθ=Eθ\mu_{\theta}=E\theta and co-variance matrix Σθ=Cov(θ)\Sigma_{\theta}=Cov(\theta) we introduce the following expression

ωθ=ωθ(μθ,Σθ)=1dθ4[dθ2(Σθ1e)dθ1(Σθ1μθ)]+rdθ4[dθ3(Σθ1μθ)dθ1(Σθ1e)],\omega^{\star}_{\theta}=\omega^{\star}_{\theta}(\mu_{\theta},\Sigma_{\theta})=\frac{1}{d_{\theta}^{4}}[d_{\theta}^{2}(\Sigma_{\theta}^{-1}e)-d_{\theta}^{1}(\Sigma_{\theta}^{-1}\mu_{\theta})]+\frac{r}{d_{\theta}^{4}}[d_{\theta}^{3}(\Sigma_{\theta}^{-1}\mu_{\theta})-d_{\theta}^{1}(\Sigma_{\theta}^{-1}e)], (5)

where

dθ1=eTΣθ1μθ,dθ2=μθTΣθ1μθ,dθ3=eTΣθ1e,dθ4=dθ2dθ3(dθ1)2.d_{\theta}^{1}=e^{T}\Sigma_{\theta}^{-1}\mu_{\theta},\;\;d_{\theta}^{2}=\mu_{\theta}^{T}\Sigma_{\theta}^{-1}\mu_{\theta},\;\;d_{\theta}^{3}=e^{T}\Sigma_{\theta}^{-1}e,\;\;d_{\theta}^{4}=d_{\theta}^{2}d_{\theta}^{3}-(d_{\theta}^{1})^{2}.
Remark 2.3.

The right-hand-side of (5) only involves the mean vector μθ\mu_{\theta} and the co-variance matrix Σθ\Sigma_{\theta} of the random vector θ\theta. Therefore ωθ=ωη\omega^{\star}_{\theta}=\omega^{\star}_{\eta} as long as θ\theta and η\eta have the same mean vectors and co-variance matrices. Here we used the random variable θ\theta in the definition of ωθ\omega^{\star}_{\theta} to indicate that the portfolio ωθ\omega_{\theta}^{\star} is the frontier portfolio under the mean-variance criteria with return vector θ\theta. Note that the above expression (5) is well defined as long as μθ0\mu_{\theta}\neq 0 and Σθ\Sigma_{\theta} is positive definite (all the eigenvalues are strictly positive numbers) as explained in Remark 2.2 above.

Our Theorem 2.4 below shows that, frontier portfolios to the problem (2), under any real-valued, law invariant, and convex risk measure, always exist and they are in the form as in (5) when the return vectors are given by (1).

Theorem 2.4.

Fix a positive integer k1k\geq 1. Let ρ\rho be any finite-valued, law-invariant, convex risk measure on LkL^{k} . Let the return vector be given by (1) with ZLkZ\in L^{k}. Then the solution of the problem (2) for each fixed real number rr is given by ωθ\omega^{\star}_{\theta} as in (5) with any random vector θ\theta with mean μθ=μ+γEZ\mu_{\theta}=\mu+\gamma EZ and co-variance matrix Σθ=Σ\Sigma_{\theta}=\Sigma.

Proof.

Denote Dr={ωRd:ωT(μ+γEZ)=r,ωTe=1}D_{r}=\{\omega\in R^{d}:\omega^{T}(\mu+\gamma EZ)=r,\omega^{T}e=1\}. On the set DrD_{r} the minimizing point of the quadratic function ωTΣω\omega^{T}\Sigma\omega is given by ω¯=:ωθ(μ+γEZ,Σ)\bar{\omega}=:\omega_{\theta}(\mu+\gamma EZ,\Sigma) as in (5). From Proposition 2.14 below, for any ωDr\omega\in D_{r} we have ω¯TX(2)ωTX\bar{\omega}^{T}X\succeq_{(2)}\omega^{T}X. The conditions on ρ\rho stated in the Theorem imply that ρ\rho is SSD-consistent. Hence we have ρ(ω¯TX)ρ(ωTX)\rho(-\bar{\omega}^{T}X)\leq\rho(-\omega^{T}X) for all ωDr\omega\in D_{r}. This completes the proof. ∎∎

Remark 2.5.

Observe that μX=EX=μ+γEZ=μθ\mu_{X}=EX=\mu+\gamma EZ=\mu_{\theta}. However the covariance matrix Cov(X)=γγTVar(Z)+ATAEZCov(X)=\gamma\gamma^{T}Var(Z)+A^{T}AEZ is different from Σθ=ATA\Sigma_{\theta}=A^{T}A. Therefore, the optimal solution of (3) is different from the optimal solution of (2). But these two solutions are similar in the sense that they differ only in the co-variance matrices ΣX\Sigma_{X} and Σθ\Sigma_{\theta}.

Remark 2.6.

The message of the Theorem 2.4 above is that the mean-risk frontier portfolios under any real-valued law invariant convex risk measure ρ\rho for return vectors XX as in (1), can be obtained by solving a Markowitz mean-variance optimal portfolio problem with an appropriately adjusted return vector θ\theta as in the Theorem 2.4 above.

The risk measure CVaRαCVaR_{\alpha} for each fixed α(0,1)\alpha\in(0,1) is a finite valued, law-invariant, convex risk measure. Therefore the result of the above Theorem 2.4 holds for this risk measure also. We state this in the following Corollary.

Corollary 2.7.

Let XX be a return vector given by (1) with ZLkZ\in L^{k} for some positive integer kk. Then for each fixed α(0,1)\alpha\in(0,1), the closed form solution of the following optimization problem

minωCVaRα(ωTX),s.t.E(ωTX)=r,ωTe=1,\begin{split}\min_{\omega}\;&CVaR_{\alpha}(-\omega^{T}X),\\ s.t.\;&E(-\omega^{T}X)=r,\\ &\omega^{T}e=1,\\ \end{split} (6)

is given by ωθ\omega_{\theta}^{\star} as in (5) with any random vector θ\theta with μθ=EY=μ+γEZ\mu_{\theta}=EY=\mu+\gamma EZ and Σθ=ATA\Sigma_{\theta}=A^{T}A.

Remark 2.8.

We remark that our Theorem 2.4 above can be applied to simplify the calculations of the optimal values of CVaRαCVaR_{\alpha} on some domains of the portfolios when returns are normal, a topic which was discussed in Theorem 2 of [28] (see also [29]). To this end, first observe that if Z=1Z=1 in (1), XX is a Normal random vector. We denote this Normal random vector by XNd(γ,Σ)X\sim N_{d}(\gamma,\Sigma) (taking μ=0\mu=0) for notations simplicity in the following discussions. Observe that with all the other model parameters γ\gamma and Σ\Sigma are fixed, the expression (5) becomes a function of the expected return level rr. Therefore for convenience we denote ωr:=ωθ\omega_{r}:=\omega_{\theta}^{\star} for θ=𝑑Nd(γ,Σ)\theta\overset{d}{=}N_{d}(\gamma,\Sigma) in expression (5). We write this ωr\omega_{r} in a different form as

ωr=k1(r)Σ1e+k2(r)Σ1γ,\omega_{r}=k_{1}(r)\Sigma^{-1}e+k_{2}(r)\Sigma^{-1}\gamma,

where

k1(r)=dθ2/dθ4rdθ1/dθ4,k2(r)=rdθ3/dθ4dθ1/dθ4.k_{1}(r)=d_{\theta}^{2}/d_{\theta}^{4}-rd_{\theta}^{1}/d_{\theta}^{4},\;\;k_{2}(r)=rd_{\theta}^{3}/d_{\theta}^{4}-d_{\theta}^{1}/d_{\theta}^{4}.

With these notations we have ωrTNnN(ωrTγ,σr2)\omega^{T}_{r}N_{n}\sim N(\omega_{r}^{T}\gamma,\sigma_{r}^{2}), where

σr2:=ωrTΣωr=k12(r)eTΣ1e+2k1(r)k2(r)eTΣ1γ+k22(r)γTΣ1γ.\sigma_{r}^{2}:=\omega_{r}^{T}\Sigma\omega_{r}=k_{1}^{2}(r)e^{T}\Sigma^{-1}e+2k_{1}(r)k_{2}(r)e^{T}\Sigma^{-1}\gamma+k_{2}^{2}(r)\gamma^{T}\Sigma^{-1}\gamma. (7)

From [27] (see also equation (2) of [19]), for a Normal random variable H=N(δ,σ2)H=N(\delta,\sigma^{2}) we have

CVaRα(H)=δ+[1σφ(zαδσ)/(1Φ(zαδσ))]σ2,CVaR_{\alpha}(H)=\delta+\Big{[}\frac{1}{\sigma}\varphi(\frac{z_{\alpha}-\delta}{\sigma})/(1-\Phi(\frac{z_{\alpha}-\delta}{\sigma}))\Big{]}\sigma^{2}, (8)

where zαz_{\alpha} is the α\alpha-quantile of the standard Normal random variable N(0,1)N(0,1). For any return level rr define

hα(r):=r+[1σrφ(zαrσr)/(1Φ(zαrσr))]σr2,h_{\alpha}(r):=r+\Big{[}\frac{1}{\sigma_{r}}\varphi(\frac{z_{\alpha}-r}{\sigma_{r}})/(1-\Phi(\frac{z_{\alpha}-r}{\sigma_{r}}))\Big{]}\sigma^{2}_{r},

where σr\sigma_{r} is given by (7). Now, consider the following domain

D={ω:ωTγr¯},D=\{\omega:\omega^{T}\gamma\geq\bar{r}\},

for some fixed real number r¯\bar{r}. Define Dγ:={ωTγ:ωD}D_{\gamma}:=\{\omega^{T}\gamma:\omega\in D\}. Observe that the domain DD has the property that for any given r0Dγr_{0}\in D_{\gamma}, any portfolio ω\omega with ωTγ=r0\omega^{T}\gamma=r_{0} lies in DD, i.e., ωD\omega\in D. On this domain, we clearly have

minωDCVaRα(ωTθ)=minrDγhα(r),min_{\omega\in D}CVaR_{\alpha}(\omega^{T}\theta)=min_{r\in D_{\gamma}}h_{\alpha}(r), (9)

where θ=𝑑Nd(γ,Σ)\theta\overset{d}{=}N_{d}(\gamma,\Sigma). Note that the right-hand-side of (9) is a minimization problem of a real valued function on a subset of the real line which is a simpler problem than finding the minimum value of Fα(ω,a)F_{\alpha}(\omega,a) in (4) of [28] as it is done in Theorem 2 of the same paper.

Remark 2.9.

Since the risk measure VaRα()VaR_{\alpha}(\cdot) is not SSD-consistent, a similar result as in Corollary 2.7 for VaRα()VaR_{\alpha}(\cdot) is difficult to construct. The following relation is well known.

CVaRα(η)=VaRα(η)+11αE(ηqα1(η))+,CVaR_{\alpha}(\eta)=VaR_{\alpha}(\eta)+\frac{1}{1-\alpha}E(\eta-q_{\alpha}^{-1}(\eta))^{+},

where E(ηqα1(η))+E(\eta-q_{\alpha}^{-1}(\eta))^{+} is referred to as the expected shortfall at level α\alpha. Clearly we have CVaRαVaRαCVaR_{\alpha}\geq VaR_{\alpha} (as stated in the paragraph that contains (P2) in [28] also). The above relation does not tell much about the frontier portfolios associated with the risk measure VaRαVaR_{\alpha}. It is not clear if the frontier portfolios under CVaRαCVaR_{\alpha} are close (in the Euclidean norm) to the frontier portfolios under VaRαVaR_{\alpha} due to the relation CVaRαVaRαCVaR_{\alpha}\geq VaR_{\alpha}.

The proof of the Theorem 2.4 above needs some preparation. We first prove the following Lemma. First, observe that from (24) in Section 3 for any H=a+bZ+cZNH=a+bZ+c\sqrt{Z}N we have

E(kH)+=0+czϕ(2)(kabzcz)f(z)𝑑z.E(k-H)^{+}=\int_{0}^{+\infty}c\sqrt{z}\phi^{(2)}(\frac{k-a-bz}{c\sqrt{z}})f(z)dz. (10)

Define

Ψ=ΨH(u,c)=:0+czϕ(2)(ucz)f(z)dz.\Psi=\Psi_{H}(u,c)=:\int_{0}^{+\infty}c\sqrt{z}\phi^{(2)}(\frac{u}{c\sqrt{z}})f(z)dz.

It is easy to check that

dΨdu=0+ϕ(1)(ucz)f(z)𝑑z>0,dΨdc=0+zϕ(0)(ucz)f(z)𝑑z>0.\frac{d\Psi}{du}=\int_{0}^{+\infty}\phi^{(1)}(\frac{u}{c\sqrt{z}})f(z)dz>0,\;\;\frac{d\Psi}{dc}=\int_{0}^{+\infty}\sqrt{z}\phi^{(0)}(\frac{u}{c\sqrt{z}})f(z)dz>0.

Therefore ΨH(u,c)\Psi_{H}(u,c) is an increasing function of uu and also of cc. We use these facts in the proof of the following Lemma.

Lemma 2.10.

Let ZZ be any random variable that satisfies βZα>0\beta\geq Z\geq\alpha>0 for some finite numbers α\alpha and β\beta. Denote η=EZ\eta=EZ and consider the NMVM models Q=a1+b1Z+c1ZNQ=a_{1}+b_{1}Z+c_{1}\sqrt{Z}N and R=a2+b2Z+c2ZNR=a_{2}+b_{2}Z+c_{2}\sqrt{Z}N, where a1,a2,b1,b2a_{1},a_{2},b_{1},b_{2} are any real numbers and c1>0,c2>0c_{1}>0,\;c_{2}>0. Then we have Q(2)RQ\succeq_{(2)}R if and only if a1+b1EZa2+b2EZa_{1}+b_{1}EZ\geq a_{2}+b_{2}EZ and c1c2c_{1}\leq c_{2}.

Proof.

If Q(2)RQ\succeq_{(2)}R then a1+b1EZa2+b2EZa_{1}+b_{1}EZ\geq a_{2}+b_{2}EZ follows from Theorem 1 of [26] and c1c2c_{1}\leq c_{2} follows from Lemma 3.13 above. To show the other direction we need to show I1=:E(kQ)+I2=:E(kR)+I_{1}=:E(k-Q)^{+}\leq I_{2}=:E(k-R)^{+} for any real number kk. From (10) we have

I1=ηβc1zϕ(2)(ka1b1zc1z)f(z)𝑑z+αηc1zϕ(2)(ka1b1zc1z)f(z)𝑑z,I2=ηβc2zϕ(2)(ka2b2zc2z)f(z)𝑑z+αηc2zϕ(2)(ka2b2zc2z)f(z)𝑑z.\begin{split}I_{1}=&\int_{\eta}^{\beta}c_{1}\sqrt{z}\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{1}\sqrt{z}})f(z)dz+\int_{\alpha}^{\eta}c_{1}\sqrt{z}\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{1}\sqrt{z}})f(z)dz,\\ I_{2}=&\int_{\eta}^{\beta}c_{2}\sqrt{z}\phi^{(2)}(\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}})f(z)dz+\int_{\alpha}^{\eta}c_{2}\sqrt{z}\phi^{(2)}(\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}})f(z)dz.\\ \end{split}

We have

I2I1=ηβc2z[ϕ(2)(ka2b2zc2z)ϕ(2)(ka1b1zc2z)]f(z)𝑑z+ηβc2z[ϕ(2)(ka1b1zc2z)c1zϕ(2)(ka1b1zc1z)]f(z)𝑑z+αηc2z[ϕ(2)(ka2b2zc2z)ϕ(2)(ka1b1zc2z)]f(z)𝑑z+αηc2z[ϕ(2)(ka1b1zc2z)c1zϕ(2)(ka1b1zc1z)]f(z)𝑑z.ηβc2z[ϕ(2)(ka2b2zc2z)ϕ(2)(ka1b1zc2z)]f(z)𝑑z+αηc2z[ϕ(2)(ka2b2zc2z)ϕ(2)(ka1b1zc2z)]f(z)𝑑z.\begin{split}I_{2}-I_{1}=&\int_{\eta}^{\beta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}})-\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})]f(z)dz\\ +&\int_{\eta}^{\beta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})-c_{1}\sqrt{z}\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{1}\sqrt{z}})]f(z)dz\\ +&\int_{\alpha}^{\eta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}})-\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})]f(z)dz\\ +&\int_{\alpha}^{\eta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})-c_{1}\sqrt{z}\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{1}\sqrt{z}})]f(z)dz.\\ \geq&\int_{\eta}^{\beta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}})-\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})]f(z)dz\\ +&\int_{\alpha}^{\eta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}})-\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})]f(z)dz.\\ \end{split}

The above inequality follows because the two terms

ηβc2z[ϕ(2)(ka1b1zc2z)c1zϕ(2)(ka1b1zc1z)]f(z)𝑑z\int_{\eta}^{\beta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})-c_{1}\sqrt{z}\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{1}\sqrt{z}})]f(z)dz

and

αηc2z[ϕ(2)(ka1b1zc2z)c1zϕ(2)(ka1b1zc1z)]f(z)𝑑z\int_{\alpha}^{\eta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})-c_{1}\sqrt{z}\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{1}\sqrt{z}})]f(z)dz

are positive numbers as the function Ψ(μ,c)\Psi(\mu,c) is an increasing function in the argument cc as explained in the paragraph preceding to this Lemma. We need to show that the term

I=:ηβc2z[ϕ(2)(ka2b2zc2z)ϕ(2)(ka1b1zc2z)]f(z)𝑑z+αηc2z[ϕ(2)(ka2b2zc2z)ϕ(2)(ka1b1zc2z)]f(z)𝑑z.\begin{split}I=:&\int_{\eta}^{\beta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}})-\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})]f(z)dz\\ +&\int_{\alpha}^{\eta}c_{2}\sqrt{z}[\phi^{(2)}(\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}})-\phi^{(2)}(\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}})]f(z)dz.\end{split}

is a positive number also. By middle value Theorem we have

I=ηβϕ(1)(1(z))[(a1a2)+(b1b2)z]f(z)𝑑z+αηϕ(1)(2(z))[(a1a2)+(b1b2)z]f(z)𝑑z,\begin{split}I=&\int_{\eta}^{\beta}\phi^{(1)}(\ell_{1}(z))[(a_{1}-a_{2})+(b_{1}-b_{2})z]f(z)dz\\ +&\int_{\alpha}^{\eta}\phi^{(1)}(\ell_{2}(z))[(a_{1}-a_{2})+(b_{1}-b_{2})z]f(z)dz,\end{split} (11)

where 1(z)\ell_{1}(z) is between ka2b2zc2z\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}} and ka1b1zc2z\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}} when z[η,β]z\in[\eta,\beta] and also 1(z)\ell_{1}(z) is between ka2b2zc2z\frac{k-a_{2}-b_{2}z}{c_{2}\sqrt{z}} and ka1b1zc2z\frac{k-a_{1}-b_{1}z}{c_{2}\sqrt{z}} when z[α,η]z\in[\alpha,\eta]. Observe that without the terms ϕ(1)(1(z))\phi^{(1)}(\ell_{1}(z)) and ϕ(1)(2(z))\phi^{(1)}(\ell_{2}(z)) in (11) we would have

ηβ[(a1a2)+(b1b2)z]f(z)𝑑z+αη[(a1a2)+(b1b2)z]f(z)𝑑z=(a1a2)+(b1b2)EZ0.\begin{split}&\int_{\eta}^{\beta}[(a_{1}-a_{2})+(b_{1}-b_{2})z]f(z)dz+\int_{\alpha}^{\eta}[(a_{1}-a_{2})+(b_{1}-b_{2})z]f(z)dz\\ &=(a_{1}-a_{2})+(b_{1}-b_{2})EZ\geq 0.\end{split}

Now, define g(z)=(a1a2)+(b1b2)zg(z)=(a_{1}-a_{2})+(b_{1}-b_{2})z. Then g(z)g(z) is an increasing or decreasing linear function of zz depending on b1b2b_{1}\geq b_{2} or b1b2b_{1}\leq b_{2}. Therefore either g(z)0g(z)\geq 0 on [η,β][\eta,\beta] and g(z)0g(z)\leq 0 on [α,η][\alpha,\eta] or g(z)0g(z)\leq 0 on [η,β][\eta,\beta] and g(z)0g(z)\geq 0 on [α,η][\alpha,\eta]. If g(z)0g(z)\geq 0 on [η,β][\eta,\beta] and g(z)0g(z)\leq 0 on [α,η][\alpha,\eta], then 1(z)2(z)\ell_{1}(z)\geq\ell_{2}(z) and therefore ϕ(1)(1(z))ϕ(1)(2(z))\phi^{(1)}(\ell_{1}(z))\geq\phi^{(1)}(\ell_{2}(z)). This means that the function g(z)g(z) is multiplied by a larger valued function ϕ(1)(1(z)\phi^{(1)}(\ell_{1}(z) when it is positive valued in [η,β][\eta,\beta] compared to it is multiplied by a smaller valued function ϕ(1)(1(z)\phi^{(1)}(\ell_{1}(z) while it takes negative values in [α,η][\alpha,\eta]. This shows that we have

ηβϕ(1)(1(z))[(a1a2)+(b1b2)z]f(z)𝑑z|αηϕ(1)(2(z))[(a1a2)+(b1b2)z]f(z)𝑑z|.\int_{\eta}^{\beta}\phi^{(1)}(\ell_{1}(z))[(a_{1}-a_{2})+(b_{1}-b_{2})z]f(z)dz\geq|\int_{\alpha}^{\eta}\phi^{(1)}(\ell_{2}(z))[(a_{1}-a_{2})+(b_{1}-b_{2})z]f(z)dz|.

This in turn implies I0I\geq 0. On the other-hand if g(z)0g(z)\leq 0 on [η,β][\eta,\beta] and g(z)0g(z)\geq 0 on [α,η][\alpha,\eta] then 1(z)2(z)\ell_{1}(z)\leq\ell_{2}(z) and therefore ϕ(1)(1(z))ϕ(1)(2(z))\phi^{(1)}(\ell_{1}(z))\leq\phi^{(1)}(\ell_{2}(z)) which implies

αηϕ(1)(2(z))[(a1a2)+(b1b2)z]f(z)𝑑z|ηβϕ(1)(1(z))[(a1a2)+(b1b2)z]f(z)𝑑z|.\int_{\alpha}^{\eta}\phi^{(1)}(\ell_{2}(z))[(a_{1}-a_{2})+(b_{1}-b_{2})z]f(z)dz\geq|\int_{\eta}^{\beta}\phi^{(1)}(\ell_{1}(z))[(a_{1}-a_{2})+(b_{1}-b_{2})z]f(z)dz|.

This shows I0I\geq 0. This completes the proof. ∎

Remark 2.11.

We should mention that the assumption on ZZ being bounded mixing random variable in the above Lemma 2.10 is needed as it guarantees that the functions kaibizciz,i=1,2,\frac{k-a_{i}-b_{i}z}{c_{i}\sqrt{z}},i=1,2, in the proof of this Lemma are bounded functions. Since limx+ϕ(2)(x)+\lim_{x\rightarrow+\infty}\phi^{(2)}(x)\rightarrow+\infty, boundedness of these functions are needed in the proof of this Lemma.

Remark 2.12.

The above Lemma 2.10 shows that when the mixing random variable is bounded ZLZ\in L^{\infty} and when

a1+b1EZ=a2+b2EZ,a_{1}+b_{1}EZ=a_{2}+b_{2}EZ,

we have Q(2)RQ\succeq_{(2)}R if and only if c1c2c_{1}\leq c_{2}.

Before we state our next result we first recall few definitions. Recall that a random variable η1\eta_{1} second order stochastically dominates another random variable η2\eta_{2}, i.e., η1(2)η2\eta_{1}\succeq_{(2)}\eta_{2}, iff t[Fη1(a)Fη2(a)]𝑑a0\int_{-\infty}^{t}[F_{\eta_{1}}(a)-F_{\eta_{2}}(a)]da\leq 0 for all tt\in\mathbb{R}. The dual of SSD is the risk-seeking stochastic dominance (RSSD) relation η1(2)η2\eta_{1}\succeq_{(2^{\prime})}\eta_{2} which is defined as t[Fη1(a)Fη2(a)]𝑑a0\int_{t}^{\infty}[F_{\eta_{1}}(a)-F_{\eta_{2}}(a)]da\leq 0 for all tt\in\mathbb{R}. Since we have t(Fη1(a)Fη2(a))𝑑a=t(Fη2(a)Fη1(a))𝑑a\int_{-\infty}^{t}(F_{\eta_{1}}(a)-F_{\eta_{2}}(a))da=\int_{-t}^{\infty}(F_{-\eta_{2}}(a)-F_{-\eta_{1}}(a))da it is clear that the relation η1(2)η2\eta_{1}\succeq_{(2)}\eta_{2} is equivalent to the relation η2(2)η1-\eta_{2}\succeq_{(2^{\prime})}-\eta_{1}. The risk measure CVaRCVaR is consistent with RSSD for any α(0,1)\alpha\in(0,1) in the sense that (See page 174 of [5] for this)

η2(2)η1CVaRα(η1)CVaRα(η2),α(0,1).\eta_{2}\succeq_{(2^{\prime})}\eta_{1}\Leftrightarrow CVaR_{\alpha}(\eta_{1})\leq CVaR_{\alpha}(\eta_{2}),\;\forall\alpha\in(0,1).

An important implication of this relation in our setting is stated in the following Lemma. This result can be found in (vi) of page 14 of [18], or in Theorem 4.A.3 of [30], or in [13].

Lemma 2.13.

For any two portfolios ω1,ω2Rd\omega_{1},\omega_{2}\in R^{d} we have

ω1TX(2)ω2TXCVaRα(ω1TX)CVaRα(ω2TX),α(0,1).\omega^{T}_{1}X\succeq_{(2)}\omega_{2}^{T}X\Leftrightarrow CVaR_{\alpha}(-\omega^{T}_{1}X)\leq CVaR_{\alpha}(-\omega^{T}_{2}X),\forall\alpha\in(0,1).

Next we state the following result.

Proposition 2.14.

Fix any positive integer k1k\geq 1. Assume ZLkZ\in L^{k} and consider the following two NMVM models Q=a1+b1Z+c1ZNQ=a_{1}+b_{1}Z+c_{1}\sqrt{Z}N and R=a2+b2Z+c2ZNR=a_{2}+b_{2}Z+c_{2}\sqrt{Z}N, where a1,a2,b1,b2a_{1},a_{2},b_{1},b_{2} are any real numbers and c1>0,c2>0c_{1}>0,\;c_{2}>0. Then the following condition

a1+b1EZa2+b2EZandc1c2,a_{1}+b_{1}EZ\geq a_{2}+b_{2}EZ\;\;\mbox{and}\;\;c_{1}\leq c_{2}, (12)

is sufficient for Q(2)RQ\succeq_{(2)}R.

Proof.

We divide the proof into two cases. First assume a1+b1EZ>a2+b2EZa_{1}+b_{1}EZ>a_{2}+b_{2}EZ. Define Zm=Z1[1/m,m]Z^{m}=Z1_{[1/m,m]} each positive integer m2m\geq 2. Then by the dominated convergence theorem we have EZmEZEZ^{m}\rightarrow EZ. Therefore there exists a positive integer m0m_{0} such that we have

a1+b1EZma2+b2EZma_{1}+b_{1}EZ^{m}\geq a_{2}+b_{2}EZ^{m}

for all mm0m\geq m_{0}. Define Qm=:a1+b1Zm+c1ZmNQ^{m}=:a_{1}+b_{1}Z^{m}+c_{1}\sqrt{Z^{m}}N and Rm=:a2+b2Zm+c2ZmNR^{m}=:a_{2}+b_{2}Z^{m}+c_{2}\sqrt{Z^{m}}N. Observe that QmQQ^{m}\rightarrow Q and RmRR^{m}\rightarrow R in LkL^{k}. From Lemma 2.10 we have Qm(2)RmQ^{m}\succeq_{(2)}R^{m} for all mm0m\geq m_{0}. This implies

CVaRα(Qm)CVaRα(Rm),mm0,CVaR_{\alpha}(-Q^{m})\leq CVaR_{\alpha}(-R^{m}),\;\;\forall\;m\geq m_{0}, (13)

for each fixed α(0,1)\alpha\in(0,1). Since for each fixed α(0,1)\alpha\in(0,1) the risk measure CVaRα()CVaR_{\alpha}(\cdot) is continuous on LkL^{k}, by taking limits to both sides of (13) as m+m\rightarrow+\infty we obtain

CVaRα(Q)CVaRα(R),CVaR_{\alpha}(-Q)\leq CVaR_{\alpha}(-R),

for each α(0,1)\alpha\in(0,1). Then Q(2)RQ\succeq_{(2)}R follows from Lemma 2.13 above.

Now assume a1+b1EZ=a2+b2EZa_{1}+b_{1}EZ=a_{2}+b_{2}EZ. Define ZmZ^{m} as above and let δm=:(a1+b1EZm)(a2+b2EZm)\delta_{m}=:(a_{1}+b_{1}EZ^{m})-(a_{2}+b_{2}EZ^{m}). Since EZmEZEZ^{m}\rightarrow EZ we have δm0\delta_{m}\rightarrow 0 as m+m\rightarrow+\infty. Define the following two NMVM models Qm=|δm|+a1+b1Zm+c1ZmNQ^{m}=|\delta_{m}|+a_{1}+b_{1}Z^{m}+c_{1}\sqrt{Z^{m}}N and Rm=a2+b2Zm+c2ZmNR^{m}=a_{2}+b_{2}Z^{m}+c_{2}\sqrt{Z^{m}}N. We clearly have |δm|+a1+b1EZma2+b2EZm|\delta_{m}|+a_{1}+b_{1}EZ^{m}\geq a_{2}+b_{2}EZ^{m}. Therefore from Lemma 2.10 we have Qm(2)RmQ^{m}\succeq_{(2)}R^{m} for all m2m\geq 2. Also observe that QmQQ^{m}\rightarrow Q and RmRR^{m}\rightarrow R in LkL^{k}. Now by following the same arguments as in the case a1+b1EZ>a2+b2EZa_{1}+b_{1}EZ>a_{2}+b_{2}EZ above we obtain Q(2)RQ\succeq_{(2)}R. This completes the proof. ∎

Remark 2.15.

Our above Proposition 2.14 shows that the condition (12) is sufficient for Q(2)RQ\succeq_{(2)}R. It is not clear if Q(2)RQ\succeq_{(2)}R also implies (12) when the mixing distribution ZZ is in LkL^{k} for some finite positive integer kk. In comparison, if ZLZ\in L^{\infty} then by Lemma 2.10, the condition Q(2)RQ\succeq_{(2)}R implies (12). If ZLkZ\in L^{k} for all k1,k\geq 1, then from our Lemma 3.13 in Section 3 below the condition a1+b1m(Z)+c1ZN(2)a2+b2m(Z)+c2ZNa_{1}+b_{1}m(Z)+c_{1}\sqrt{Z}N\succeq_{(2)}a_{2}+b_{2}m(Z)+c_{2}\sqrt{Z}N implies a1+b1Em(Z)a2+b2Em(Z)a_{1}+b_{1}Em(Z)\geq a_{2}+b_{2}Em(Z) and c1c2c_{1}\leq c_{2} for any positively valued bounded Borel function m()m(\cdot).

Remark 2.16.

Fix any positive integer k1k\geq 1 and assume ZLkZ\in L^{k}. Consider the following two NMVM models Q=a1+b1Z+cZNQ=a_{1}+b_{1}Z+c\sqrt{Z}N and R=a2+b2Z+cZNR=a_{2}+b_{2}Z+c\sqrt{Z}N, where a1,a2,b1,b2a_{1},a_{2},b_{1},b_{2} are any real numbers and c>0c>0. Then Q(2)RQ\succeq_{(2)}R if and only if a1+b1EZa2+b2EZa_{1}+b_{1}EZ\geq a_{2}+b_{2}EZ. This easily follows from Proposition 2.14 and Theorem 1 of [26]

As an application of our results above we present the following example.

Example 2.17.

Let RR be the return vector of dd risky assets and assume that

RGHd(λ,α,β,δ,μ,Σ),R\sim GH_{d}(\lambda,\alpha,\beta,\delta,\mu,\Sigma),

a multi-dimensional hyperbolic distribution. For the definition and allowed parameter ranges of this distribution see Chapter 2 of [15]. As stated in the first paragraph at page 78 of [15], RR has the following NMVM representation

R=𝑑μ+ZΣβ+ZANd,R\overset{d}{=}\mu+Z\Sigma\beta+\sqrt{Z}AN_{d},

where A=Σ12A=\Sigma^{-\frac{1}{2}}. Consider the problem (6) for X=RX=R for any fixed real number rr and for any fixed α(0,1)\alpha\in(0,1). Our Theorem 2.4 shows that this problem has closed form solution and it is given by ωθ(μ+ΣβEZ,Σ)\omega_{\theta}(\mu+\Sigma\beta EZ,\;\Sigma) as in (5) . Note that the mixing distribution ZZ is ZGIG(λ,δ,α2βTΣβ)Z\sim GIG(\lambda,\delta,\sqrt{\alpha^{2}-\beta^{T}\Sigma\beta}) and we assume here that the parameters λ,α,β,δ,μ,Σ\lambda,\alpha,\beta,\delta,\mu,\Sigma are such that ZLkZ\in L^{k} for some positive integer k1k\geq 1, see page 11 of [15] for the moments of GIG distributions.

Remark 2.18.

We remark here that the conclusion of the above Example 2.17 is also true for other risk measures as long as it is law-invariant and SSD-consistent. For example, the following class of risk measures satisfy these conditions.

ρ^(X)=supμ(0,1]CVaRβ(X)μ(dβ),\hat{\rho}(X)=\sup_{\mu\in\mathcal{M}}\int_{(0,1]}CVaR_{\beta}(X)\mu(d\beta),

where \mathcal{M} is any subset of the set 1((0,1])\mathcal{M}_{1}((0,1]) of probability measures on (0,1](0,1], see Section 5.1 and the preceding sections of [12] for more details of these risk measures. We remark here that any finite valued law-invariant coherent risk measure has representation as above, see the paragraph before Corollary 5.1 of [12] for example.

3 High-order stochastic dominance

Stochastic dominance (SD) relation among random variables are defined by the point wise comparison of their distribution functions. It has rich applications in economics and finance topics such as option valuation, portfolio insurance, risk modeling etc. As such investigating necessary as well as sufficient conditions for SD within random variables is important. Our aim in this section is to investigate necessary as well as sufficient conditions for high-order SD relations within the class of one dimensional NMVM models.

Denoting by Uk={u::(1)j1u(j)(x)0,x,1jk}U_{k}=\{u:\mathbb{R}\rightarrow\mathbb{R}:(-1)^{j-1}u^{(j)}(x)\geq 0,\forall x\in\mathbb{R},1\leq j\leq k\}, the kk times differentiable utility functions with alternating signs of derivatives for each k1k\geq 1, we say that a scalar random variable η1\eta_{1} nthn^{\prime}th order stochastically dominates another scaler random variable η2\eta_{2} if Eu(η1)Eu(η2)Eu(\eta_{1})\geq Eu(\eta_{2}) for all uUku\in U_{k}. We use the notation η1(k)η2\eta_{1}\succeq_{(k)}\eta_{2} to denote this SD relation in our paper.

This SD relation among random variables can also be defined by using their high-order cumulative distribution functions. For any given real-valued random variable η\eta denote by Fη(1)(x)F^{(1)}_{\eta}(x) the right-continuous distribution function of it. For each integer k2k\geq 2 define the following functions

Fη(k)(x)=xFη(k1)(s)𝑑s,x,F^{(k)}_{\eta}(x)=\int_{-\infty}^{x}F^{(k-1)}_{\eta}(s)ds,\;\;\forall x\in\mathbb{R},

recursively. We call Fη(k)(x)F^{(k)}_{\eta}(x) the kth order cumulative distribution function of η\eta and we use the short-hand notation kCDFk-CDF to denote it from now on. For any two random variables η1\eta_{1} and η2\eta_{2}, the relation η1(k)η2\eta_{1}\succeq_{(k)}\eta_{2} can equivalently be defined by the following condition

Fη1(k)(x)Fη2(k)(x),x.F^{(k)}_{\eta_{1}}(x)\leq F^{(k)}_{\eta_{2}}(x),\;\;\forall x\in\mathbb{R}. (14)

Clearly the relation (14) is a high dimensional problem in the sense that the inequality there needs to be checked for all real number xx. Therefore investigating kSD relations among random variables is not a trivial issue. In this section, to obtain some results on kSD relations among NMVM models, we directly check the relation (14) after deriving and investigating the properties of their kCDFk-CDF for all positive integers kk. We refer to the following papers [3], [4], [25], [26] and the references there for further details on these kSD dominance relations.

3.1 High-order CDFs

The main purpose of this subsection is to compute the high-order CDFs of normal, elliptical, and NMVM models. Our calculations in this section mainly relies on some important characterizations of the kSD property developed in the paper [26]. Here we first review some of the important results in [26] that are useful for us. We use the same notations as in this paper. For each kk\in\mathbb{N} (here \mathbb{N} denotes the set of positive integers from now on) and for any random variable HLkH\in L^{k} define GH(k)(x)=(xH)+kG^{(k)}_{H}(x)=||(x-H)^{+}||_{k} for all xx\in\mathbb{R}, where x+x^{+} denotes the positive part of xx. For any positive integer k2k\geq 2 and any two H,QLkH,Q\in L^{k}, the paper [26] in their Proposition 1 shows that the kSD is equivalent to

GH(k1)(x)GQ(k1)(x),x.G_{H}^{(k-1)}(x)\leq G_{Q}^{(k-1)}(x),\;\forall x\in\mathbb{R}. (15)

According to the Proposition 6 of [26], the function GH(k)(x)G_{H}^{(k)}(x) is an increasing convex function with limxGH(k)(x)=0\lim_{x\rightarrow-\infty}G_{H}^{(k)}(x)=0 and GH(k)(x)xEHG_{H}^{(k)}(x)\geq x-EH for all xx\in\mathbb{R}. It was also shown in the same Proposition that xEHx-EH is a right asymptotic line of the function GH(k)(x)G_{H}^{(k)}(x), i.e.,

limx+(GH(k)(x)(xEH))=0.\lim_{x\rightarrow+\infty}\left(G_{H}^{(k)}(x)-(x-EH)\right)=0.

For the remainder of this section, we calculate kCDFk-CDF for some classes of random variables explicitly for any kk\in\mathbb{N}. Clearly these results are related to the kSD property through (14).

Normal random variables: When HN(0,1)H\sim N(0,1), we denote FH(k)(x)F_{H}^{(k)}(x) by ϕ(k)(x)\phi^{(k)}(x). We have

ϕ(k)(x)=1(k1)!E[(xN)+]k1,\phi^{(k)}(x)=\frac{1}{(k-1)!}E[(x-N)^{+}]^{k-1}, (16)

where NN is the standard normal random variable. We show the following simple Lemma.

Lemma 3.1.

For any k2k\geq 2, we have

ϕ(k)(x)=1k1[xϕ(k1)(x)+ϕ(k2)(x)],\phi^{(k)}(x)=\frac{1}{k-1}[x\phi^{(k-1)}(x)+\phi^{(k-2)}(x)], (17)

where ϕ(0)(x)=12πex22\phi^{(0)}(x)=\frac{1}{\sqrt{2\pi}}e^{-\frac{x^{2}}{2}} and ϕ(1)(x)=Φ(x)\phi^{(1)}(x)=\Phi(x).

Proof.

We use induction. When k=2k=2, we have

ϕ(2)(x)=xϕ(1)(s)ds=sϕ(1)(s)/xxsϕ(0)(s)ds=xϕ(1)(x)+ϕ(0)(x),\begin{split}\phi^{(2)}(x)=&\int_{-\infty}^{x}\phi^{(1)}(s)ds=s\phi^{(1)}(s)/_{-\infty}^{x}-\int_{-\infty}^{x}s\phi^{(0)}(s)ds\\ =&x\phi^{(1)}(x)+\phi^{(0)}(x),\\ \end{split}

where we have used lims[sϕ(1)(s)]=limsϕ(1)(s)/1s=0\lim_{s\rightarrow-\infty}[s\phi^{(1)}(s)]=\lim_{s\rightarrow-\infty}\phi^{(1)}(s)/\frac{1}{s}=0 which follows from L’Hopital’s rule. Assume (17) is true for kk and we show that it is also true for k+1k+1. To this end, we first integrate both sides of (17) and then apply (17). We obtain

ϕ(k+1)(x)=xϕ(k)(s)𝑑s=1k1xsϕ(k1)(s)𝑑s+1k1ϕ(k1)(x)=1k1[sϕ(k)](s)/x1k1ϕ(k+1)(x)+1k1ϕ(k1)(x)=1k1[xϕ(k)(x)]1k1ϕ(k+1)(x)+1k1ϕ(k1)(x),\begin{split}\phi^{(k+1)}(x)=&\int_{-\infty}^{x}\phi^{(k)}(s)ds=\frac{1}{k-1}\int_{-\infty}^{x}s\phi^{(k-1)}(s)ds+\frac{1}{k-1}\phi^{(k-1)}(x)\\ &=\frac{1}{k-1}[s\phi^{(k)}](s)/_{-\infty}^{x}-\frac{1}{k-1}\phi^{(k+1)}(x)+\frac{1}{k-1}\phi^{(k-1)}(x)\\ &=\frac{1}{k-1}[x\phi^{(k)}(x)]-\frac{1}{k-1}\phi^{(k+1)}(x)+\frac{1}{k-1}\phi^{(k-1)}(x),\end{split} (18)

where we have used lims[sϕ(1)(s)]=limsϕ(k)(s)/1s=0\lim_{s\rightarrow-\infty}[s\phi^{(1)}(s)]=\lim_{s\rightarrow-\infty}\phi^{(k)}(s)/\frac{1}{s}=0 which follows from multiple applications of L’Hopital’s rule. From equation (18) we obtain

ϕ(k+1)(x)=1k[xϕ(k)(x)+ϕ(k1)(x)],\phi^{(k+1)}(x)=\frac{1}{k}[x\phi^{(k)}(x)+\phi^{(k-1)}(x)],

and this completes the proof. ∎

Remark 3.2.

The relation (17) leads us to

ϕ(k)(0)={1357(2i3)(2i1)12πk=2i,14(2i2)(2i)12k=2i+1.\phi^{(k)}(0)=\left\{\begin{array}[]{ll}\frac{1}{3\cdot 5\cdot 7\cdots(2i-3)\cdot(2i-1)}\frac{1}{\sqrt{2\pi}}&k=2i,\\ \frac{1}{4\cdots(2i-2)\cdot(2i)}\frac{1}{2}&k=2i+1.\\ \end{array}\right.

We observe that ϕ(k)(0)\phi^{(k)}(0) is a decreasing sequence that goes to zero. From the relation (17) we have ϕ(k)(x)xk1ϕ(k1)(x)\phi^{(k)}(x)\geq\frac{x}{k-1}\phi^{(k-1)}(x) and therefore when xk1x\geq k-1, we have ϕ(k)(x)ϕ(k1)(x)\phi^{(k)}(x)\geq\phi^{(k-1)}(x). This shows that ϕ(k)(x)\phi^{(k)}(x) and ϕ(k1)(x)\phi^{(k-1)}(x) intersects at some point x>0x>0.

Next, by using (17), we can obtain the following expression for ϕ(k)(x)\phi^{(k)}(x)

Lemma 3.3.

For any k2k\geq 2 we have

ϕ(k)(x)=pk1(x)Φ(x)+qk2(x)ϕ(0)(x),\phi^{(k)}(x)=p_{k-1}(x)\Phi(x)+q_{k-2}(x)\phi^{(0)}(x), (19)

where pk1(x)p_{k-1}(x) is a k1k-1’th order polynomial that satisfies

pj(x)=xjpj1(x)+1jpj2(x),j2,p1(x)=x,p0(x)=1,p_{j}(x)=\frac{x}{j}p_{j-1}(x)+\frac{1}{j}p_{j-2}(x),\;\;j\geq 2,\;\;\;p_{1}(x)=x,\;\;p_{0}(x)=1,

and qk2(x)q_{k-2}(x) is a k2k-2’th order polynomial that satisfies

qi(x)=xi+1qj1(x)+1i+1qi2(x),i2,q1(x)=x2,q0(x)=1.q_{i}(x)=\frac{x}{i+1}q_{j-1}(x)+\frac{1}{i+1}q_{i-2}(x),\;\;i\geq 2,\;\;\;q_{1}(x)=\frac{x}{2},\;\;q_{0}(x)=1.
Proof.

When k=2k=2, from (17) it is easy to see ϕ(2)(x)=xΦ(x)+ϕ(0)(x)=p1(x)Φ(x)+q0(x)ϕ(0)(x)\phi^{(2)}(x)=x\Phi(x)+\phi^{(0)}(x)=p_{1}(x)\Phi(x)+q_{0}(x)\phi^{(0)}(x). When k=3k=3, we can easily calculate ϕ(3)(x)=x2+12ϕ(2)(x)+x2ϕ(0)(x)=p2(x)Φ(x)+q1(x)ϕ(0)(x)\phi^{(3)}(x)=\frac{x^{2}+1}{2}\phi^{(2)}(x)+\frac{x}{2}\phi^{(0)}(x)=p_{2}(x)\Phi(x)+q_{1}(x)\phi^{(0)}(x) again by using (17). Now assume (19) is true for all 2,3,,k2,3,\cdots,k and we would like to prove it for k+1k+1. By (17) we have

ϕ(k+1)(x)=xkϕ(k)(x)+1kϕ(k1)(x)=xk[pk1(x)Φ(x)+qk2(x)ϕ(0)(x)]+1k[pk2(x)Φ(x)+qk3(x)ϕ(0)(x)]=[xkpk1(x)+1kpk2(x)]Φ(x)+[xkqk2(x)+1kqk3(x)]ϕ(0)(x).=pk(x)Φ(x)+qk1(x)ϕ(0)(x),\begin{split}\phi^{(k+1)}(x)&=\frac{x}{k}\phi^{(k)}(x)+\frac{1}{k}\phi^{(k-1)}(x)\\ &=\frac{x}{k}[p_{k-1}(x)\Phi(x)+q_{k-2}(x)\phi^{(0)}(x)]+\frac{1}{k}[p_{k-2}(x)\Phi(x)+q_{k-3}(x)\phi^{(0)}(x)]\\ &=[\frac{x}{k}p_{k-1}(x)+\frac{1}{k}p_{k-2}(x)]\Phi(x)+[\frac{x}{k}q_{k-2}(x)+\frac{1}{k}q_{k-3}(x)]\phi^{(0)}(x).\\ &=p_{k}(x)\Phi(x)+q_{k-1}(x)\phi^{(0)}(x),\end{split}

where pk(x)=:xkpk1(x)+1kpk2(x)p_{k}(x)=:\frac{x}{k}p_{k-1}(x)+\frac{1}{k}p_{k-2}(x) and qk1(x)=:xkqk2(x)+1kqk3(x)q_{k-1}(x)=:\frac{x}{k}q_{k-2}(x)+\frac{1}{k}q_{k-3}(x). Clearly pk(x)p_{k}(x) is a kk^{\prime}th order polynomial and qk1(s)q_{k-1}(s) is a (k1)(k-1)^{\prime}th order polynomial. ∎

Lemma 3.4.

The function y(x)=ϕ(k)(x)y(x)=\phi^{(k)}(x) satisfies

y′′+xy(k1)y=0,y(0)=ϕ(k)(0),y(0)=ϕ(k1)(0),\begin{split}&y^{{}^{\prime\prime}}+xy^{\prime}-(k-1)y=0,\\ &y(0)=\phi^{(k)}(0),\;y^{\prime}(0)=\phi^{(k-1)}(0),\\ \end{split} (20)

and the polynomial solution y(x)=j=0+ajxjy(x)=\sum_{j=0}^{+\infty}a_{j}x^{j} of (20) is given by

aj+3=(k1)(j+1)(j+2)(j+3)aj+1,j=0,1,,a_{j+3}=\frac{(k-1)-(j+1)}{(j+2)(j+3)}a_{j+1},\;\;j=0,1,\cdots,
a2=k12a0,a0=ϕk(0),a1=ϕk1(0).a_{2}=\frac{k-1}{2}a_{0},\;\;a_{0}=\phi^{k}(0),\;\;a_{1}=\phi^{k-1}(0).
Proof.

The equation (17) can be written as (k1)ϕ(k)(x)=xϕ(k1)(x)+ϕ(k2)(x)(k-1)\phi^{(k)}(x)=x\phi^{(k-1)}(x)+\phi^{(k-2)}(x). With y=:ϕ(k)(x)y=:\phi^{(k)}(x) observe that y=ϕ(k1)(x)y^{\prime}=\phi^{(k-1)}(x) and y′′=ϕ(k2)(x)y^{{}^{\prime\prime}}=\phi^{(k-2)}(x). Therefore the equation in (20) holds. To find its polynomial solution we plug y(x)=j=0+ajxjy(x)=\sum_{j=0}^{+\infty}a_{j}x^{j} into (20) and obtain a new polynomial that equals to zero. Then all the co-efficients of this new polynomial are zero. This gives us the expressions for aja_{j}. ∎

Remark 3.5.

Observe that ak+1=0a_{k+1}=0 and hence ak+2j+1=0a_{k+2j+1}=0 for all j0j\geq 0. We have

ak+2j=(1)j35(2j3)(2j1)(k+2j)!12π,a_{k+2j}=(-1)^{j}\frac{3\cdot 5\cdots(2j-3)\cdot(2j-1)}{(k+2j)!}\frac{1}{\sqrt{2\pi}},

and

a0=ϕ(k)(0),a1=ϕ(k1)(0),,aj=1j!ϕ(kj)(0),,ak=1k!ϕ(0)(0)=1k!12π.a_{0}=\phi^{(k)}(0),a_{1}=\phi^{(k-1)}(0),\cdots,a_{j}=\frac{1}{j!}\phi^{(k-j)}(0),\cdots,a_{k}=\frac{1}{k!}\phi^{(0)}(0)=\frac{1}{k!}\frac{1}{\sqrt{2\pi}}.

In the case that HN(μ,σ2)H\sim N(\mu,\sigma^{2}), we denote FH(k)(x)F_{H}^{(k)}(x) by φ(k)(x;μ,σ2)\varphi^{(k)}(x;\mu,\sigma^{2}). By using (16) , we can easily obtain

φ(k)(x;μ,σ2)=σk1ϕ(k)(xμσ).\varphi^{(k)}(x;\mu,\sigma^{2})=\sigma^{k-1}\phi^{(k)}(\frac{x-\mu}{\sigma}).

By letting y(x)=φ(k)(x;μ,σ2)y(x)=\varphi^{(k)}(x;\mu,\sigma^{2}), one can easily show that it satisfies the following equation

σ2y′′+(xμ)y(k1)y=0,\sigma^{2}y^{{}^{\prime\prime}}+(x-\mu)y^{\prime}-(k-1)y=0,

with the initial conditions

y(0)=σk1ϕ(k)(μσ),y(0)=σk2ϕ(k1)(μσ).y(0)=\sigma^{k-1}\phi^{(k)}(-\frac{\mu}{\sigma}),\;y^{\prime}(0)=\sigma^{k-2}\phi^{(k-1)}(-\frac{\mu}{\sigma}).

Elliptical random variables: When Hμ+σZN(0,1)H\sim\mu+\sigma\sqrt{Z}N(0,1), where ZLk1Z\in L^{k-1} is a positive random variable independent from N(0,1)N(0,1), we denote F(k)(x)F^{(k)}(x) by φe(x;μ,σ)\varphi_{e}(x;\mu,\sigma). From (16), we have

φe(k)(x;μ,σ)=σk10+zk12ϕ(k)(xμσz)fZ(z)𝑑z.\varphi_{e}^{(k)}(x;\mu,\sigma)=\sigma^{k-1}\int_{0}^{+\infty}z^{\frac{k-1}{2}}\phi^{(k)}(\frac{x-\mu}{\sigma\sqrt{z}})f_{Z}(z)dz.
Lemma 3.6.

For each k2k\geq 2, when ZLk1Z\in L^{k-1} we have

dφe(k)(x;μ,σ)dμ=σk20+zk22ϕ(k1)(xμσz)fZ(z)𝑑z<0,dφe(k)(x;μ,σ)dσ=σk20+zk12ϕ(k2)(xμσz)fZ(z)𝑑z>0.\begin{split}\frac{d\varphi_{e}^{(k)}(x;\mu,\sigma)}{d\mu}=&-\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k-2}{2}}\phi^{(k-1)}(\frac{x-\mu}{\sigma\sqrt{z}})f_{Z}(z)dz<0,\\ \frac{d\varphi_{e}^{(k)}(x;\mu,\sigma)}{d\sigma}=&\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k-1}{2}}\phi^{(k-2)}(\frac{x-\mu}{\sigma\sqrt{z}})f_{Z}(z)dz>0.\end{split} (21)
Proof.

The first part of (21) is direct. To see the second part note that

dφe(k)(x;μ,σ)dσ=(k1)σk20+zk12ϕ(k)(xμσz)fZ(z)𝑑z+σk10+zk12(xμσ2z)ϕ(k1)(xμσz)fZ(z)𝑑z=σk20+zk12[(k1)ϕ(k)(xμσz)(xμσz)ϕ(k1)(xμσz)]fZ(z)𝑑z.\begin{split}\frac{d\varphi_{e}^{(k)}(x;\mu,\sigma)}{d\sigma}=&(k-1)\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k-1}{2}}\phi^{(k)}(\frac{x-\mu}{\sigma\sqrt{z}})f_{Z}(z)dz\\ +&\sigma^{k-1}\int_{0}^{+\infty}z^{\frac{k-1}{2}}(-\frac{x-\mu}{\sigma^{2}\sqrt{z}})\phi^{(k-1)}(\frac{x-\mu}{\sigma\sqrt{z}})f_{Z}(z)dz\\ =&\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k-1}{2}}[(k-1)\phi^{(k)}(\frac{x-\mu}{\sigma\sqrt{z}})-(\frac{x-\mu}{\sigma\sqrt{z}})\phi^{(k-1)}(\frac{x-\mu}{\sigma\sqrt{z}})]f_{Z}(z)dz.\end{split} (22)

From (17) we have

(k1)ϕ(k)(xμσz)xμσzϕ(k1)(xμσz)=ϕ(k2)(xμσz).(k-1)\phi^{(k)}(\frac{x-\mu}{\sigma\sqrt{z}})-\frac{x-\mu}{\sigma\sqrt{z}}\phi^{(k-1)}(\frac{x-\mu}{\sigma\sqrt{z}})=\phi^{(k-2)}(\frac{x-\mu}{\sigma\sqrt{z}}). (23)

Then we plug (23) into (22). This gives us the second equation in the Proposition. ∎

Remark 3.7.

Lemma 3.6 shows that for any two random variables H=μ1+σ1ZNH=\mu_{1}+\sigma_{1}ZN and Q=μ2+σ2ZNQ=\mu_{2}+\sigma_{2}ZN with ZLk1Z\in L^{k-1}, the relation μ1μ2\mu_{1}\geq\mu_{2} and σ2σ1>0\sigma_{2}\geq\sigma_{1}>0 implies HkQH\succeq_{k}Q. This result clear follows from Proposition 2.14 in Section 2 above. Here our Lemma 3.6 compares the high-order CDFs of HH and QQ directly which can be seen as an alternative approach.

NMVM random variables: When Hμ+γZ+σZN(0,1)H\sim\mu+\gamma Z+\sigma\sqrt{Z}N(0,1), we denote F(k)(x)F^{(k)}(x) by φh(k)(x;μ,γ,σ)\varphi_{h}^{(k)}(x;\mu,\gamma,\sigma) (taking the first letter in hyperbolic distributions). From (16), we have

φh(k)(x;μ,γ,σ)=σk10+zk12ϕ(k)(xμγzσz)fZ(z)𝑑z\varphi_{h}^{(k)}(x;\mu,\gamma,\sigma)=\sigma^{k-1}\int^{+\infty}_{0}z^{\frac{k-1}{2}}\phi^{(k)}(\frac{x-\mu-\gamma z}{\sigma\sqrt{z}})f_{Z}(z)dz (24)

By direct calculation and by using (17), we obtain the following result

Lemma 3.8.

For each positive integer k2k\geq 2, when ZLk1Z\in L^{k-1} we have

dφh(k)(x;μ,γ,σ)dμ=σk20+zk22ϕ(k1)(xμγzσz)fZ(z)𝑑z<0,dφh(k)(x;μ,γ,σ)dγ=σk20+zk2ϕ(k1)(ημγzσz)fZ(z)𝑑z<0,dφh(k)(x;μ,γ,σ)dσ=σk20+zk12ϕ(k2)(ημγzσz)fZ(z)𝑑z>0.\begin{split}\frac{d\varphi_{h}^{(k)}(x;\mu,\gamma,\sigma)}{d\mu}=&-\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k-2}{2}}\phi^{(k-1)}(\frac{x-\mu-\gamma z}{\sigma\sqrt{z}})f_{Z}(z)dz<0,\\ \frac{d\varphi_{h}^{(k)}(x;\mu,\gamma,\sigma)}{d\gamma}=&-\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k}{2}}\phi^{(k-1)}(\frac{\eta-\mu-\gamma z}{\sigma\sqrt{z}})f_{Z}(z)dz<0,\\ \frac{d\varphi_{h}^{(k)}(x;\mu,\gamma,\sigma)}{d\sigma}=&\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k-1}{2}}\phi^{(k-2)}(\frac{\eta-\mu-\gamma z}{\sigma\sqrt{z}})f_{Z}(z)dz>0.\\ \end{split} (25)
Proof.

The first two equations of (25) follow from taking the corresponding derivatives of (24). The third equation is obtained by taking the derivative of (24) with respect to σ\sigma and by using the relation (17) similar to the proof of Proposition 3.6. ∎

Remark 3.9.

Lemma 3.8 implies that for any two random variables Hμ1+γ1Z+σ1ZN(0,1)H\sim\mu_{1}+\gamma_{1}Z+\sigma_{1}\sqrt{Z}N(0,1) and Qμ2+γ2Z+σ2ZN(0,1)Q\sim\mu_{2}+\gamma_{2}Z+\sigma_{2}\sqrt{Z}N(0,1) the condition μ1μ2,γ1γ2,σ2σ1>0\mu_{1}\geq\mu_{2},\gamma_{1}\geq\gamma_{2},\sigma_{2}\geq\sigma_{1}>0 implies H(k)QH\succeq_{(k)}Q, a result that also follows from Proposition 2.14 above.

3.2 High-order SD

As stated earlier the condition (14), which needs to be checked for all the real numbers xx, shows that SD is an infinite dimensional problem and hence necessary as well as sufficient conditions for SD are difficult to obtain. For some special cases of random variables some characterizations of the second order stochastic dominance relation is well known in the literature. For example, if N1N(μ1,σ1)N_{1}\sim N(\mu_{1},\sigma_{1}) and N2N(μ2,σ2)N_{2}\sim N(\mu_{2},\sigma_{2}), then N1(2)N2N_{1}\succeq_{(2)}N_{2} if an only if μ1μ2\mu_{1}\geq\mu_{2} and σ1σ2\sigma_{1}\leq\sigma_{2}, see Proposition 2.59 of [13], Theorem 6.2 of [20] for instance. For the general case of random variables such convenient necessary and sufficient conditions for SD are difficult to construct. In this section we introduce some necessary as well as sufficient conditions for kSD for the case of NMVM models.

From its definition it is easy to see that the n1thn_{1}^{\prime}th order SD relation is a weaker condition than the n2thn_{2}^{\prime}th order SD relation for any n1n2n_{1}\geq n_{2} as Un1Un2U_{n_{1}}\subset U_{n_{2}}. Therefore it is not immediately clear that if N1(k)N2N_{1}\succeq_{(k)}N_{2} also implies the relation μ1μ2,σ1σ2,\mu_{1}\geq\mu_{2},\sigma_{1}\leq\sigma_{2}, for k3k\geq 3 in the case of normal random variables. In this section, as an offshoot of a general result on NVMM, we will show that in fact N1(k+1)N2N_{1}\succeq_{(k+1)}N_{2} is equivalent to μ1μ2\mu_{1}\geq\mu_{2} and σ1σ2\sigma_{1}\leq\sigma_{2} for each kk\in\mathbb{N}, see Proposition 3.10 below for this. Our proof here gives alternative approach for the proof of the second order stochastic dominance characterization on normal random variables ( see Theorem 6.2 of [20] and Theorem 3.1 of [5]), and in the meantime it also extends similar type of characterizations to kSD (especially for k3k\geq 3 ) relations on elliptical random variables.

Before we proceed with our calculations, we first recall a relevant result from the paper [26]. With GH(k)(x)G_{H}^{(k)}(x) defined as in sub-section 3.1 above, the value of GH(k)(x)G_{H}^{(k)}(x) at x=EHx=EH is called central semideviation of HH and it is denoted by δ¯H(k)=GH(k)(EH)\bar{\delta}_{H}^{(k)}=G_{H}^{(k)}(EH) (again using the same notation in the paper [26]). That is we have

δ¯H(k)=(EHH)+k=[E((EHH)+)k]1k.\bar{\delta}_{H}^{(k)}=||(EH-H)^{+}||_{k}=[E\left((EH-H)^{+}\right)^{k}]^{\frac{1}{k}}.

Proposition 4 of [26] shows that δ¯H(k)\bar{\delta}_{H}^{(k)} is a convex function, i.e., δ¯tH+(1t)Q(k)tδ¯H(k)+(1t)δ¯Q(k)\bar{\delta}_{tH+(1-t)Q}^{(k)}\leq t\bar{\delta}_{H}^{(k)}+(1-t)\bar{\delta}_{Q}^{(k)} for any t[0,1]t\in[0,1] and any H,QLkH,Q\in L^{k}. Corollary 2 of the same paper shows that if H(k+1)QH\succeq_{(k+1)}Q then

EHδ¯H(n)EQδ¯Q(n),EHEQ,EH-\bar{\delta}_{H}^{(n)}\geq EQ-\bar{\delta}_{Q}^{(n)},\;\;\;EH\geq EQ, (26)

for all nkn\geq k as long as HLnH\in L^{n}. This relation (26) plays an important role in our discussions in this sub-section.

As an immediate application of this result, we show the following Proposition first.

Proposition 3.10.

Let Hμ1+σ1ZN(0,1)H\sim\mu_{1}+\sigma_{1}ZN(0,1) and Qμ2+σ2ZN(0,1)Q\sim\mu_{2}+\sigma_{2}ZN(0,1) be two elliptical random variables with μ1,μ2,σ1>0,σ2>0,\mu_{1},\mu_{2}\in\mathbb{R},\sigma_{1}>0,\sigma_{2}>0, and ZZ is any positive random variable with EZj<EZ^{j}<\infty for all positive integers jj. Then for each k2k\geq 2, X(k)YX\succeq_{(k)}Y if and only if μ1μ2\mu_{1}\geq\mu_{2} and σ1σ2\sigma_{1}\leq\sigma_{2}.

Proof.

The relation μ1μ2\mu_{1}\geq\mu_{2} and σ1σ2\sigma_{1}\leq\sigma_{2} implies X(k)YX\succeq_{(k)}Y for each k2k\geq 2 follows from Proposition 3.6 above. To see the other direction, note that EH=μ1EH=\mu_{1} and EQ=μ2EQ=\mu_{2} and so the relation μ1μ2\mu_{1}\geq\mu_{2} follows from Theorem 1 of [26]. To see the other relation σ1σ2\sigma_{1}\leq\sigma_{2}, note that the central semi-deviations of HH and QQ are δ¯H(j)=σ1(ZN)+j\bar{\delta}^{(j)}_{H}=\sigma_{1}||(ZN)^{+}||_{j} and δ¯Q(j)=σ2(ZN)+j\bar{\delta}^{(j)}_{Q}=\sigma_{2}||(ZN)^{+}||_{j}. Corollary 2 of [26] implies

μ1σ1(ZN)+jμ2σ2(ZN)+j,jk.\mu_{1}-\sigma_{1}||(ZN)^{+}||_{j}\geq\mu_{2}-\sigma_{2}||(ZN)^{+}||_{j},\forall j\geq k. (27)

Now, since (ZN)+(ZN)^{+} is an unbounded random variable we have limj(ZN)+j=\lim_{j\rightarrow\infty}||(ZN)^{+}||_{j}=\infty. Dividing both sides of (27) by (ZN)+j||(ZN)^{+}||_{j} and letting jj\rightarrow\infty we obtain σ1σ2-\sigma_{1}\geq-\sigma_{2}. This completes the proof. ∎

Remark 3.11.

The significant part of this Proposition 3.10 is that the relation X(k)YX\succeq_{(k)}Y implies μ1μ2\mu_{1}\geq\mu_{2} and σ1σ2\sigma_{1}\leq\sigma_{2} for any fixed k2k\geq 2 as long as ZLjZ\in L^{j} for all j1j\geq 1, a result which seems to be new in the literature.

Next we discuss kSD relations within the class of one dimensional NMVM models. First we introduce some notations. Let m()m(\cdot) denote any Borel function from (0,+)(0,+\infty) to (0,+)(0,+\infty). Denote

Zm=:m(Z).Z_{m}=:m(Z).

For any real numbers a1,b1,a2,b2,a_{1},b_{1},a_{2},b_{2}, define X¯m=a1+b1Zm+ZN\bar{X}_{m}=a_{1}+b_{1}Z_{m}+\sqrt{Z}N and Y¯m=a2+b2Zm+ZN\bar{Y}_{m}=a_{2}+b_{2}Z_{m}+\sqrt{Z}N. We first prove the following Lemma.

Lemma 3.12.

Consider the model (1) and assume that ZLjZ\in L^{j} for all positive integers jj. Then for any bounded Borel function mm, we have

limk+(X¯m)+k(Y¯m)+k=1.\lim_{k\rightarrow+\infty}\frac{||(\bar{X}_{m})^{+}||_{k}}{||(\bar{Y}_{m})^{+}||_{k}}=1. (28)
Proof.

Observe that X¯mY¯m=a1a2+(b1b2)Zm\bar{X}_{m}-\bar{Y}_{m}=a_{1}-a_{2}+(b_{1}-b_{2})Z_{m} are bounded random variables. By using (a+b)+a++b+(a+b)^{+}\leq a^{+}+b^{+} for any real numbers and the triangle inequality for norms, we have

(X¯m)+k=(X¯mY¯m+Y¯m)+k(X¯mY¯m)+k+(Y¯m)+k.||(\bar{X}_{m})^{+}||_{k}=||(\bar{X}_{m}-\bar{Y}_{m}+\bar{Y}_{m})^{+}||_{k}\leq||(\bar{X}_{m}-\bar{Y}_{m})^{+}||_{k}+||(\bar{Y}_{m})^{+}||_{k}.

From this it follows that

(X¯m)+k(Y¯m)+k1+(X¯mY¯m)+k(Y¯m)+k.\frac{||(\bar{X}_{m})^{+}||_{k}}{||(\bar{Y}_{m})^{+}||_{k}}\leq 1+\frac{||(\bar{X}_{m}-\bar{Y}_{m})^{+}||_{k}}{||(\bar{Y}_{m})^{+}||_{k}}. (29)

Since (X¯mY¯m)+(\bar{X}_{m}-\bar{Y}_{m})^{+} are bounded random variables, we have supk1(X¯mY¯m)+k<\sup_{k\geq 1}||(\bar{X}_{m}-\bar{Y}_{m})^{+}||_{k}<\infty and since (Y¯m)+(\bar{Y}_{m})^{+} are unbounded random variables we have limkY¯mk\lim_{k\rightarrow\infty}||\bar{Y}_{m}||_{k}\rightarrow\infty. Therefore from (29) we conclude that

lim¯k(X¯m)+k(Y¯m)+k1.\overline{\lim}_{k\rightarrow\infty}\frac{||(\bar{X}_{m})^{+}||_{k}}{||(\bar{Y}_{m})^{+}||_{k}}\leq 1. (30)

Following the same idea, we have

(Y¯m)+k=(Y¯mX¯m+X¯m)+k(X¯mY¯m)+k+(X¯m)+k.||(\bar{Y}_{m})^{+}||_{k}=||(\bar{Y}_{m}-\bar{X}_{m}+\bar{X}_{m})^{+}||_{k}\leq||(\bar{X}_{m}-\bar{Y}_{m})^{+}||_{k}+||(\bar{X}_{m})^{+}||_{k}.

From this we obtain

(X¯m)+k(Y¯m)+k(X¯m)+k(X¯mY¯m)+k+(X¯m)+k=1(X¯mY¯m)+k/(X¯m)+k+1.\frac{||(\bar{X}_{m})^{+}||_{k}}{||(\bar{Y}_{m})^{+}||_{k}}\geq\frac{||(\bar{X}_{m})^{+}||_{k}}{||(\bar{X}_{m}-\bar{Y}_{m})^{+}||_{k}+||(\bar{X}_{m})^{+}||_{k}}=\frac{1}{||(\bar{X}_{m}-\bar{Y}_{m})^{+}||_{k}/||(\bar{X}_{m})^{+}||_{k}+1}. (31)

Since (X¯m)+(\bar{X}_{m})^{+} are unbounded random variables we have limk(X¯m)+k=\lim_{k\rightarrow\infty}||(\bar{X}_{m})^{+}||_{k}=\infty and therefore (X¯mY¯m)+k/(X¯m)+k0||(\bar{X}_{m}-\bar{Y}_{m})^{+}||_{k}/||(\bar{X}_{m})^{+}||_{k}\rightarrow 0 as kk\rightarrow\infty. Therefore from (31) we conclude that

lim¯k(X¯m)+k(Y¯m)+k1.\underline{\lim}_{k\rightarrow\infty}\frac{||(\bar{X}_{m})^{+}||_{k}}{||(\bar{Y}_{m})^{+}||_{k}}\geq 1. (32)

Now, from (30) and (32) we obtain (28). ∎

Next, for any bounded Borel function mm and any real numbers a1,a2,b1,b2,a_{1},a_{2},b_{1},b_{2}, and c1>0,c2>0c_{1}>0,c_{2}>0, let X~m=a1+b1Zm+c1ZN\tilde{X}_{m}=a_{1}+b_{1}Z_{m}+c_{1}\sqrt{Z}N and Y~m=a2+b2Zm+c2ZN\tilde{Y}_{m}=a_{2}+b_{2}Z_{m}+c_{2}\sqrt{Z}N. The following Proposition is our main result in this Section.

Proposition 3.13.

Consider the model (1) and assume that ZLjZ\in L^{j} for all positive integers jj. Then for any bounded Borel function mm and for each kk\in\mathbb{N}, the relation X~m(k+1)Y~m\tilde{X}_{m}\succeq_{(k+1)}\tilde{Y}_{m} implies a1+b1EZma2+b2EZma_{1}+b_{1}EZ_{m}\geq a_{2}+b_{2}EZ_{m} and c1c2c_{1}\leq c_{2}.

Proof.

Since EX~m=a1+b1EZmE\tilde{X}_{m}=a_{1}+b_{1}EZ_{m} and EY~m=a2+b2EZmE\tilde{Y}_{m}=a_{2}+b_{2}EZ_{m}, the relation a1+b1EZma2+b2EZma_{1}+b_{1}EZ_{m}\geq a_{2}+b_{2}EZ_{m} follows from Theorem 1 of [26]. To show c1c2c_{1}\leq c_{2}, we use Corollary 2 of the same paper [26]. First observe that for any integer j>kj>k the central semi-deviations of order jj are equal to

δ¯X~m(j)=c1(b1c1EZmb1c1Zm+ZN)+j,δ¯Y~m(j)=c2(b2c2EZmb2c2Zm+ZN)+j.\bar{\delta}_{\tilde{X}_{m}}^{(j)}=c_{1}||(\frac{b_{1}}{c_{1}}EZ_{m}-\frac{b_{1}}{c_{1}}Z_{m}+ZN)^{+}||_{j},\;\;\bar{\delta}_{\tilde{Y}_{m}}^{(j)}=c_{2}||(\frac{b_{2}}{c_{2}}EZ_{m}-\frac{b_{2}}{c_{2}}Z_{m}+ZN)^{+}||_{j}.

Denote Dj=:||(b1c1EZmb1c1Zm+ZN)+||jD_{j}=:||(\frac{b_{1}}{c_{1}}EZ_{m}-\frac{b_{1}}{c_{1}}Z_{m}+ZN)^{+}||_{j} and Ej=:||(b2c2EZmb2c2Zm+ZN)+||jE_{j}=:||(\frac{b_{2}}{c_{2}}EZ_{m}-\frac{b_{2}}{c_{2}}Z_{m}+ZN)^{+}||_{j} and observe that Lemma 3.12 implies limjDjEj=1\lim_{j\rightarrow\infty}\frac{D_{j}}{E_{j}}=1. Also since (b1c1EZmb1c1Zm+ZN)+(\frac{b_{1}}{c_{1}}EZ_{m}-\frac{b_{1}}{c_{1}}Z_{m}+ZN)^{+} and (b2c2EZmb2c2Zm+ZN)+(\frac{b_{2}}{c_{2}}EZ_{m}-\frac{b_{2}}{c_{2}}Z_{m}+ZN)^{+} are unbounded random variables we have limj+Dj=+\lim_{j\rightarrow+\infty}D_{j}=+\infty and limj+Ej=+\lim_{j\rightarrow+\infty}E_{j}=+\infty. Corollary 2 of [26] implies

a1+b1EZmc1Dja2+b2EZmc2Ej,a_{1}+b_{1}EZ_{m}-c_{1}D_{j}\geq a_{2}+b_{2}EZ_{m}-c_{2}E_{j}, (33)

for any jkj\geq k. Now dividing both sides of (33) by EjE_{j} and letting jj\rightarrow\infty we obtain c1c2-c_{1}\geq-c_{2}. This shows that c1c2c_{1}\leq c_{2}. ∎

Now for any real numbers b1,b2,b_{1},b_{2}, and any c1>0,c2>0c_{1}>0,c_{2}>0, let X^m=b1Zm+c1ZN\hat{X}_{m}=b_{1}Z_{m}+c_{1}\sqrt{Z}N and Y^m=b2Zm+c2ZN\hat{Y}_{m}=b_{2}Z_{m}+c_{2}\sqrt{Z}N. We have the following Corollary.

Corollary 3.14.

For any positive valued and bounded Borel function mm and for each kk\in\mathbb{N}, we have

X^m(k+1)Y^mb1b2,c1c2.\hat{X}_{m}\succeq_{(k+1)}\hat{Y}_{m}\Leftrightarrow b_{1}\geq b_{2},\;\;c_{1}\leq c_{2}.
Proof.

If X^m(k+1)Y^m\hat{X}_{m}\succeq_{(k+1)}\hat{Y}_{m}, then from Lemma 3.13 we have b1EZmb2EZmb_{1}EZ_{m}\geq b_{2}EZ_{m} and c1c2c_{1}\leq c_{2}. This is equivalent to b1b2b_{1}\geq b_{2} and c1c2c_{1}\leq c_{2} as EZm>0EZ_{m}>0. To show the other direction, observe that for any Ym=:γZm+σZNY_{m}=:\gamma Z_{m}+\sigma\sqrt{Z}N with γ,σ>0\gamma\in\mathbb{R},\;\sigma>0, we have

φYm(k)(x;γ,σ)=σk10+zk12ϕ(k)(xγzmσz)fZ(z)𝑑z.\varphi_{Y_{m}}^{(k)}(x;\gamma,\sigma)=\sigma^{k-1}\int^{+\infty}_{0}z^{\frac{k-1}{2}}\phi^{(k)}(\frac{x-\gamma z_{m}}{\sigma\sqrt{z}})f_{Z}(z)dz.

Now following the same idea as in the proof of Proposition 3.8, we can show that

dφYm(k)(x;γ,σ)dγ=σk20+zk22zmϕ(k1)(xγzmσz)fZ(z)𝑑z<0,dφYm(k)(x;γ,σ)dσ=σk20+zk12ϕ(k2)(xγzmσz)fZ(z)𝑑z>0.\begin{split}\frac{d\varphi_{Y_{m}}^{(k)}(x;\gamma,\sigma)}{d\gamma}=&-\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k-2}{2}}z_{m}\phi^{(k-1)}(\frac{x-\gamma z_{m}}{\sigma\sqrt{z}})f_{Z}(z)dz<0,\\ \frac{d\varphi_{Y_{m}}^{(k)}(x;\gamma,\sigma)}{d\sigma}=&\sigma^{k-2}\int_{0}^{+\infty}z^{\frac{k-1}{2}}\phi^{(k-2)}(\frac{x-\gamma z_{m}}{\sigma\sqrt{z}})f_{Z}(z)dz>0.\\ \end{split}

This shows that b1b2b_{1}\geq b_{2} and c1c2c_{1}\leq c_{2} implies X^m(k+1)Y^m\hat{X}_{m}\succeq_{(k+1)}\hat{Y}_{m}. ∎

We remark here that in our calculations in this paper we have used the probability density functions fZ(z)f_{Z}(z) of the mixing distributions ZZ. Recall that only random variables with absolutely continuous cumulative distribution functions possess probability density functions. Our results in this paper however holds in the case of mixing random variables ZZ that satisfy the integrability conditions stated in each results without requiring that they have density functions. All the calculations can be carried out by conditioning argument as in

GX(k)=EZxXzk,G^{(k)}_{X}=E^{Z}||x-X_{z}||_{k},

where Xz=a+bz+czN(0,1)X_{z}=a+bz+c\sqrt{z}N(0,1) for all zz in the support of ZZ for example.

Acknowledgements

The author would like to thank for the comments and directions provided by Alexander Schied in the early stages of this project.

References

  • Aas and Haff [2006] K. Aas and I. H. Haff. The generalized hyperbolic skew student’st-distribution. Journal of financial econometrics, 4(2):275–309, 2006.
  • Barndorff-Nilesen [1977] O. E. Barndorff-Nilesen. Exponentail decreasing distributions for the logarithm of particle size. Proceedings of the Royal Society London A, 353:401–409, 1977.
  • Dentcheva and Ruszczynski [2003] D. Dentcheva and A. Ruszczynski. Optimization with stochastic dominance constraints. SIAM Journal on Optimization, 14(2):548–566, 2003.
  • Dentcheva and Ruszczynski [2006] D. Dentcheva and A. Ruszczynski. Portfolio optimization with stochastic dominance constraints. Journal of Banking and Finance, 30(2):433–451, 2006.
  • Dhompongsa et al. [2017] S. Dhompongsa, W. K. Wong, H. T. Nguyen, and S. Sriboonchitta. Stochastic dominance and applications to finance, risk and economics. 2017.
  • Eberlein [2001] E. Eberlein. Application of generalized hyperbolic lévy motions to finance. In Lévy processes, pages 319–336. Springer, 2001.
  • Eberlein and Keller [1995] E. Eberlein and U. Keller. Hyperbolic distributions in finance. Bernoulli, pages 281–299, 1995.
  • Eberlein and Kluge [2006] E. Eberlein and W. Kluge. Exact pricing formulae for caps and swaptions in a lévy term structure model. Journal of computational finance, 9:99–125, 2006.
  • Eberlein and Riable [1999] E. Eberlein and S. Riable. Term structure models driven by general lévy processes. Mathematical Finance, 9:31–54, 1999.
  • Embrechts et al. [2021] P. Embrechts, A. Schied, and R. Wang. Robustness in the optimization of risk measures. Operations Research, 70(1):95–110, 2021.
  • Farkas et al. [2014] W. Farkas, P. Koch-Medina, and C. Munari. Beyond cash-additive risk measures: when changing the numeraire fails. Finance and Stochastic, 18(1):145–173, 2014.
  • Föllmer and Knispel [2013] H. Föllmer and T. Knispel. Convex risk measures: Basic facsts, law-invariance and beyond, asymptotics for large portfolios. Handbook of the fundamentals of financial decision making: Part II, 2013.
  • Föllmer and Schied [2004] H. Föllmer and A. Schied. Stochastic Finance: An Introduction in Discrete Time. Walter de Gruyter, 2004.
  • G.Svindland [2009] G.Svindland. Subgradients of law-invariant convex risk measures on l1l^{1}. Statistics and Decisions, 27(2):169–199, 2009.
  • Hammerstein [2010] E. Hammerstein. Generalized hyperbolic distributions: theory and applications to CDO pricing. PhD thesis, Citeseer, 2010.
  • Huang and Litzenberger [1988] C. F. Huang and R. H. Litzenberger. Foundations for financial economics, volume 1. Prentice-Hall, Inc., 1988.
  • Kaina and Rüschendorf [2009] M. Kaina and L. Rüschendorf. On convex risk measures on lpl^{p} spaces. Math Meth Oper Res, 69:475–495, 2009.
  • Kopa [2013] M. Kopa. Stochastic Dominance In Portfolio Efficiency Testing. Ph.D Thesis, 2013.
  • Landsman and Valdez [2003] Z. Landsman and E. A. Valdez. Tail conditional expectations for elliptical distributions. North American Actuarial Journal, 7:55–71, 2003.
  • Levy [2016] H. Levy. Stochastic Dominance, volume 3. Springer Nature, 2016.
  • Madan and Seneta [1990] D. Madan and E. Seneta. The variance gamma (vg) model for share market returns. Journal of business, pages 511–524, 1990.
  • Madan et al. [1998] D. Madan, P. Carr, and E. Chang. The variance gamma process and option pricing. Review of Finance, 2(1):79–105, 1998.
  • Mao and Wang [2020] T. Mao and R. Wang. Risk aversion in regulatory capital principles. SIAM Journal on Financial Mathematics, 11(1):169–200, 2020.
  • Mcneil et al. [2015] A. J. Mcneil, R. Frey, and P. Embrechts. Quantitative risk management: Concepts, techniques and tools: Revised edition. Introductory Chapters, 2015.
  • Ogryczak and Ruszczyński [1999] W. Ogryczak and A. Ruszczyński. From stochastic dominance to mean-risk models: Semideviations as risk measures. European Journal of Operational Research, 116:33–50, 1999.
  • Ogryczak and Ruszczyński [2001] W. Ogryczak and A. Ruszczyński. On consistency of stochastic dominance and mean–semideviation models. Math. Program., 89:217–232, 2001.
  • Panjer [2002] H. H. Panjer. Measurement of Risk, Solvency Requirements, and Allocation of Capital within Financial Conglomerates. PhD thesis, Institute of Insurance and Pension Research, University of Waterloo Research Report 01-15., 2002.
  • Rockafellar and Uryasev [2000] R. T. Rockafellar and S. Uryasev. Optimization of conditional value-at-risk. Journal of risk, 2:21–42, 2000.
  • Rockafellar and Uryasev [2002] R. T. Rockafellar and S. Uryasev. Conditional value-at-risk for general loss distributions. Journal of banking & finance, 26(7):1443–1471, 2002.
  • Shaked and Shanthikumar [1994] M. Shaked and J. Shanthikumar. Stochastic orders and their applications. New York: Academic Press, 1994.