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

Exponential utility maximization in small/large financial markets

Miklós Rásonyi1\text{Mikl\'{o}s R\'{a}sonyi}^{1} Hasanjan Sayit2\text{Hasanjan Sayit}^{2}
Rényi Institute, Budapest, Hungary1{}^{1}\text{Rényi Institute, Budapest, Hungary}
Xi’Jiao Liverpool University, Suzhou, China2{}^{2}\text{Xi'Jiao Liverpool University, Suzhou, China}
(September 20, 2021)
Abstract

Obtaining utility maximizing optimal portfolio in closed form is a challenging issue when the return vector follows a more general distribution than the normal one. In this note, we give closed form expression, in markets based on finitely many assets, for optimal portfolios that maximize exponential utility function when the return vector follows normal mean-variance mixture models. Especially, our approach expresses the closed form solution in terms of the Laplace transformation of the mixing distribution of the normal mean-variance mixture model and no any distributional assumptions on the mixing distribution are made.

We then consider large financial markets based on normal mean-variance mixture models also and show that the optimal exponential utilities based on small markets converge to the optimal exponential utility in the large financial market. This shows, in particular, that to reach the best utility level investors need to diversify their investment to include infinitely many assets into their portfolio and with portfolios based on only finitely many assets they never be able to reach the optimum level of utility.

Keywords: Expected utility; Mean-variance mixtures; Hara utility functions; Large financial markets; Martingale measures.

JEL Classification: G11

1 Introduction

We consider a frictionless financial market with d+1d+1 assets. We assume the first asset is a risk-free asset with risk-free interest rate rfr_{f} and the remaining dd assets are risky assets with returns modelled by an dd-dimensional random vector XX. In this note, we assume that XX follows normal mean-variance mixture (NMVM) distribution as follows

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

where μd\mu\in\mathbb{R}^{d} is location parameter, γd\gamma\in\mathbb{R}^{d} controls the skewness, ZGZ\sim G is a non-negative random variable with distribution function GG, Ad×dA\in\mathbb{R}^{d\times d} is a symmetric and positive definite d×dd\times d matrix of real numbers, NN(0,I)N\sim N(0,I) is a dd-dimensional Gaussian random vector with identity co-variance matrix II in d×d\mathbb{R}^{d}\times\mathbb{R}^{d}, and NN is independent from the mixing distribution ZZ.

In this paper we use the following notations. For any 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} in d\mathbb{R}^{d}, where the superscript TT stands for the transpose of a vector, <x,y>=xTy=i=1dxiyi<x,y>=x^{T}y=\sum_{i=1}^{d}x_{i}y_{i} denotes the scalar product of the vectors xx and yy, and |x|=i=1dxi2|x|=\sqrt{\sum_{i=1}^{d}x_{i}^{2}} denotes the Euclidean norm of the vector xx. We sometimes use the short hand notation XN(μ+γz,zΣ)GX\sim N(\mu+\gamma z,z\Sigma)\circ G for (1), where Σ=ATA\Sigma=A^{T}A. \mathbb{R} denotes the set of real numbers and +=[0,+)\mathbb{R}_{+}=[0,+\infty) denotes the set of non-negative real-numbers. Following the same notations of [13], 𝒥\mathcal{J} denotes the family of infinitely divisible random variables on +\mathbb{R}_{+}, 𝒮\mathcal{S} denotes the set of self-decomposable random variables on +\mathbb{R}_{+}, and 𝒢\mathcal{G} denotes the class of generalized gamma convolutions (GGCs) on +\mathbb{R}_{+} that will be introduced later. The Laplace transformation of any distribution GG is denoted by G(s)=esyG(dy)\mathcal{L}_{G}(s)=\int e^{-sy}G(dy). A gamma random variable with density function f(x)=1Γ(α)βαxα1ex/βf(x)=\frac{1}{\Gamma(\alpha)\beta^{\alpha}}x^{\alpha-1}e^{-x/\beta} is denoted by G=G(α,β)G=G(\alpha,\beta).

A prominent example of the NMVM models is generalized hyperbolic (GH) distributions, where the mixing distribution ZZ follows generalized inverse Gaussian (GIG) distribution denoted as GIG(λ,a,b)GIG(\lambda,a,b). The probability density function of a GIG distribution, denoted by fGIG(λ,a,b)f_{GIG}(\lambda,a,b), takes the following form

fGIG(x;λ,a,b)=(ba)λ1Kλ(ab)xλ1e12(a2x1+b2x)1(0,+)(x),f_{GIG}(x;\lambda,a,b)=(\frac{b}{a})^{\lambda}\frac{1}{K_{\lambda}(ab)}x^{\lambda-1}e^{-\frac{1}{2}(a^{2}x^{-1}+b^{2}x)}1_{(0,+\infty)}(x), (2)

where Kλ(x)K_{\lambda}(x) denotes the modified Bessel function of third kind with index λ\lambda and the allowed parameter ranges for λ,a,b\lambda,a,b in (2) are (i) a0,b>0a\geq 0,b>0 if λ>0\lambda>0, (ii) a>0,b0a>0,b\geq 0 if λ<0\lambda<0, (iii) a>0,b>0a>0,b>0 if λ=0\lambda=0. Here the case a=0a=0 in (i) or the case b=0b=0 in (ii) above need to be understood in limiting cases of (2) and in these special cases we have

fGIG(x;λ,0,b)=(b22)λxλ1Γ(λ)eb22x1(0,+)(x),λ>0,fGIG(x;λ,a,0)=(2a2)λxλ1Γ(λ)ea22x1(0,+)(x),λ<0,\begin{split}f_{GIG}(x;\lambda,0,b)&=(\frac{b^{2}}{2})^{\lambda}\frac{x^{\lambda-1}}{\Gamma(\lambda)}e^{-\frac{b^{2}}{2}x}1_{(0,+\infty)}(x),\;\;\;\lambda>0,\\ f_{GIG}(x;\lambda,a,0)&=(\frac{2}{a^{2}})^{\lambda}\frac{x^{\lambda-1}}{\Gamma(-\lambda)}e^{-\frac{a^{2}}{2x}}1_{(0,+\infty)}(x),\;\;\;\lambda<0,\end{split} (3)

where Γ(x)\Gamma(x) denotes the Gamma function. Here fGIG(x;λ,0,b)f_{GIG}(x;\lambda,0,b) is the density function of a Gamma distribution G(λ,2b2)G(\lambda,\frac{2}{b^{2}}) and fGIG(x;λ,a,0)f_{GIG}(x;\lambda,a,0) is the density function of a inverse Gamma distribution iG(λ,a22)iG(\lambda,\frac{a^{2}}{2}).

The GH distribution in dimension dd is denoted by GHd(λ,α,β,δ,μ,Σ)GH_{d}(\lambda,\alpha,\beta,\delta,\mu,\Sigma) and it satisfies GHd(λ,α,β,δ,μ,Σ)N(μ+zΣβ,zΣ)GIG(λ,δ,α2βTΣβ)GH_{d}(\lambda,\alpha,\beta,\delta,\mu,\Sigma)\sim N(\mu+z\Sigma\beta,z\Sigma)\circ GIG(\lambda,\delta,\sqrt{\alpha^{2}-\beta^{T}\Sigma\beta}). The parameter ranges of this distribution is λ,α,δ+,β,μd\lambda\in\mathbb{R},\;\alpha,\delta\in\mathbb{R}_{+},\;\beta,\mu\in\mathbb{R}^{d} and (i’) δ0, 0βTΣβ<α\delta\geq 0,\;0\leq\sqrt{\beta^{T}\Sigma\beta}<\alpha if λ>0\lambda>0, (ii’) δ>0, 0βTΣβ<α\delta>0,\;0\leq\sqrt{\beta^{T}\Sigma\beta}<\alpha if λ=0\lambda=0, (iii’) δ>0, 0βTΣβα\delta>0,\;0\leq\sqrt{\beta^{T}\Sigma\beta}\leq\alpha if λ<0\lambda<0. The class of GH distributions include two popular models in finance: if λ=12\lambda=-\frac{1}{2} we have normal inverse Gaussian distribution which is denoted by NIGd(α,β,δ,μ,Σ)NIG_{d}(\alpha,\beta,\delta,\mu,\Sigma) and when λ=1+d2\lambda=\frac{1+d}{2} we have the class of hyperbolic distributions denoted by HYPd(α,β,δ,μ,Σ)HYP_{d}(\alpha,\beta,\delta,\mu,\Sigma). As in the case of the GIG distributions, the case δ=0\delta=0 in (i’) above and the case βTΣβ=α\sqrt{\beta^{T}\Sigma\beta}=\alpha or α=0\alpha=0 in (iii’) above needs to be understood as limiting cases of the GH distributions. If λ>0,δ0\lambda>0,\delta\rightarrow 0 in case (i’) above then

GHd(λ,α,β,δ,μ,Σ)𝑤Nd(μ+zΣβ,zΣ)G(λ,α2βTΣβ2)=:VGd(λ,α,β,μ,Σ),GH_{d}(\lambda,\alpha,\beta,\delta,\mu,\Sigma)\overset{w}{\rightarrow}N_{d}(\mu+z\Sigma\beta,z\Sigma)\circ G(\lambda,\frac{\alpha^{2}-\beta^{T}\Sigma\beta}{2})=:VG_{d}(\lambda,\alpha,\beta,\mu,\Sigma), (4)

where =𝑤\overset{w}{=} denotes weak convergence of distributions and VGdVG_{d} represents the class of variance gamma distributions. If λ<0\lambda<0 and α0\alpha\rightarrow 0 as well as β0\beta\rightarrow 0 in case (iii’) above we have the shifted tt distributions with degrees of freedom 2λ-2\lambda

GHd(λ,α,β,δ,μ,Σ)𝑤N(μ,zΣ)iG(λ,δ22)=:td(λ,δ,μ,Σ).GH_{d}(\lambda,\alpha,\beta,\delta,\mu,\Sigma)\overset{w}{\rightarrow}N(\mu,z\Sigma)\circ iG(\lambda,\frac{\delta^{2}}{2})=:t_{d}(\lambda,\delta,\mu,\Sigma). (5)

If α,δ\alpha\rightarrow\infty,\delta\rightarrow\infty and δασ2<\frac{\delta}{\alpha}\rightarrow\sigma^{2}<\infty, we have the following that shows that the Normal random vectors are limiting cases of the GH distributions

GHd(λ,α,β,δ,μ,Σ)𝑤N(μ+zΣβ,zΣ)ϵσ2=:N(μ+σ2Σβ,σ2Σ),GH_{d}(\lambda,\alpha,\beta,\delta,\mu,\Sigma)\overset{w}{\rightarrow}N(\mu+z\Sigma\beta,z\Sigma)\circ\epsilon_{\sigma^{2}}=:N(\mu+\sigma^{2}\Sigma\beta,\sigma^{2}\Sigma), (6)

where ϵσ2\epsilon_{\sigma^{2}} is the dirac function that equals to 11 when z=σ2z=\sigma^{2} and equals to zero otherwise, see Chapter 2 of [10] for the details of these. All these normal inverse Gaussian, hyperbolic, variance gamma, and student tt distributions are very popular models in finance, see [12], [1], [3], [8], [11], [21], [20], [14], [22] for this.

The class of GIG distributions belong to the class of GGCs. A positive random variable ZZ is a GGC, without translation term, if there exists a positive Radon measure ν\nu on +\mathbb{R}_{+} such that

Z(s)=EesZ=e0ln(1+sz)ν(dz),\mathcal{L}_{Z}(s)=Ee^{-sZ}=e^{-\int_{0}^{\infty}\ln(1+\frac{s}{z})\nu(dz)}, (7)

with

01|lnx|ν(dx)<,11xν(dx)<.\int_{0}^{1}|lnx|\nu(dx)<\infty,\;\;\int_{1}^{\infty}\frac{1}{x}\nu(dx)<\infty. (8)

The measure ν\nu is called Thorin’s measure associated with ZZ. For the definition of the GGCs see the survey paper [13]. In Proposition 1.1 of [13], it was shown that any GGC random variable can be written as Wiener-Gamma integral

Z=0h(s)𝑑γs,Z=\int_{0}^{\infty}h(s)d\gamma_{s}, (9)

where h(s):++h(s):\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} is a deterministic function with 0ln(1+h(s))𝑑s<\int_{0}^{\infty}ln(1+h(s))ds<\infty and {γs}\{\gamma_{s}\} is a standard Gamma process with Lévy measure exdxx,x>0e^{-x}\frac{dx}{x},x>0.

Proposition 1.23 of [10] shows that the class of GIG random variables belongs to the class GGC. It provides the description of the corresponding Thorin’s measures (in terms of the functions UGIGU_{GIG} in the Proposition) for all the cases of parameters of GIG. The class of GGC distributions are rich as stated in the introduction of [13] and we have the relation 𝒢𝒮𝒥\mathcal{G}\subset\mathcal{S}\subset\mathcal{J}. In our model (1) the mixing distribution ZZ can be any distribution in 𝒥\mathcal{J}. In fact, ZZ can be any non-negative random variable.

Given an initial endowment W0>0W_{0}>0, the investor must determine the portfolio weights x on the dd risky assets such that the expected utility of the next period wealth is maximized. The wealth that corresponds to portfolio weight xx on the risky assets is given by

W(x)=W0[1+(1xT1)rf+xTX]=W0(1+rf)+W0[xT(X1rf)]\begin{split}W(x)=&W_{0}[1+(1-x^{T}1)r_{f}+x^{T}X]\\ =&W_{0}(1+r_{f})+W_{0}[x^{T}(X-\textbf{1}r_{f})]\end{split} (10)

and the investor’s problem is

maxxDEU(W(x)),\max_{x\in D}\;EU(W(x)),\\ (11)

for some domain DD of the portfolio set DD. Note here that xx represents the portfolio weights on the risky assets and 1xT11-x^{T}\textbf{1} is the proportion of the initial wealth invested on the risk free asset. The portfolio weights xx on risky assets are allowed to be any vector in DD.

The main goal of this paper is to discuss the solution of the problem (11) for exponential utility function UU when the returns of the risky assets have NMVM distribution as in (1). These type of utility maximization problems in one period models were studied in many papers in the past, see [17], [18], [15], [29], [2]. Especially, the recent paper [3] made an interesting observation that, with generalized hyperbolic models and with exponential utility, the optimal portfolios of the corresponding expected utility maximization problems can be written as a sum of two portfolios that are determined by the location and skewness parameters of the model (1) separately. The present paper, extends their result to more general class of NMVM models as a compliment.

The paper is organized as follows. In section 2 below we present closed form solution for optimal portfolio when the utility function UU is exponential. In section 3, we show that the optimal expected utilities in small financial markets converge to an overall best expected utility in a large financial market. In section 4 we present examples as applications of our results.

2 Closed form solution for optimal portfolios under exponential utility

In this section, we study the solution of the problem (11) when the utility function of the investor is exponential

U(W)=eaW,a>0,U(W)=-e^{-aW},a>0, (12)

and when the investment opportunity set consists of the above stated d+1d+1 assets. Below we obtain an expression that relates EU(W)EU(W) to the Laplace transformation of the mixing distribution ZZ as in (14) below. First observe that we have

W(x)=𝑑W0(1+rf)+W0[xT(μ1rf)+xTγZ+xTΣxZN(0,1)].W(x)\overset{d}{=}W_{0}(1+r_{f})+W_{0}[x^{T}(\mu-\textbf{1}r_{f})+x^{T}\gamma Z+\sqrt{x^{T}\Sigma x}\sqrt{Z}N(0,1)]. (13)
Lemma 2.1.

For any portfolio xdx\in\mathbb{R}^{d} such that EU(W(x))EU(W(x)) is finite we have

EU(W(x))=eaW0(1+rf)eaW0xT(μ1rf)Z(aW0xTγa2W022xTΣx),EU(W(x))=-e^{-aW_{0}(1+r_{f})}e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\mathcal{L}_{Z}\Big{(}aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x\Big{)}, (14)

where Z(s)=EesZ\mathcal{L}_{Z}(s)=Ee^{-sZ} is the Laplace transformation of ZZ.

Proof.

From (13), we have

EU(W(x))=EeaW0(1+rf)aW0[xT(μ1rf)+xTγZ+xTΣxZN(0,1)]=eaW0(1+rf)eaW0xT(μ1rf)0+EeaW0xTγzaW0xTΣxzN(0,1)fZ(z)𝑑z=eaW0(1+rf)eaW0xT(μ1rf)0+eaW0xTγzEeaW0xTΣxzN(0,1)fZ(z)𝑑z=eaW0(1+rf)eaW0xT(μ1rf)0+eaW0xTγzea2W022xTΣxzfZ(z)𝑑z=eaW0(1+rf)eaW0xT(μ1rf)0+e(aW0xTγa2W022xTΣx)zfZ(z)𝑑z=eaW0(1+rf)eaW0xT(μ1rf)Z(aW0xTγa2W022xTΣx).\begin{split}EU(W(x))=&-Ee^{-aW_{0}(1+r_{f})-aW_{0}\Big{[}x^{T}(\mu-\textbf{1}r_{f})+x^{T}\gamma Z+\sqrt{x^{T}\Sigma x}\sqrt{Z}N(0,1)\Big{]}}\\ =&-e^{-aW_{0}(1+r_{f})}e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\int_{0}^{+\infty}Ee^{-aW_{0}x^{T}\gamma z-aW_{0}\sqrt{x^{T}\Sigma x}\sqrt{z}N(0,1)}f_{Z}(z)dz\\ =&-e^{-aW_{0}(1+r_{f})}e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\int_{0}^{+\infty}e^{-aW_{0}x^{T}\gamma z}Ee^{-aW_{0}\sqrt{x^{T}\Sigma x}\sqrt{z}N(0,1)}f_{Z}(z)dz\\ =&-e^{-aW_{0}(1+r_{f})}e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\int_{0}^{+\infty}e^{-aW_{0}x^{T}\gamma z}e^{\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma xz}f_{Z}(z)dz\\ =&-e^{-aW_{0}(1+r_{f})}e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\int_{0}^{+\infty}e^{-(aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x)z}f_{Z}(z)dz\\ =&-e^{-aW_{0}(1+r_{f})}e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\mathcal{L}_{Z}(aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x).\\ \end{split}

Remark 2.2.

If μ1rf=0\mu-\textbf{1}r_{f}=0 in our model (1), from (14) we have

EU(W(x))=eaW0(1+rf)Z(aW0xTγa2W022xTΣx).EU(W(x))=-e^{-aW_{0}(1+r_{f})}\mathcal{L}_{Z}\Big{(}aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x\Big{)}.

Since Z(s)\mathcal{L}_{Z}(s) is a strictly decreasing function, the expected utility maximization problem becomes the maximization problem of the quadratic function aW0xTγa2W022xTΣxaW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x in this case. Especially, if the risk-free interest rate rfr_{f} is zero and our model (1) is such that the location parameter μ\mu is zero, then the utility optimizing portfolio can be found by optimizing a quadratic function. Therefore for the rest of the paper, we assume that our model (1) is such that μ1rf0\mu-\textbf{1}r_{f}\neq 0. Also we assume that Z0Z\neq 0 with positive probability.

Remark 2.3.

By using the relation (11) and by checking the first order condition for optimality it is easy to see that the optimal portfolio xx^{\star} satisfies the following relation

x=1aW0[Σ1γZ(g(x))Z(g(x))Σ1(μ1rf)],x^{\star}=\frac{1}{aW_{0}}[\Sigma^{-1}\gamma-\frac{\mathcal{L}_{Z}(g(x^{\star}))}{\mathcal{L}_{Z}^{\prime}(g(x^{\star}))}\Sigma^{-1}(\mu-\textbf{1}r_{f})], (15)

where g(x)g(x) is given in the expression (16) below. There are several questions that one needs to address when applying the direct approach (15) in obtaining the optimal portfolio xx^{\star}: (i) if the function xEU(W(x))x\rightarrow EU(W(x)) is continuously differentiable (ii) if the optimal portfolio is the interior point of the corresponding domain (iii) if the equation (15) has unique solution. After these questions are addressed the next challenge becomes how to compute xx^{\star} numerically. This problem is not trivial if the dimension dd is a large number, i.e., xdx\in\mathbb{R}^{d} for large dd. To overcome these problems, in this paper we take different approach and obtain xx^{\star} in near closed form: to calculate xx^{\star} we only need to find the minimizing point of a convex function on the real line.

The above Lemma 2.1, expresses the expected utility in terms of a linear function xT(μ1rf)x^{T}(\mu-\textbf{1}r_{f}) and a quadratic function aW0xTγa2W022xTΣxaW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x of the portfolio xnx\in\mathbb{R}^{n}. For convenience, we introduce the following notations

g(x)=:aW0xTγa2W022xTΣx,G(x)=:eaW0xT(μ1rf)Z(aW0xTγa2W022xTΣx),=eaW0xT(μ1rf)Z(g(x)).\begin{split}g(x)=:&aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x,\\ G(x)=:&e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\mathcal{L}_{Z}\Big{(}aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x\Big{)},\\ =&e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\mathcal{L}_{Z}\Big{(}g(x)\Big{)}.\end{split} (16)

Then the relation (14) becomes

EU(W)=eaW0(1+rf)G(x)=eaW0(1+rf)eaW0xT(μ1rf)Z(g(x)).EU(W)=-e^{aW_{0}(1+r_{f})}G(x)=-e^{aW_{0}(1+r_{f})}e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\mathcal{L}_{Z}\big{(}g(x)\big{)}. (17)

Therefore we have the following obvious relation

argmaxxDEU(W)=argminxDG(x)\arg\max_{x\in D}EU(W)=\arg\min_{x\in D}G(x) (18)

for any domain DdD\in\mathbb{R}^{d} of the portfolio set. Note here that the equality in (18) means the equality of two sets if the optimizing points are more than one.

Our goal in this section is to give closed form solution for the problem (11) for some domains of the portfolio set. Before we start our analysis, we first present the following example.

Example 2.4.

Consider the model (1) with γ=0\gamma=0 and with the mixing distribution ZeN(0,1)Z\sim e^{N(0,1)}. Then for any x0x\neq 0 we have

EU(W(x))=.EU(W(x))=-\infty.

To see this, assume that there is a x0x\neq 0 such that EU(W(x))EU(W(x)) is finite. Then by Lemma 2.1 we have

EU(W(x))=eaW0(1+rf)eaW0xT(μ1rf)Z(a2W022xTΣx).EU(W(x))=-e^{-aW_{0}(1+r_{f})}e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\mathcal{L}_{Z}\Big{(}-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x\Big{)}.

For any x0x\neq 0 we have xTΣx>0x^{T}\Sigma x>0 as Σ\Sigma is positive definite by the assumption of the model (1). Now it is well known that when ZeN(0,1)Z\sim e^{N(0,1)} we have Z(s)=+\mathcal{L}_{Z}(s)=+\infty whenever s<0s<0. Therefore Z(a2W022xTΣx)=+\mathcal{L}_{Z}\Big{(}-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x\Big{)}=+\infty whenever x0x\neq 0 and this contradicts with the finiteness assumption of EU(W(x))EU(W(x)) made above. Thus we have EU(W(x))=EU(W(x))=-\infty whenever x0x\neq 0. Therefore the problem (11) does not have a solution when the domain DD does not include the zero vector in it. But if 0D0\in D, then x=0x=0 is the optimal portfolio and maxxDEU(W(x))=eaW0(1+rf)\max_{x\in D}\;EU(W(x))=-e^{-aW_{0}(1+r_{f})}. This case corresponds to investing all the initial wealth W0W_{0} on the risk-free asset as an optimal portfolio. We remark here that since γ=0\gamma=0 by Jensen’s inequality we have

EU(W(x))U(EW(x))=U(W0(1+rf)+W0xT(μ1rf)).EU(W(x))\leq U(EW(x))=U(W_{0}(1+r_{f})+W_{0}x^{T}(\mu-\textbf{1}r_{f})).

From this relation it is difficult to see that 0 is the expected utility optimizing portfolio when ZeN(0,1)Z\sim e^{N(0,1)}. But with the assistance of Lemma 2.1 above it becomes trivial to determine that 0 is the optimal portfolio as discussed earlier.

The above Example 2.4 shows that when the model (1) satisfies the conditions in the example and when 0D0\in D, the zero portfolio x=0x=0 is an optimal portfolio as when x0x\neq 0 one has EU(W(x))=EU(W(x))=-\infty always. It is obvious that, in this case, the function xEU(W(x))x\rightarrow EU(W(x)) is not differentiable at x=0x=0. Therefore we call x=0x=0 irregular solution for the optimization problem (18). Before we give formal definition of irregularity, we first introduce the following definition.

Definition 2.5.

For any mixing distribution ZZ, if Z(s)<\mathcal{L}_{Z}(s)<\infty for all ss\in\mathbb{R} we set s^=\hat{s}=-\infty and if Z(s)<\mathcal{L}_{Z}(s)<\infty for some ss\in\mathbb{R} and Z(s)=+\mathcal{L}_{Z}(s)=+\infty for some ss\in\mathbb{R}, we let s^\hat{s} be the real number such that

Z(s)=EesZ<,s>s^andZ(s)=EesZ=+,s<s^.\mathcal{L}_{Z}(s)=Ee^{-sZ}<\infty,\;\forall s>\hat{s}\;\;\mbox{and}\;\;\mathcal{L}_{Z}(s)=Ee^{-sZ}=+\infty,\;\forall s<\hat{s}. (19)

We call s^\hat{s} the critical value (CV) (we use the acronym CV-L from now on, where LL implies that it is CV in the context of Laplace transformation. One can also define this CV in the context of moment generating functions and in this case an acronym CV-M can be used) of ZZ under Laplace transformation in this paper. Observe that since ZZ is non-negative random variable we always have s^0\hat{s}\leq 0.

Remark 2.6.

In the above definition 2.5, the value of Z(s)\mathcal{L}_{Z}(s) at s=s^s=\hat{s} is not specified. Both of the cases Z(s^)<\mathcal{L}_{Z}(\hat{s})<\infty and Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty are possible. For example if ZeN(0,1)Z\sim e^{N(0,1)}, then s^=0\hat{s}=0 and clearly Z(0)=1<\mathcal{L}_{Z}(0)=1<\infty. If Zxα1ex/β/[Γ(α)βα]Z\sim x^{\alpha-1}e^{-x/\beta}/[\Gamma(\alpha)\beta^{\alpha}] is a Gamma distribution, then Z(s)=1/[(1+βs)α]\mathcal{L}_{Z}(s)=1/[(1+\beta s)^{\alpha}]. In this case s^=1/β\hat{s}=-1/\beta and we have Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty.

Below we define some domains for the portfolio set.

Sa=:{xd:aW0xTγa2W022xTΣx>s^},Sa=:{xd:aW0xTγa2W022xTΣx=s^},S¯a=:SaSa.\begin{split}S_{a}=:&\{x\in\mathbb{R}^{d}:aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x>\hat{s}\},\\ \partial S_{a}=:&\{x\in\mathbb{R}^{d}:aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x=\hat{s}\},\\ \bar{S}_{a}=:&S_{a}\cup\partial S_{a}.\end{split} (20)
Remark 2.7.

Our main objective in this section is to find closed form solution for the optimal portfolio for the problem

maxxdEU(W(x)).\max_{x\in\mathbb{R}^{d}}\;EU(W(x)). (21)

The following relations are easy to see

maxxdEU(W(x))=maxxSaEU(W(x)),\max_{x\in\mathbb{R}^{d}}\;EU(W(x))=\max_{x\in S_{a}}EU(W(x)), (22)

if Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty and

maxxdEU(W(x))=maxxS¯aEU(W(x)),\max_{x\in\mathbb{R}^{d}}\;EU(W(x))=\max_{x\in\bar{S}_{a}}EU(W(x)), (23)

if Z(s^)<+\mathcal{L}_{Z}(\hat{s})<+\infty. Observe here that if s^<0\hat{s}<0, then SaS_{a} is a nonempty set as the zero vector x=0x=0 is in it. If s^=0\hat{s}=0, then the set S¯a\bar{S}_{a} is nonempty as x=0x=0 is in it.

In this section we attempt to give closed form solutions for the problems (22) and (23) above. Our approach for this is based on the following idea: we fix the term xT(μ1rf)x^{T}(\mu-\textbf{1}r_{f}) at some constant level cc and optimize the quadratic term aW0xTγa2W022xTΣxaW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x in (14). More specifically, we solve the following optimization problem

maxxaW0xTγa2W022xTΣx,s.t.xT(μrf1)=c,\begin{split}\max_{x}\;\;&aW_{0}x^{T}\gamma-\frac{a^{2}W_{0}^{2}}{2}x^{T}\Sigma x,\\ s.t.\;\;&x^{T}(\mu-r_{f}\textbf{1})=c,\\ \end{split} (24)

first and plug in the solution, which we denote by xcx_{c}, into the expression (14) so that the utility maximization problem becomes an optimization problem of a function of one variable cc.

Lemma 2.8.

Consider the optimization problem (21). Let x¯d\bar{x}\in\mathbb{R}^{d} be a solution for this problem. Then x¯\bar{x} solves (24) for some cc.

Proof.

Define c¯=:x¯T(μłrf)\bar{c}=:\bar{x}^{T}(\mu-\l r_{f}). Let x~\tilde{x} be the solution to the problem (24) with cc replaced by c¯\bar{c} (here the solution is unique as Σ\Sigma is positive definite by assumption). By the optimality of x~\tilde{x}, we have g(x¯)g(x~)g(\bar{x})\leq g(\tilde{x}). Since Z(s)\mathcal{L}_{Z}(s) is a decreasing function we have Z(g(x~))Z(g(x¯))\mathcal{L}_{Z}(g(\tilde{x}))\leq\mathcal{L}_{Z}(g(\bar{x})). Since c¯=x¯T(μłrf)=x~T(μłrf)\bar{c}=\bar{x}^{T}(\mu-\l r_{f})=\tilde{x}^{T}(\mu-\l r_{f}) we have G(x~)G(x¯)G(\tilde{x})\leq G(\bar{x}). This shows that EU(W(x~))EU(W(x¯))EU(W(\tilde{x}))\geq EU(W(\bar{x})). But x¯\bar{x} is optimal for (11) with D=dD=\mathbb{R}^{d}. Therefore we should have EU(W(x~))=EU(W(x¯))EU(W(\tilde{x}))=EU(W(\bar{x})). This implies G(x~)=G(x¯)G(\tilde{x})=G(\bar{x}) and this in turn implies g(x¯)=g(x~)g(\bar{x})=g(\tilde{x}) again due to c¯=x¯T(μłrf)=x~T(μłrf)\bar{c}=\bar{x}^{T}(\mu-\l r_{f})=\tilde{x}^{T}(\mu-\l r_{f}). The uniqueness of the optimization point for (24) then implies x¯=x~\bar{x}=\tilde{x}. ∎

Remark 2.9.

The Lemma 2.8 above gives a characterization of the optimal portfolios for the problem (11). But it doesn’t tell us if the optimal portfolio for the problem (2.8) is unique. It shows only that any optimal portfolio for the problem (11) solves a quadratic optimization problem (24) for some appropriate cc. Now consider the case of example 2.4. In the setting of this example, consider the utility maximization problem (11). Since 0d0\in\mathbb{R}^{d}, as explained in the Example 2.4, the vector x^=0\hat{x}=0 is the solution for the optimization problem (11). Now let xx^{\star} be the optimal solution for the problem (24) with c=0c=0 (which means (x)T(μrf1))=0(x^{\star})^{T}(\mu-r_{f}\textbf{1}))=0). Then we should have g(x)g(x^)g(x^{\star})\geq g(\hat{x}). But if g(x)>g(x^)g(x^{\star})>g(\hat{x}), then x^=0\hat{x}=0 can not be optimal solution for (11). Therefore we should have g(x)=g(x^)g(x^{\star})=g(\hat{x}). The uniqueness of the optimal solution for (24) with c=0c=0 then implies x=x^=0x^{\star}=\hat{x}=0.

Definition 2.10.

Consider the optimization problem (11) for some given model (1) and for some domain DdD\subset\mathbb{R}^{d}. Let s^\hat{s} denote the CV-L of the mixing distribution ZZ. Let xDx^{\star}\in D be a solution for (11). We say that xx^{\star} is irregular if g(x)=s^g(x^{\star})=\hat{s}. If g(x)>s^g(x^{\star})>\hat{s}, we call the solution xx^{\star} regular.

Remark 2.11.

Clearly the definition of irregular and regular solutions depend on the CV-L number s^\hat{s} of the mixing distribution ZZ in (1). If Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty, then the solution to (11) can not be irregular. Therefore, the irregularity can happen only when Z(s^)<+\mathcal{L}_{Z}(\hat{s})<+\infty. Observe that the solution x=0x=0 in Example 2.4 is an irregular solution.

Remark 2.12.

Consider the optimization problem (11). From Lemma 2.8, any optimal portfolio xx^{\star} is a solution for the quadratic optimization problem (24) with xT(μrf1)=cx^{T}(\mu-r_{f}\textbf{1})=c^{\star} for some fixed cc^{\star}. If xx^{\star} is irregular, then g(x)=s^g(x^{\star})=\hat{s}. The optimality and uniqueness (on the hyperplane xT(μrf1)=cx^{T}(\mu-r_{f}\textbf{1})=c^{\star}) of xx^{\star} implies that we have g(x)<g(x)=s^g(x)<g(x^{\star})=\hat{s} for all xxx\neq x^{\star} on the hyperplane xT(μrf1)=cx^{T}(\mu-r_{f}\textbf{1})=c^{\star}. Therefore we have EU(W(x))=EU(W(x))=-\infty for all xxx\neq x^{\star} on the hyperplane xT(μrf1)=cx^{T}(\mu-r_{f}\textbf{1})=c^{\star}. From this we conclude that if the optimal portfolio for the problem (24) is irregular, then any small neighborhood of this portfolio contains some portfolios with infinite expected utility. In comparison, if the optimal portfolio is regular, then it has a small ball around it with finite expected value for each portfolio in this small ball.

As it was shown in Lemma 2.8, the solutions of the utility maximization (11) can be obtained by solving the quadratic optimization problem (24). For a given optimization problem (11), if we know the corresponding cc in (24) such that the solution of (24) is the solution of (11), then we just need to solve the optimization problem (24) to obtain the optimal portfolio. But figuring out such an cc is not a trivial issue. We first prove following Lemma.

Lemma 2.13.

For any real number cc, when xT(μ1rf)=cx^{T}(\mu-\textbf{1}r_{f})=c, the maximizing point xcx_{c} of g(x)g(x) is given by

xc=1aW0[Σ1γqcΣ1(μ1rf)],x_{c}=\frac{1}{aW_{0}}\Big{[}\Sigma^{-1}\gamma-q_{c}\Sigma^{-1}(\mu-\textbf{1}r_{f})\Big{]}, (25)

and we have

g(xc)=12γTΣ1γqc22(μ1rf)TΣ1(μ1rf),g(x_{c})=\frac{1}{2}\gamma^{T}\Sigma^{-1}\gamma-\frac{q^{2}_{c}}{2}(\mu-\textbf{1}r_{f})^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f}), (26)

where

qc=γTΣ1(μ1rf)aW0c(μ1rf)TΣ1(μ1rf).q_{c}=\frac{\gamma^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f})-aW_{0}c}{(\mu-\textbf{1}r_{f})^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f})}. (27)
Proof.

We form the Lagrangian L=g(x)+λ(cxT(μ1rf))L=g(x)+\lambda(c-x^{T}(\mu-\textbf{1}r_{f})) with the Lagrangian parameter λ\lambda. Denoting the maximizing point by xcx_{c}, the first order condition gives

xc=1aW0Σ1γλa2W02Σ1(μ1rf).x_{c}=\frac{1}{aW_{0}}\Sigma^{-1}\gamma-\frac{\lambda}{a^{2}W_{0}^{2}}\Sigma^{-1}(\mu-\textbf{1}r_{f}). (28)

We plug xcx_{c} into xcT(μ1rf)=cx^{T}_{c}(\mu-\textbf{1}r_{f})=c and obtain

c=1aW0γTΣ1(μ1rf)λa2W02(μ1rf)TΣ1(μ1rf).c=\frac{1}{aW_{0}}\gamma^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f})-\frac{\lambda}{a^{2}W_{0}^{2}}(\mu-\textbf{1}r_{f})^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f}). (29)

From this we find λ\lambda as follows

λ=aW0γTΣ1(μ1rf)ca2W02(μ1rf)TΣ1(μ1rf).\lambda=\frac{aW_{0}\gamma^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f})-ca^{2}W_{0}^{2}}{(\mu-\textbf{1}r_{f})^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f})}. (30)

Then we plug λ\lambda into the expression (28) of xcx_{c} above and obtain (25). To obtain (26), we plug xcx_{c} into g(x)g(x) in (16). After doing some algebra we obtain

g(xc)=12γTΣ1γ12qc2(μ1rf)TΣ1(μ1rf),g(x_{c})=\frac{1}{2}\gamma^{T}\Sigma^{-1}\gamma-\frac{1}{2}q_{c}^{2}(\mu-\textbf{1}r_{f})^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f}), (31)

with qcq_{c} given as in (27). This completes the proof. ∎

For the rest of the paper, as in [3], for convenience, we use the following notations

𝒜=γTΣ1γ,𝒞=(μ1rf)TΣ1(μ1rf),=γTΣ1(μ1rf).\mathcal{A}=\gamma^{T}\Sigma^{-1}\gamma,\;\mathcal{C}=(\mu-\textbf{1}r_{f})^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f}),\;\mathcal{B}=\gamma^{T}\Sigma^{-1}(\mu-\textbf{1}r_{f}). (32)

We first observe that 𝒞>0\mathcal{C}>0 due to the assumption in Remark 2.2 and the assumption on positive definiteness of Σ\Sigma. With these notations we have

g(xc)=𝒜2qc22𝒞,qc=𝒞aW0𝒞c.g(x_{c})=\frac{\mathcal{A}}{2}-\frac{q_{c}^{2}}{2}\mathcal{C},\;\;q_{c}=\frac{\mathcal{B}}{\mathcal{C}}-\frac{aW_{0}}{\mathcal{C}}c. (33)

From the relation (33), we express cc as a function of qcq_{c} as follows

c=1aW0[𝒞qc].c=\frac{1}{aW_{0}}[\mathcal{B}-\mathcal{C}q_{c}]. (34)

We define the following function

Q(θ)=e𝒞θZ[12𝒜θ22𝒞],Q(\theta)=e^{\mathcal{C}\theta}\mathcal{L}_{Z}\Big{[}\frac{1}{2}\mathcal{A}-\frac{\theta^{2}}{2}\mathcal{C}\Big{]}, (35)

and we define θ^=:𝒜2s^𝒞\hat{\theta}=:\sqrt{\frac{\mathcal{A}-2\hat{s}}{\mathcal{C}}}, where s^\hat{s} is the IN of ZZ. If s^=\hat{s}=-\infty, the θ^\hat{\theta} is understood to be equal to ++\infty. Note here that s^0\hat{s}\leq 0 as ZZ is non-negative random variable. Therefore θ^\hat{\theta} is well defined. If Z(s^)<+\mathcal{L}_{Z}(\hat{s})<+\infty, Q(θ)Q(\theta) is finite iff 12𝒜θ22𝒞s^\frac{1}{2}\mathcal{A}-\frac{\theta^{2}}{2}\mathcal{C}\geq\hat{s} and this translates into: Q(θ)Q(\theta) is finite iff θ[θ^,θ^]\theta\in[-\hat{\theta},\hat{\theta}]. If Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty, Q(θ)Q(\theta) is finite iff 12𝒜θ22𝒞>s^\frac{1}{2}\mathcal{A}-\frac{\theta^{2}}{2}\mathcal{C}>\hat{s} and this translates into: Q(θ)Q(\theta) is finite iff θ(θ^,θ^)\theta\in(-\hat{\theta},\hat{\theta}).

Next we prove the following Lemma that relates QQ to GG.

Lemma 2.14.

Let xcx_{c} be the solution for the problem (24) for a given cc. Assume xcSax_{c}\in S_{a} if Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty and xcS¯ax_{c}\in\bar{S}_{a} if Z(s^)<+\mathcal{L}_{Z}(\hat{s})<+\infty. Then for any xx with xT(μ1rf)=cx^{T}(\mu-\textbf{1}r_{f})=c, we have

eQ(qc)G(x),e^{-\mathcal{B}}Q(q_{c})\leq G(x), (36)

where qcq_{c} is given by (27) and \mathcal{B} is given by (32). We also have eQ(qc)=G(xc)e^{-\mathcal{B}}Q(q_{c})=G(x_{c}).

Proof.

Note that G(x)=eaW0xT(μ1rf)Z(g(x))G(x)=e^{-aW_{0}x^{T}(\mu-\textbf{1}r_{f})}\mathcal{L}_{Z}(g(x)). The stated conditions on xcx_{c} in the Lemma insures that G(xc)=eaW0cZ(g(xc))G(x_{c})=e^{-aW_{0}c}\mathcal{L}_{Z}(g(x_{c})) is finite. Since g(x)g(xc)g(x)\leq g(x_{c}) for any xx with xT(μ1rf)=cx^{T}(\mu-\textbf{1}r_{f})=c by the definition of xcx_{c} (the optimizing point) and also since Z(s)\mathcal{L}_{Z}(s) is a decreasing function of ss we have

G(xc)G(x)G(x_{c})\leq G(x) (37)

for any xx with xT(μ1rf)=cx^{T}(\mu-\textbf{1}r_{f})=c. We plug the cc in (34) into the expression of G(xc)G(x_{c}) and obtain

G(xc)=ee𝒞qcZ[12𝒜qc22𝒞]=eQ(qc).G(x_{c})=e^{-\mathcal{B}}e^{\mathcal{C}q_{c}}\mathcal{L}_{Z}\Big{[}\frac{1}{2}\mathcal{A}-\frac{q_{c}^{2}}{2}\mathcal{C}\Big{]}=e^{-\mathcal{B}}Q(q_{c}). (38)

Remark 2.15.

The above Lemma 2.14 shows that the function G(x)G(x) achieves its unique (as the solution for (24) is unique in a hyperplane) minimum value on the hyperplane xT(μrf1)=cx^{T}(\mu-r_{f}\textbf{1})=c at xcx_{c} and its minimum value is given by eQ(qc)e^{\mathcal{-B}}Q(q_{c}) with qcq_{c} in (33). For any θ0[θ^,θ^]\theta_{0}\in[-\hat{\theta},\hat{\theta}], we can let c0c_{0} be such that qc0=θ0q_{c_{0}}=\theta_{0}. Let x0x_{0} be the optimal solution of (24) with cc replaced by c0c_{0}. From Lemma 2.13, we have g(x0)=12𝒜qc022𝒞g(x_{0})=\frac{1}{2}\mathcal{A}-\frac{q_{c_{0}}^{2}}{2}\mathcal{C}. If |qc0|=θ^|q_{c_{0}}|=\hat{\theta}, then g(x0)=s^g(x_{0})=\hat{s}. If |qc0|<θ^|q_{c_{0}}|<\hat{\theta}, then g(x0)>s^g(x_{0})>\hat{s}.

Theorem 2.16.

Consider the optimization problem (21). A portfolio xx^{\star} is a solution for (21) if and only if

x=1aW0[Σ1γqminΣ1(μ1rf)],x^{\star}=\frac{1}{aW_{0}}\Big{[}\Sigma^{-1}\gamma-q_{min}\Sigma^{-1}(\mu-\textbf{1}r_{f})\Big{]}, (39)

for some

qminargminθΘQ(θ),q_{min}\in\arg min_{\theta\in\Theta}Q(\theta), (40)

where Θ=[θ^,θ^]\Theta=[-\hat{\theta},\hat{\theta}] if θ^=𝒜2s^𝒞<\hat{\theta}=\sqrt{\frac{\mathcal{A}-2\hat{s}}{\mathcal{C}}}<\infty and Θ=(,+)\Theta=(-\infty,+\infty) if θ^=+\hat{\theta}=+\infty. Here s^\hat{s} is the CV-L of the mixing distribution ZZ.

Proof.

First we show that if x^\hat{x} is a solution for (21), then x^\hat{x} is given by (39). By Lemma 2.8, x^\hat{x} is a solution for the optimization problem (24) with some c=c^c=\hat{c}. By Lemma 2.13, x^\hat{x} takes the following form

x^=1aW0[Σ1γq^Σ1(μ1rf)],\hat{x}=\frac{1}{aW_{0}}\Big{[}\Sigma^{-1}\gamma-\hat{q}\Sigma^{-1}(\mu-\textbf{1}r_{f})\Big{]},

with q^=/𝒞(aW0/𝒞)c^\hat{q}=\mathcal{B}/\mathcal{C}-(aW_{0}/\mathcal{C})\hat{c}. Again by Lemma 2.13 we have (see (33))

g(x^)=𝒜2(q^)22𝒞.g(\hat{x})=\frac{\mathcal{A}}{2}-\frac{(\hat{q})^{2}}{2}\mathcal{C}.

Since x^\hat{x} is a solution for (21) we have G(x^)<G(\hat{x})<\infty and this implies g(x^)s^g(\hat{x})\geq\hat{s} if s^\hat{s} is finite and g(x^)>s^g(\hat{x})>\hat{s} if s^=\hat{s}=-\infty (note that g(x^)=g(\hat{x})=-\infty implies G(x^)=+G(\hat{x})=+\infty due to the assumption Z0Z\neq 0 in Remark 2.2 and G(x^)=eaW0x^T(μrf1)Z(g(x^))G(\hat{x})=e^{-aW_{0}\hat{x}^{T}(\mu-r_{f}\textbf{1})}\mathcal{L}_{Z}(g(\hat{x}))). The expression of g(x^)g(\hat{x}) above then implies q^Θ\hat{q}\in\Theta (note here that for the case θ^=+\hat{\theta}=+\infty, we can’t have q^2=+\hat{q}^{2}=+\infty as g(x^)g(\hat{x}) is finite as explained above).

Now we need to show q^argminθΘQ(θ)\hat{q}\in\arg min_{\theta\in\Theta}Q(\theta). From Lemma 2.14, we have G(x^)=eQ(q^)G(\hat{x})=e^{-\mathcal{B}}Q(\hat{q}). Take any θ0Θ\theta_{0}\in\Theta (including the case Θ=(,+)\Theta=(-\infty,+\infty)). Let c0c_{0} be such that θ0=qc0\theta_{0}=q_{c_{0}} (see Remark 2.15). Let x0x_{0} be the solution for (24) with cc replaced by c0c_{0}. By Lemma 2.13 we have g(x0)=𝒜2(qc0)22𝒞g(x_{0})=\frac{\mathcal{A}}{2}-\frac{(q_{c_{0}})^{2}}{2}\mathcal{C}. Since θ0=qc0Θ\theta_{0}=q_{c_{0}}\in\Theta, we have g(x0)s^g(x_{0})\geq\hat{s} if s^\hat{s} is finite and g(x0)>s^g(x_{0})>\hat{s} if s^=\hat{s}=-\infty. Therefore either x0Sax_{0}\in S_{a} or x0S¯ax_{0}\in\bar{S}_{a}. Then by Lemma 2.14 we have G(x0)=eQ(qc0)G(x_{0})=e^{-\mathcal{B}}Q(q_{c_{0}}). Since x^\hat{x} is the optimal portfolio it is the minimizing point for the function G(x)G(x) (see (18) for this). Therefore we have G(x^)G(x0)G(\hat{x})\leq G(x_{0}). This implies Q(q^)Q(qc0)=Q(θ0)Q(\hat{q})\leq Q(q_{c_{0}})=Q(\theta_{0}). Since θ0\theta_{0} is arbitrary, we conclude that q^argminθΘQ(θ)\hat{q}\in\arg min_{\theta\in\Theta}Q(\theta).

Next we show that any portfolio of the form (39) is an optimal portfolio for (21). Fix an arbitrary qmargminθΘQ(θ)q_{m}\in\arg min_{\theta\in\Theta}Q(\theta). Then qm[θ^,θ^]q_{m}\in[-\hat{\theta},\hat{\theta}] if θ^\hat{\theta} is finite and qm(,+)q_{m}\in(-\infty,+\infty) if θ^=+\hat{\theta}=+\infty. Let cmc_{m} be such that qm=qcmq_{m}=q_{c_{m}} and let xmx_{m} be the solution of (24) with cc replaced by cmc_{m}. By Lemma 2.13, we have

xm=1aW0[Σ1γqmΣ1(μ1rf)],x_{m}=\frac{1}{aW_{0}}\Big{[}\Sigma^{-1}\gamma-q_{m}\Sigma^{-1}(\mu-\textbf{1}r_{f})\Big{]},

and g(xm)=𝒜2qm22𝒞g(x_{m})=\frac{\mathcal{A}}{2}-\frac{q_{m}^{2}}{2}\mathcal{C}. The condition on qmq_{m} above implies g(xm)s^g(x_{m})\geq\hat{s} if s^\hat{s} is finite and g(xm)>g(x_{m})>-\infty if s^=\hat{s}=-\infty. Therefore either xmSax_{m}\in S_{a} or xmS¯ax_{m}\in\bar{S}_{a}. By Lemma 2.14 we have G(xm)=eQ(qm)G(x_{m})=e^{-\mathcal{B}}Q(q_{m}) which is a finite number. To show xmx_{m} is an optimal portfolio we need to show G(xm)G(x)G(x_{m})\leq G(x) for any xx that G(x)G(x) is finite (note that either G(x)=+G(x)=+\infty or it is finite). Fix an arbitrary x¯\bar{x} with G(x¯)<+G(\bar{x})<+\infty. Let c¯=x¯T(μrf1)\bar{c}=\bar{x}^{T}(\mu-r_{f}\textbf{1}). Let xc¯x_{\bar{c}} be the solution of (24) with cc replaced by c¯\bar{c}. Since G(x)<G(x)<\infty we either have xS¯ax\in\bar{S}_{a} or xSax\in S_{a}. This means that xc¯S¯ax_{\bar{c}}\in\bar{S}_{a}. By Lemma 2.13 we have g(xc¯)=𝒜2qc¯22𝒞g(x_{\bar{c}})=\frac{\mathcal{A}}{2}-\frac{q_{\bar{c}}^{2}}{2}\mathcal{C}, where qc¯q_{\bar{c}} is given by (33) with cc replaced by c¯\bar{c}. Therefore we have qc¯[θ^,θ^]q_{\bar{c}}\in[-\hat{\theta},\hat{\theta}] if θ^\hat{\theta} is finite and qc¯(,+)q_{\bar{c}}\in(-\infty,+\infty) if θ^=+\hat{\theta}=+\infty. By the definition of qmq_{m}, we have Q(qm)Q(qc¯)Q(q_{m})\leq Q(q_{\bar{c}}). Therefore we have G(xm)=eQ(qm)eQ(qc¯)=G(x¯)G(x_{m})=e^{-\mathcal{B}}Q(q_{m})\leq e^{-\mathcal{B}}Q(q_{\bar{c}})=G(\bar{x}). ∎

Proposition 2.17.

Consider the optimization problem (21). If xx^{\star} is a regular solution for (21) then

x=1aW0[Σ1γqminΣ1(μ1rf)],x^{\star}=\frac{1}{aW_{0}}\Big{[}\Sigma^{-1}\gamma-q_{min}\Sigma^{-1}(\mu-\textbf{1}r_{f})\Big{]}, (41)

for some

qminargminθ(θ^,θ^)Q(θ),q_{min}\in\arg min_{\theta\in(-\hat{\theta},\hat{\theta})}Q(\theta), (42)

where θ^=:𝒜2s^𝒞\hat{\theta}=:\sqrt{\frac{\mathcal{A}-2\hat{s}}{\mathcal{C}}} and s^\hat{s} is the CV-L of the mixing distribution ZZ.

Proof.

Let x^\hat{x} be a regular solution. By Lemma 2.8, x^\hat{x} is a solution for the optimization problem (24) with some c=c^c=\hat{c}. By Lemma 2.13, x^\hat{x} takes the following form

x^=1aW0[Σ1γq^Σ1(μ1rf)],\hat{x}=\frac{1}{aW_{0}}\Big{[}\Sigma^{-1}\gamma-\hat{q}\Sigma^{-1}(\mu-\textbf{1}r_{f})\Big{]},

with q^=/𝒞(aW0/𝒞)c^\hat{q}=\mathcal{B}/\mathcal{C}-(aW_{0}/\mathcal{C})\hat{c}. Again by Lemma 2.13 we have (see (33))

g(x^)=𝒜2(q^)22𝒞.g(\hat{x})=\frac{\mathcal{A}}{2}-\frac{(\hat{q})^{2}}{2}\mathcal{C}.

Since x^\hat{x} is regular we have g(x^)>s^g(\hat{x})>\hat{s}. From this we conclude q^(θ^,θ^)\hat{q}\in(-\hat{\theta},\hat{\theta}). From Lemma 2.14, we have G(x^)=eQ(q^)G(\hat{x})=e^{-\mathcal{B}}Q(\hat{q}). Note that q^=qc^\hat{q}=q_{\hat{c}}. Now we show that q^argminθ(θ^,θ^)Q(θ)\hat{q}\in\arg min_{\theta\in(-\hat{\theta},\hat{\theta})}Q(\theta). Take any θ0(θ^,θ^)\theta_{0}\in(-\hat{\theta},\hat{\theta}). Let c0c_{0} be such that θ0=qc0\theta_{0}=q_{c_{0}} (see Remark 2.15). Let x0x_{0} be the solution for (24) with cc replaced by c0c_{0}. By Lemma 2.13 we have g(x0)=𝒜2(qc0)22𝒞g(x_{0})=\frac{\mathcal{A}}{2}-\frac{(q_{c_{0}})^{2}}{2}\mathcal{C}. Since θ0=qc0(θ^,θ^)\theta_{0}=q_{c_{0}}\in(-\hat{\theta},\hat{\theta}), we have g(x0)>s^g(x_{0})>\hat{s}. Therefore x0Sax_{0}\in S_{a}. Then by Lemma 2.14 we have G(x0)=eQ(qc0)G(x_{0})=e^{-\mathcal{B}}Q(q_{c_{0}}). Since x^\hat{x} is the optimal portfolio it is the minimizing point for the function G(x)G(x) (see (18) for this). Therefore we have G(x^)G(x0)G(\hat{x})\leq G(x_{0}). This implies Q(q^)Q(qc0)=Q(θ0)Q(\hat{q})\leq Q(q_{c_{0}})=Q(\theta_{0}). Since θ0\theta_{0} is arbitrary, we conclude that q^argminθ(θ^,θ^)Q(θ)\hat{q}\in\arg min_{\theta\in(-\hat{\theta},\hat{\theta})}Q(\theta). ∎

Remark 2.18.

Let us look at the case of Example 2.4. From the analysis in this example the optimal solution for the problem (21) is x=0x^{\star}=0 and it is unique. Here we would like to check that this optimal portfolio x=0x^{\star}=0 can also be derived from (39). To see this, note that in this example γ=0\gamma=0. Therefore we have Q(θ)=e𝒞θZ(θ22𝒞)Q(\theta)=e^{\mathcal{C}\theta}\mathcal{L}_{Z}(-\frac{\theta^{2}}{2}\mathcal{C}) and qc=aW0𝒞cq_{c}=-\frac{aW_{0}}{\mathcal{C}}c. Observe that 0{xT(μ1rf):xn}0\in\{x^{T}(\mu-\textbf{1}r_{f}):x\in\mathbb{R}^{n}\}. Also for any θ0\theta\neq 0 we have Q(θ)=+Q(\theta)=+\infty as the CV-L of ZeN(0,1)Z\sim e^{N(0,1)} is s^=0\hat{s}=0. Therefore argminθΘQ(θ)\arg min_{\theta\in\Theta}Q(\theta) has only one element qmin=0q_{min}=0. Then (39) gives x¯=0\bar{x}^{\star}=0 as the only optimal solution. Observe that in fact in this example we have 𝒜=0\mathcal{A}=0 and therefore θ^=0\hat{\theta}=0. Thus q¯min=argminθ{0}Q(θ)=0\bar{q}_{min}=\arg min_{\theta\in\{0\}}Q(\theta)=0.

Remark 2.19.

We remark here that our closed form formula (39) expresses the optimal portfolio in terms of the critical value (see definition 2.10) of the mixing distribution ZZ and its Laplace transformation which is hidden in the function Q(θ)Q(\theta). This has some advantage in determining the optimal portfolio for some cases of models (1), see our Corollary 4.5 below for this.

3 Large financial markets

In the previous section we gave closed form solution for the optimal portfolio for an exponential utility maximizer in a market that contains one risk-free asset and finitely many risky assets with return vector that follow (1). Our Theorem 2.16 gives complete characterization of the optimal portfolio in such small markets.

The next natural question to ask is what happens if the consumer with exponential utility wants to increase her expected utility as much as possible by adding as many as necessary assets into her portfolio. We can best investigate this possibility by working in mathematical models with countably infinitely many assets.

In this section we consider a sequence of economies with increasing number of assets. In the nnth economy, there are nn risky assets and one riskless asset. The return vector of the risky assets in the nnth economy satisfies (1). A consumer with exponential utility maximizes her expected utility based on the n+1n+1 assets in each nnth economy. Our main concern in this section is to investigate if the optimal expected utility of the consumer converges to a limit as nn\to\infty and we would like to identify this limit as the optimizer in the market with infinitely many assets.

Such “stability” of optimal investment problems was proved in [7] for a wide range of models. The methods of [7], however, cannot deal with exponential utilities. So we need to apply somewhat different, new arguments.

Our main result in this section shows that the consumer can achieve the maximum possible (in a market where she can trade on countably infinite risky assets) expected utility by following the sequence of optimal trading strategies in each nnth economy, which are shown to converge to a limit (see our Lemma 3.6 below). We call this limit portfolio the “overall best optimal portfolio” in this paper.

An economy that allows to trade on countably infinite risky assets is called a large financial market in the literature. They serve well to describe e.g. bonds of various maturities. A first model of this type, the “Arbitrage Pricing Model” (APM) goes back to [26]. We consider a slight extension of that model in the present section. As the main result of this section, we will show that the exponential utiliy maximization problem in large financial market can be approximated by similar problems for finitely many assets (and the latter can be solved by the results of the previous sections).

Before we state and prove our main result of this section, we first specify the structure of our nnth economy for all nn. Return on the bank account is R0:=rfR_{0}:=r_{f} where rf0r_{f}\geq 0 is the risk-free interest rate. For simplicity we assume rf=0r_{f}=0 henceforth. For i=1i=1, R1:=γ1Z+μ1+β¯1Zε1R_{1}:=\gamma_{1}Z+\mu_{1}+\bar{\beta}_{1}\sqrt{Z}\varepsilon_{1} is the return on the “market portfolio”, which may be thought of as an investment into an index. For i2i\geq 2, let the return on risky asset ii be given by

Ri=γiZ+μi+βiZε1+β¯iZεi.\displaystyle R_{i}=\gamma_{i}Z+\mu_{i}+\beta_{i}\sqrt{Z}\varepsilon_{1}+\bar{\beta}_{i}\sqrt{Z}\varepsilon_{i}. (43)

Here the (εi)i1(\varepsilon_{i})_{i\geq 1} are assumed independent standard Gaussian, ZZ is a positive random variable, independent of the εi\varepsilon_{i}, βi\beta_{i}, i2i\geq 2, β¯i0,γi,μi\bar{\beta}_{i}\neq 0,\gamma_{i},\mu_{i}, i1i\geq 1 are constants. The classical APM corresponds to Z1Z\equiv 1. We refer to [26] for further discussions on that model.

We consider investment strategies in finite market segments. A strategy investing in the first nn assets is a sequence of numbers ϕ0,ϕ1,,ϕn\phi_{0},\phi_{1},\ldots,\phi_{n}. For simplicity, we assume 0 initial capital and also that every asset has price 11 at time 0. Self-financing imposes i=0nϕi=0\sum_{i=0}^{n}\phi_{i}=0 so a strategy is, in fact, described by ϕ1,,ϕn\phi_{1},\ldots,\phi_{n} which can be arbitrary real numbers. The return on the portfolio ϕ\phi is thus

V(ϕ)=i=1nϕiRi,V(\phi)=\sum_{i=1}^{n}\phi_{i}R_{i},

noting also that R0=0R_{0}=0 is assumed.

For utility maximization to be well-posed, one should assume a certain arbitrage-free property for the market. Notice that a probability QnPQ_{n}\sim P is a martingale measure for the first nn assets (that is, EQn[Ri]=0E_{Q_{n}}[R_{i}]=0 for all 1in1\leq i\leq n) provided that

EQn[ε1|Z=z]=b1(z):=γ1zβ¯1μ1zβ¯1,z(0,)E_{Q_{n}}[\varepsilon_{1}|Z=z]=b_{1}(z):=-\frac{\gamma_{1}\sqrt{z}}{\bar{\beta}_{1}}-\frac{\mu_{1}}{\sqrt{z}\bar{\beta}_{1}},\ z\in(0,\infty) (44)

and, for each i2i\geq 2,

EQn[εi|Z=z]=bi(z):=γizβ¯iμizβ¯iβib1(z)zβ¯i,z(0,).E_{Q_{n}}[\varepsilon_{i}|Z=z]=b_{i}(z):=-\frac{\gamma_{i}\sqrt{z}}{\bar{\beta}_{i}}-\frac{\mu_{i}}{\sqrt{z}\bar{\beta}_{i}}-\frac{\beta_{i}b_{1}(z)\sqrt{z}}{\bar{\beta}_{i}},\ z\in(0,\infty). (45)

Now notice that, in fact, the set of such V(ϕ)V(\phi) coincides with the set of

V(h):=i=1nhiZ(εibi(Z))V(h):=\sum_{i=1}^{n}h_{i}\sqrt{Z}(\varepsilon_{i}-b_{i}(Z))

where h1,,hnh_{1},\ldots,h_{n} are arbitrary real numbers. We denote by HnH_{n} the set of all nn-tuples (h1,,hn)(h_{1},\ldots,h_{n}). It is more convenient to use this “hh-parametrization” in the sequel.

Assumption 3.1.

There are 0<c<C0<c<C such that cZCc\leq Z\leq C.

Let us define di:=supz[c,C]|bi(z)|d_{i}:=\sup_{z\in[c,C]}|b_{i}(z)|, i1i\geq 1. The next assumption is similar in spirit to the no-arbitrage condition derived in [26], see also [25].

Assumption 3.2.

We stipulate i=1di2<.\sum_{i=1}^{\infty}d_{i}^{2}<\infty.

Fact. If XX is standard normal then E[eθXθ2/2]=1E[e^{-\theta X-\theta^{2}/2}]=1 and E[XeθXθ2/2]=θE[Xe^{-\theta X-\theta^{2}/2}]=\theta, for all θ\theta\in\mathbb{R}. Notice also that, for all p1p\geq 1,

E[epθXpθ2/2]=e(p2p)θ2/2.E[e^{-p\theta X-p\theta^{2}/2}]=e^{(p^{2}-p)\theta^{2}/2}. (46)

Let us now define

fn(z):=exp(i=1n[bi(z)εi+bi(z)2]).f_{n}(z):=\exp\left(-\sum_{i=1}^{n}[b_{i}(z)\varepsilon_{i}+b_{i}(z)^{2}]\right).

Clearly, E[fn(z)]=1E[f_{n}(z)]=1 and E[fn(z)εi]=bi(z)E[f_{n}(z)\varepsilon_{i}]=b_{i}(z) for i=1,,ni=1,\ldots,n. Then QnQ_{n} defined by dQn/dP:=fn(Z)dQ_{n}/dP:=f_{n}(Z) will be a martingale measure for the first nn assets. Indeed,

E[fn(Z)]=[c,C]E[fn(z)]Law(Z)(dz)=1E[f_{n}(Z)]=\int_{[c,C]}E[f_{n}(z)]\,\mathrm{Law}(Z)(dz)=1

and

E[fn(Z)εi|Z=z]=E[(εibi(z))ebi(z)εibi(z)2/2]=0, 1in.E[f_{n}(Z)\varepsilon_{i}|Z=z]=E[(\varepsilon_{i}-b_{i}(z))e^{-b_{i}(z)\varepsilon_{i}-b_{i}(z)^{2}/2}]=0,\ 1\leq i\leq n.

It follows from (46) and from Assumption 3.2 that supnE[(dQn/dP)2]<\sup_{n}E[(dQ_{n}/dP)^{2}]<\infty hence dQ/dP:=limndQn/dPdQ/dP:=\lim_{n\to\infty}dQ_{n}/dP exists almost surely and in L2L^{2}, and this is a martingale measure for all the assets, that is, EQ[Ri]=0E_{Q}[R_{i}]=0 for all i1i\geq 1. Note also that E[(dQ/dP)2]<E[(dQ/dP)^{2}]<\infty.

Using the previous sections, we may find hnHnh^{*}_{n}\in H_{n} such that

Un:=E[eV(hn)]=minhHnE[eV(h)].U_{n}:=E[e^{-V(h^{*}_{n})}]=\min_{h\in H_{n}}E[e^{-V(h)}].

If we wish to find (asymptotically) optimal strategies for this large financial market then we also need to verify that UnU:=infhn1HnE[eV(h)]U_{n}\to U:=\inf_{h\in\cup_{n\geq 1}H_{n}}E[e^{-V(h)}] as nn\to\infty.

Let us introduce

2:={(hi)i1,hi,i1,i=1hi2<}\ell_{2}:=\left\{(h_{i})_{i\geq 1},\,h_{i}\in\mathbb{R},\,i\geq 1,\,\sum_{i=1}^{\infty}h_{i}^{2}<\infty\right\}

which is a Hilbert space with the norm h2:=i=1hi2||h||_{\ell_{2}}:=\sqrt{\sum_{i=1}^{\infty}h_{i}^{2}}. We may and will identify each (h1,,hn)Hn(h_{1},\ldots,h_{n})\in H_{n} with (h1,h2,)2(h_{1},h_{2},\ldots)\in\ell_{2} for all n1n\geq 1. Also define d:=(d1,d2,)2d:=(d_{1},d_{2},\ldots)\in\ell_{2}.

Theorem 3.3.

Under Assumptions 3.1 and 3.2, one has UnUU_{n}\to U, nn\to\infty.

Proof.

It follows from Lemma 3.6 below that there is h¯2\bar{h}^{*}\in\ell_{2} such that U=E[eV(h¯)]U=E[e^{-V(\bar{h}^{*})}]. Define now h~n:=(h¯1,,h¯n)Hn\tilde{h}_{n}:=(\bar{h}^{*}_{1},\ldots,\bar{h}^{*}_{n})\in H_{n}. It is clear that UnUU_{n}\geq U and E[eV(h~n)]UnE[e^{-V(\tilde{h}_{n})}]\geq U_{n} for all n1n\geq 1. Hence it remains to establish E[eV(h~n)]UE[e^{-V(\tilde{h}_{n})}]\to U.

Noting that V(h~n)V(h)V(\tilde{h}_{n})\to V(h^{*}) almost surely, it suffices to show that supnE[e2V(h~n)]<\sup_{n\in\mathbb{N}}E[e^{-2V(\tilde{h}_{n})}]<\infty. This follows from

E[e2V(h~n)]e2Ch~n2d2E[e2Ch~n2|N|]e2Ch2d2E[e2Ch2|N|],\displaystyle E[e^{-2V(\tilde{h}_{n})}]\leq e^{2\sqrt{C}||\tilde{h}_{n}||_{2}||d||_{2}}E[e^{2\sqrt{C}||\tilde{h}_{n}||_{2}|N|}]\leq e^{2\sqrt{C}||h^{*}||_{2}||d||_{2}}E[e^{2\sqrt{C}||h^{*}||_{2}|N|}],

where NN is a standard normal random variable. ∎

Remark 3.4.

The main message of Theorem 3.3 is that the sequence of optimal expected utilities in the small markets defined above is a convergent sequence, the limit being a finite number. This means that after the consumer increases the number of assets in her/his portfolio to a certain level, a further increase of the number of assets will not bring significant expected utility increments. It is not trivial to have some estimations on the number of assets needed for the optimal expected utility to be sufficiently close to the overall best utility level. It would be interesting to see how fast this sequence converges to the overall best utility level UU. We leave this for further discussions.

Lemma 3.5.

There exists α>0\alpha>0, such that for all h2h\in\ell_{2} with h2=1,\|h\|_{\ell_{2}}=1, P(V(h)α)αP(V(h)\leq-\alpha)\geq\alpha holds.

Proof.

We follow closely the proof of Proposition 3.2 in [7], see also [6]. We argue by contradiction. Assume that for all n1n\geq 1, there is gn=(gn(1),gn(2),)n1Hng_{n}=(g_{n}(1),g_{n}(2),\ldots)\in\cup_{n\geq 1}H_{n} with gn2=1\|g_{n}\|_{\ell_{2}}=1 and P(V(gn)1/n)1/nP\left(V(g_{n})\leq-{1}/{n}\right)\leq{1}/{n}.
Clearly, V(gn)0V(g_{n})^{-}\to 0 in probability as nn\to\infty. We claim that EQ[V(gn)]0E_{Q}[V(g_{n})^{-}]\to 0. By the Cauchy-Schwarz inequality

EQ[V(gn)]\displaystyle E_{Q}[V(g_{n})^{-}] \displaystyle\leq dQ/dPL2(P)(E[(V(gn))2])1/2.\displaystyle\|dQ/dP\|_{L^{2}(P)}\left(E[(V(g_{n})^{-})^{2}]\right)^{1/2}.

However,

V(gn)|V(gn)|C[|N|+d2]V(g_{n})^{-}\leq|V(g_{n})|\leq\sqrt{C}[|N|+||d||_{2}] (47)

for some standard normal NN. This implies E[(V(gn))2]E[(V(g_{n})^{-})^{2}], nn\to\infty and hence our claim.

Since EQ[V(gn)]=0E_{Q}[V(g_{n})]=0 by the martingale measure property of QQ, we also get that EQ[V(gn)+]0E_{Q}[V(g_{n})^{+}]\to 0. It follows that EQ[|V(gn)|]0,E_{Q}[|V(g_{n})|]\to 0, hence V(gn)V(g_{n}) goes to zero QQ-a.s. (along a subsequence) and, as QQ is equivalent to PP, PP-a.s. Using that |V(gn)|2|V(g_{n})|^{2}, nn\in\mathbb{N} is uniformly PP-integrable by (47), we get E[V(gn)2]0.E[V(g_{n})^{2}]\to 0. An auxiliary calculation gives

E[V(gn)2]=gn22E[Z]+i=1gn2(i)E[bi2(Z)Z]E[Z]>0,E[V(g_{n})^{2}]=\|g_{n}\|_{\ell_{2}}^{2}E[Z]+\sum_{i=1}^{\infty}g^{2}_{n}(i)E[b_{i}^{2}(Z)Z]\geq E[Z]>0,

a contradiction showing our lemma. ∎

Lemma 3.6.

There is h2h^{*}\in\ell_{2} such that U=E[eV(h)]U=E[e^{-V(h^{*})}].

Proof.

There are hnjHjh_{n}\in\cup_{j\in\mathbb{N}}H_{j}, nn\in\mathbb{N} such that E[eV(hn)]UE[e^{-V(h_{n})}]\to U. If we had supnhn2=\sup_{n}||h_{n}||_{\ell_{2}}=\infty then (taking a subsequence still denoted by nn), hn2||h_{n}||_{\ell_{2}}\to\infty, nn\to\infty. By Lemma 3.5,

P(V(hn)αhn2)αP(V(h_{n})\leq-\alpha||h_{n}||_{\ell_{2}})\geq\alpha

and this implies E[eV(hn)]E[e^{-V(h_{n})}]\to\infty, which contradicts E[eV(hn)]UE[e0]=1E[e^{-V(h_{n})}]\to U\leq E[e^{0}]=1.

Then necessarily supnhn2<\sup_{n}||h_{n}||_{\ell_{2}}<\infty and the Banach-Saks theorem implies that convex combinations h¯n\bar{h}_{n} of the hnh_{n} converge to some h2h^{*}\in\ell_{2} (in the norm of 2\ell_{2}). By Fatou’s lemma,

E[eV(h)]lim infnE[eV(h¯n)]lim infnE[eV(hn)]=U,E[e^{-V(h^{*})}]\leq\liminf_{n\to\infty}E[e^{-V(\bar{h}_{n})}]\leq\liminf_{n\to\infty}E[e^{-V({h}_{n})}]=U,

using also convexity of the exponential function. This proves the statement. ∎

4 Applications and examples

Our Theorem 2.16 gives closed form expression for the optimal portfolios for the problem (21) by using the function Q(θ)Q(\theta) defined in (35). In this section, we first study some properties of this function. Then we present some examples.

Let Z(s)=EesZ\mathcal{M}_{Z}(s)=Ee^{sZ} and 𝒦Z(s)=lnZ(s)\mathcal{K}_{Z}(s)=\ln\mathcal{M}_{Z}(s) denote the moment generating function (MGF) and the cumulant generating function (KGF) of the mixing distribution ZZ respectively. We have the following obvious relation

Q(θ)=e𝒞θZ(𝒞2θ2𝒜2),lnQ(θ)=𝒞θ+𝒦Z(𝒞2θ2𝒜2).Q(\theta)=e^{\mathcal{C}\theta}\mathcal{M}_{Z}(\frac{\mathcal{C}}{2}\theta^{2}-\frac{\mathcal{A}}{2}),\;\;\ln Q(\theta)=\mathcal{C}\theta+\mathcal{K}_{Z}(\frac{\mathcal{C}}{2}\theta^{2}-\frac{\mathcal{A}}{2}).

Therefore the minimizing points of Q(θ)Q(\theta) in (40) can also be found by using the MGF or KGF of ZZ. In the following Lemma we state some properties of the function Q(θ)Q(\theta).

Lemma 4.1.

Consider the model (1) with a non-trivial mixing distribution ZZ. Let s^\hat{s} denote the CV-L of ZZ and θ^\hat{\theta} is defined as in Section 2. Let the function Q(θ)Q(\theta) be defined by (35). Assume our model (1) is such that either 𝒜0\mathcal{A}\neq 0 or s^0\hat{s}\neq 0 which insures θ^=(𝒜2s^)/𝒞0\hat{\theta}=\sqrt{(\mathcal{A}-2\hat{s})/\mathcal{C}}\neq 0 and hence (θ^,θ^)(-\hat{\theta},\hat{\theta}) is a non-empty open interval. Then we have the following.

  • a)

    The function Q(θ)Q(\theta) is infinitely differentiable on (θ^,θ^)(-\hat{\theta},\hat{\theta}). If s^\hat{s} is finite and Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty or if s^=\hat{s}=-\infty, we have

    limθθ^Q(θ)=+,limθθ^+Q(θ)=+.\lim_{\theta\rightarrow\hat{\theta}^{-}}Q(\theta)=+\infty,\;\;\;\lim_{\theta\rightarrow-\hat{\theta}^{+}}Q(\theta)=+\infty. (48)

    When s^\hat{s} is finite and Z(s^)<\mathcal{L}_{Z}(\hat{s})<\infty we have Q(θ^)<Q(\hat{\theta})<\infty and Q(θ^)<Q(-\hat{\theta})<\infty. When s^\hat{s} is finite and θ[θ^,θ^]\theta\notin[-\hat{\theta},\hat{\theta}] we have Q(θ)=+Q(\theta)=+\infty.

  • b)

    The function Q(θ)Q(\theta) is strictly increasing on [0,θ^][0,\hat{\theta}] when s^\hat{s} is finite. It is strictly increasing on [0,+)[0,+\infty) when s^=\hat{s}=-\infty. We have Q(0)0Q^{\prime}(0)\neq 0 which implies the qminq_{min} in (39) can not be zero under the stated conditions in the Lemma.

  • c)

    The function Q(θ)Q(\theta) is strictly convex on the open interval (θ^,θ^)(-\hat{\theta},\hat{\theta}) when s^\hat{s} is finite and (s^)=+\mathcal{L}(\hat{s})=+\infty or when s^=\hat{s}=-\infty. Q(θ)Q(\theta) is strictly convex on [θ^,θ^][-\hat{\theta},\hat{\theta}] when s^\hat{s} is finite and (s^)<\mathcal{L}(\hat{s})<\infty.

Proof.

a) It is sufficient to prove that the function θZ(𝒜2𝒞2θ2)\theta\rightarrow\mathcal{L}_{Z}(\frac{\mathcal{A}}{2}-\frac{\mathcal{C}}{2}\theta^{2}) is infinitely differentiable when θ(θ^,θ^)\theta\in(-\hat{\theta},\hat{\theta}). This function is a composition of two functions sZ(s)s\rightarrow\mathcal{L}_{Z}(s) and θ𝒜2𝒞2θ2\theta\rightarrow\frac{\mathcal{A}}{2}-\frac{\mathcal{C}}{2}\theta^{2}. So it is sufficient to prove the infinite differentiability of sZ(s)s\rightarrow\mathcal{L}_{Z}(s) in the corresponding domain. If Z(s)\mathcal{L}_{Z}(s) is kk’th order differentiable then we would have Z(k)(s)=(s)kE[ZkesZ]\mathcal{L}_{Z}^{(k)}(s)=(-s)^{k}E[Z^{k}e^{-sZ}]. To justify the change of the order of derivative with expectation for this we need to show E[ZkesZ]<E[Z^{k}e^{-sZ}]<\infty. Let us look at the case s^0\hat{s}\neq 0 first. In this case we have EesZ<Ee^{sZ}<\infty in (,|s^|)(-\infty,|\hat{s}|). Thus all the moments of ZZ are finite. This implies E[ZkesZ]<E[Z^{k}e^{-sZ}]<\infty for any positive integer kk and all s(s^,+)s\in(\hat{s},+\infty). If θ(θ^,θ^)\theta\in(-\hat{\theta},\hat{\theta}), then 𝒜2𝒞2θ2(s^,𝒜2)\frac{\mathcal{A}}{2}-\frac{\mathcal{C}}{2}\theta^{2}\in(\hat{s},\frac{\mathcal{A}}{2}). Therefore when s^0\hat{s}\neq 0, the infinite differentiability of Q(θ)Q(\theta) follows. Now let us look at the case s^=0\hat{s}=0. In this case θ^=𝒜𝒞\hat{\theta}=\sqrt{\frac{\mathcal{A}}{\mathcal{C}}} and for any θ(θ^,θ^)\theta\in(-\hat{\theta},\hat{\theta}) we have 𝒜2𝒞2θ2(0,𝒜2)\frac{\mathcal{A}}{2}-\frac{\mathcal{C}}{2}\theta^{2}\in(0,\frac{\mathcal{A}}{2}). Therefore it is sufficient to prove infinite differentiability of Z(s)\mathcal{L}_{Z}(s) on (0,𝒜2)(0,\frac{\mathcal{A}}{2}). Fix an arbitrary positive integer kk. When s(0,𝒜2)s\in(0,\frac{\mathcal{A}}{2}) we have Zk/esZ=(Zk/esZ)1{ZM}+(Zk/esZ)1{Z>M}Z^{k}/e^{sZ}=(Z^{k}/e^{sZ})1_{\{Z\leq M\}}+(Z^{k}/e^{sZ})1_{\{Z>M\}} for any positive number MM. For sufficiently large M=M0M=M_{0}, we have (Zk/esZ)1{Z>M0}1(Z^{k}/e^{sZ})1_{\{Z>M_{0}\}}\leq 1 and Zk/esZ=(Zk/esZ)1{ZM0}Z^{k}/e^{sZ}=(Z^{k}/e^{sZ})1_{\{Z\leq M_{0}\}} is a bounded random variable. Thus E(ZkesZ)<E(Z^{k}e^{-sZ})<\infty for any positive integer kk when s(0,𝒜2)s\in(0,\frac{\mathcal{A}}{2}). This shows that θZ(𝒜2𝒞2θ2)\theta\rightarrow\mathcal{L}_{Z}(\frac{\mathcal{A}}{2}-\frac{\mathcal{C}}{2}\theta^{2}) is infinitely differentiable when s^=0\hat{s}=0 also.

When s^\hat{s} is finite and when θθ^\theta\rightarrow\hat{\theta} from the left-hand-side or when θθ^\theta\rightarrow-\hat{\theta} from the right-hand-side, the function 𝒜2𝒞2θ2\frac{\mathcal{A}}{2}-\frac{\mathcal{C}}{2}\theta^{2} decreasingly converges to s^\hat{s} (in some neighborhood of s^\hat{s}). Then the monotone convergence theorem gives the claim (48). Now assume s^=\hat{s}=-\infty which happens when the mixing distribution ZZ is a bounded non-trivial random variable. The result limθ+Q(θ)=+\lim_{\theta\rightarrow+\infty}Q(\theta)=+\infty is clear as both e𝒞θe^{\mathcal{C}\theta} and Z(𝒜2θ22𝒞)\mathcal{L}_{Z}(\frac{\mathcal{A}}{2}-\frac{\theta^{2}}{2}\mathcal{C}) go to ++\infty. The limit limθQ(θ)=+\lim_{\theta\rightarrow-\infty}Q(\theta)=+\infty is less clear as e𝒞θ0e^{\mathcal{C}\theta}\rightarrow 0 and Z(𝒜2θ22𝒞)+\mathcal{L}_{Z}(\frac{\mathcal{A}}{2}-\frac{\theta^{2}}{2}\mathcal{C})\rightarrow+\infty in this case. But since Z0Z\neq 0 with positive probability, we have a positive number δ>0\delta>0 with P(Zδ)>0P(Z\geq\delta)>0. We have the following

Q(θ)=Ee[𝒞2θ2𝒜2]Z+𝒞θe[𝒞2θ2𝒜2]δ+𝒞θP(Zδ),Q(\theta)=Ee^{[\frac{\mathcal{C}}{2}\theta^{2}-\frac{\mathcal{A}}{2}]Z+\mathcal{C}\theta}\geq e^{[\frac{\mathcal{C}}{2}\theta^{2}-\frac{\mathcal{A}}{2}]\delta+\mathcal{C}\theta}P(Z\geq\delta), (49)

for all θ\theta with 𝒞2θ2𝒜2>0\frac{\mathcal{C}}{2}\theta^{2}-\frac{\mathcal{A}}{2}>0. Then, since the right-hand-side of (49) goes to ++\infty when θ\theta\rightarrow-\infty, the claim follows. The remaining property of QQ in part a) above is obvious by the definition of θ^\hat{\theta}.

b) For any θ(θ^,θ^)\theta\in(-\hat{\theta},\hat{\theta}) we have

Q(θ)=𝒞e𝒞θZ[𝒜2θ22𝒞]θ𝒞e𝒞θZ[𝒜2θ22𝒞].Q^{\prime}(\theta)=\mathcal{C}e^{\mathcal{C}\theta}\mathcal{L}_{Z}[\frac{\mathcal{A}}{2}-\frac{\theta^{2}}{2}\mathcal{C}]-\theta\mathcal{C}e^{\mathcal{C}\theta}\mathcal{L}^{\prime}_{Z}[\frac{\mathcal{A}}{2}-\frac{\theta^{2}}{2}\mathcal{C}]. (50)

Observe that 0(θ^,θ^)0\in(-\hat{\theta},\hat{\theta}) always (in both cases s^0\hat{s}\neq 0 and s^=0\hat{s}=0). Therefore Q(0)Q^{\prime}(0) always exists and from (50) we see that Q(0)0Q^{\prime}(0)\neq 0. Now since Z(s)\mathcal{L}_{Z}(s) is a strictly decreasing function we have Z(s)<0\mathcal{L}^{\prime}_{Z}(s)<0. Therefore Q(θ)Q^{\prime}(\theta) is finite and Q(θ)>0Q^{\prime}(\theta)>0 when θ(0,θ^)\theta\in(0,\hat{\theta}). At θ=0\theta=0, we have Q(0)=𝒞Z(𝒜/2)Q(0)=\mathcal{C}\mathcal{L}_{Z}(\mathcal{A}/2) and clearly we have Q(0)<Q(θ)Q(0)<Q(\theta) for all θ(0,θ^)\theta\in(0,\hat{\theta}). At θ=θ^\theta=\hat{\theta}, we have Q(θ)=Z(s^)Q(\theta)=\mathcal{L}_{Z}(\hat{s}) which is either ++\infty or finite. When it is finite we have Q(θ)<Q(θ^)Q(\theta)<Q(\hat{\theta}) for all θ[0,θ^)\theta\in[0,\hat{\theta}) also.

c) Define fz(θ)=:e𝒞2zθ2+𝒞θ𝒜2zf_{z}(\theta)=:e^{\frac{\mathcal{C}}{2}z\theta^{2}+\mathcal{C}\theta-\frac{\mathcal{A}}{2}z} for any real number z0z\geq 0 and for all θ\theta\in\mathbb{R}. We have fz(θ)=(𝒞zθ+𝒞)e𝒞2zθ2+𝒞θ𝒜2zf^{\prime}_{z}(\theta)=(\mathcal{C}z\theta+\mathcal{C})e^{\frac{\mathcal{C}}{2}z\theta^{2}+\mathcal{C}\theta-\frac{\mathcal{A}}{2}z} and fz′′(θ)=𝒞ze𝒞2zθ2+𝒞θ𝒜2z+(𝒞zθ+𝒞)2e𝒞2zθ2+𝒞θ𝒜2z>0f^{{}^{\prime\prime}}_{z}(\theta)=\mathcal{C}ze^{\frac{\mathcal{C}}{2}z\theta^{2}+\mathcal{C}\theta-\frac{\mathcal{A}}{2}z}+(\mathcal{C}z\theta+\mathcal{C})^{2}e^{\frac{\mathcal{C}}{2}z\theta^{2}+\mathcal{C}\theta-\frac{\mathcal{A}}{2}z}>0 for any z0z\geq 0. Therefore fz(θ)f_{z}(\theta) is a strictly convex function for any fixed z0z\geq 0. Therefore we have

fz(λθ1+(1λ)θ2)<λfz(θ1)+(1λ)fz(θ2)f_{z}(\lambda\theta_{1}+(1-\lambda)\theta_{2})<\lambda f_{z}(\theta_{1})+(1-\lambda)f_{z}(\theta_{2})

for any λ[0,1]\lambda\in[0,1] and for all θ1,θ2\theta_{1},\theta_{2}\in\mathbb{R} for each fixed z0z\geq 0. This strict inequality also holds when z=Zz=Z. Also observe that when s^\hat{s} is finite and Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty or when s^=\hat{s}=-\infty, for θ1,θ2(θ^,θ^)\theta_{1},\theta_{2}\in(-\hat{\theta},\hat{\theta}) we have EfZ(θ1)<Ef_{Z}(\theta_{1})<\infty and EfZ(θ2)<Ef_{Z}(\theta_{2})<\infty. When s^\hat{s} is finite and Z(s^)<\mathcal{L}_{Z}(\hat{s})<\infty, for all θ,θ2[θ^,θ^]\theta_{,}\theta_{2}\in[-\hat{\theta},\hat{\theta}] we have EfZ(θ1)<Ef_{Z}(\theta_{1})<\infty and EfZ(θ2)<Ef_{Z}(\theta_{2})<\infty. We take expectation to the above inequality when z=Zz=Z and obtain Q(λθ1+(1λ)θ2)<λ1Q(θ1)+(1λ)Q(θ2)Q(\lambda\theta_{1}+(1-\lambda)\theta_{2})<\lambda_{1}Q(\theta_{1})+(1-\lambda)Q(\theta_{2}). This shows the strict convexity of Q(θ)Q(\theta) that is stated in the Lemma. ∎

Remark 4.2.

The main message of the above Lemma 4.1 is that the optimal solution for the problem (21) is always unique. Now assume Z(s^)<\mathcal{L}_{Z}(\hat{s})<\infty. In this case, if the optimal portfolio xx^{\star} for the problem (21) is irregular then the qminq_{min} in (39) satisfy qmin=θ^q_{min}=-\hat{\theta}. This means that θ^-\hat{\theta} is the minimizing point of Q(θ)Q(\theta) in [θ^,θ^][-\hat{\theta},\hat{\theta}]. As Q(θ)Q(\theta) is a strictly convex function on [θ^,θ^][-\hat{\theta},\hat{\theta}] as shown in the above Lemma 4.1, we conclude that Q(θ)Q(\theta) is a strictly increasing, strictly convex function on [θ^,θ^][-\hat{\theta},\hat{\theta}]. In comparision, when the solution for (21) is regular, then the corresponding Q(θ)Q(\theta) is strictly convex but not strictly increasing on [θ^,θ^][-\hat{\theta},\hat{\theta}].

Example 4.3.

Assume the mixing distribution ZZ in our model (1) takes finitely many values {zi}1im\{z_{i}\}_{1\leq i\leq m} with corresponding probabilities (pi)1im(p_{i})_{1\leq i\leq m}. Then XX in (1) is a mixture of Normal random vectors

Xi=1mpiNd(μ+γzi,ziΣ).X\sim\sum_{i=1}^{m}p_{i}N_{d}(\mu+\gamma z_{i},z_{i}\Sigma). (51)

In this case, the function Q(θ)Q(\theta) takes the following form

Q(θ)=i=1mpie(θ22C12A)zi+θC.Q(\theta)=\sum_{i=1}^{m}p_{i}e^{(\frac{\theta^{2}}{2}C-\frac{1}{2}A)z_{i}+\theta C}. (52)

From part c) of the above Lemma 4.1 we know that the function Q(θ)Q(\theta) is strictly convex on (,+)(-\infty,+\infty). Thus the solution for the optimization problem (21) is unique and this unique solution is given by (41) with qmin=argminθ(,0)Q(θ)q_{min}=\arg min_{\theta\in(-\infty,0)}Q(\theta). Now, assume Z=1Z=1 with probability one instead. Then Z(s)=es\mathcal{L}_{Z}(s)=e^{-s} and in this case it is easy to see that

Q(θ)=eC2(θ2+2θ)A2.Q(\theta)=e^{\frac{C}{2}(\theta^{2}+2\theta)-\frac{A}{2}}.

The minimizing point of this function is θ=1\theta=-1 and so qmin=1q_{min}=-1. Then, from (39), the optimal portfolio is given by

x=1aW0Σ1(γ+μ1rf).x^{\star}=\frac{1}{aW_{0}}\Sigma^{-1}(\gamma+\mu-\textbf{1}r_{f}).

Note here that since we assumed Z=1Z=1, the XX in (1) is a Gaussian random vector and therefore one can obtain the above optimal portfolio by direct calculation as our utility function is exponential. However, our above approach seems more convenient.

In the next example, we look at the case of GH models.

Example 4.4.

Lets look at the case of the model (1) when the mixing distribution ZZ is given by GIG models. First assume ZiG(λ,a22)Z\sim iG(\lambda,\frac{a^{2}}{2}), the inverse Gaussian distribution. In this case we have λ<0\lambda<0 by the definition of inverse Gaussian random variable. From Proposition 9 of [10] we have Z(s)=(2a2s)λ2Kλ(a2s)Γ(λ)\mathcal{L}_{Z}(s)=(\frac{2}{a\sqrt{2s}})^{\lambda}\frac{2K_{\lambda}(a\sqrt{2s})}{\Gamma(-\lambda)} and therefore Q(θ)=e𝒞θ(2a𝒜𝒞θ2)λ2Kλ(a𝒜𝒞θ2)Γ(λ)Q(\theta)=e^{\mathcal{C}\theta}(\frac{2}{a\sqrt{\mathcal{A}-\mathcal{C}\theta^{2}}})^{\lambda}\frac{2K_{\lambda}(a\sqrt{\mathcal{A}-\mathcal{C}\theta^{2}})}{\Gamma(-\lambda)}. In this case, the CV-L is s^=0\hat{s}=0 and θ^=𝒜/𝒞\hat{\theta}=\sqrt{\mathcal{A}/\mathcal{C}}. If γ=0\gamma=0, as discussed in the Example 2.4, the optimal solution for (21) is x=0x^{\star}=0. In this case, this solution x=0x^{\star}=0 is an irregular solution. Note that in this case 𝒜=0\mathcal{A}=0 and therefore θ^=0\hat{\theta}=0. If γ0\gamma\neq 0, then θ^>0\hat{\theta}>0 and in this case the qminq_{min} in (39) is given by qmin=argminθ[𝒜/𝒞,0)Q(θ)q_{min}=\arg min_{\theta\in[-\sqrt{\mathcal{A}/\mathcal{C}},0)}Q(\theta) (due to Lemma 4.1). Note that either by using the fact s^=0\hat{s}=0 or by using the property (A. 8) in [10] directly, one can easily check that (2a𝒜𝒞θ2)λ2Kλ(a𝒜𝒞θ2)Γ(λ)1(\frac{2}{a\sqrt{\mathcal{A}-\mathcal{C}\theta^{2}}})^{\lambda}\frac{2K_{\lambda}(a\sqrt{\mathcal{A}-\mathcal{C}\theta^{2}})}{\Gamma(-\lambda)}\rightarrow 1 when θ2𝒜/𝒞\theta^{2}\rightarrow\mathcal{A}/\mathcal{C}. Therefore Q(𝒜𝒞)=e𝒜𝒞Q(-\sqrt{\frac{\mathcal{A}}{\mathcal{C}}})=e^{-\sqrt{\mathcal{A}\mathcal{C}}}. In this case, it is not clear if qmin=𝒜𝒞q_{min}=-\sqrt{\frac{\mathcal{A}}{\mathcal{C}}} (the solution xx^{\star} is irregular) or qmin(𝒜𝒞,0)q_{min}\in(-\sqrt{\frac{\mathcal{A}}{\mathcal{C}}},0) (the solution xx^{\star} is regular).

Now let us look at the case ZGIG(λ,a,b)Z\sim GIG(\lambda,a,b) when a>0,b>0a>0,b>0. Again from Proposition 9 of [10] we have Z(s)=(bb2+2s)λKλ(ab2+2s)Kλ(ab)\mathcal{L}_{Z}(s)=(\frac{b}{\sqrt{b^{2}+2s}})^{\lambda}\frac{K_{\lambda}(a\sqrt{b^{2}+2s})}{K_{\lambda}(ab)} and Q(θ)=e𝒞θ(bb2+𝒜𝒞θ2)λKλ(ab2+𝒜𝒞θ2)Kλ(ab)Q(\theta)=e^{\mathcal{C}\theta}(\frac{b}{\sqrt{b^{2}+\mathcal{A}-\mathcal{C}\theta^{2}}})^{\lambda}\frac{K_{\lambda}(a\sqrt{b^{2}+\mathcal{A}-\mathcal{C}\theta^{2}})}{K_{\lambda}(ab)}. In this case s^=b2/2\hat{s}=-b^{2}/2 and θ^=𝒜+b2𝒞\hat{\theta}=\sqrt{\frac{\mathcal{A}+b^{2}}{\mathcal{C}}}. One can easily check Z(s^)=+\mathcal{L}_{Z}(\hat{s})=+\infty in this case. Therefore the unique optimal solution for (21) is given by (41) and it is regular.

Corollary 4.5.

Consider the model (1) with γ=0\gamma=0. In this case the distribution of XX is Elliptical distribution. Assume the CV-L of the mixing distribution ZZ is s^=0\hat{s}=0. Then the corresponding optimization problem (21) has a unique solution x=0x^{\star}=0. The CV-L of ZZ is s^=0\hat{s}=0 if EZn=+EZ^{n}=+\infty for some positive integer nn.

Proof.

Observe that in this case 𝒜=0\mathcal{A}=0 and therefore θ^=0\hat{\theta}=0. Then [θ^,θ^]={0}[-\hat{\theta},\hat{\theta}]=\{0\}. Therefore qminq_{min} in (39) is qmin=0q_{min}=0. As γ=0\gamma=0 also by assumption, we have x=0x^{\star}=0 by (39). It is clear that this solution is unique. If s^0\hat{s}\neq 0, then the Laplace transformation of ZZ is finite in (,|s^|)(-\infty,|\hat{s}|) and this would imply that all the moments of ZZ is finite. Therefore infinity of one of the moments of ZZ imply s^=0\hat{s}=0. ∎

Example 4.6.

(Stable distributions) Lets look at the case of α\alpha-stable distributions. Here we look at the 1- parametrization of the stable distributions (see Definition 1.5 of [24]). For other parametrizations see [24]. A distribution WW follows α\alpha-stable distribution with parameters α(0,2]\alpha\in(0,2], β[1,1]\beta\in[-1,1], σ>0\sigma>0, uu\in\mathbb{R} and we write WS(α,β,σ,u)W\sim S(\alpha,\beta,\sigma,u) if its characteristic function is given by

ϕ(t)=EeitW={eσα|t|α[1iβsign(t)tan(πα2)]+ituα1,eσ|t|[1+iβ2πsign(t)ln|t|]+ituα=1.\phi(t)=Ee^{itW}=\left\{\begin{array}[]{ll}e^{-\sigma^{\alpha}|t|^{\alpha}\left[1-i\beta sign(t)tan(\frac{\pi\alpha}{2})\right]+itu}&\mbox{$\alpha\neq 1$},\\ e^{-\sigma|t|\left[1+i\beta\frac{2}{\pi}sign(t)\ln|t|\right]+itu}&\mbox{$\alpha=1$}.\end{array}\right. (53)

When α=2\alpha=2, a stable distribution is a Normal distribution. When α(0,2)\alpha\in(0,2), EW2=+EW^{2}=+\infty for all β[1,1],σ>0,u\beta\in[-1,1],\sigma>0,u\in\mathbb{R}. Therefore for the mixing distributions Z=|W|,α(0,2),β[1,1],σ>0,uZ=|W|,\alpha\in(0,2),\beta\in[-1,1],\sigma>0,u\in\mathbb{R}, the corresponding CV-L is s^=0\hat{s}=0. Thus when γ=0\gamma=0 and when Z=|W|,α(0,2),β[1,1],σ>0,u,Z=|W|,\alpha\in(0,2),\beta\in[-1,1],\sigma>0,u\in\mathbb{R}, in the model (1), the optimization problem (21) has a unique solution x=0x^{\star}=0. This means that when the mixing distribution ZZ in (1) is equal to the absolute value of a stable distribution with α(0,2)\alpha\in(0,2) and when γ=0\gamma=0, then the optimal portfolio for an exponential utility maximizer is to invest all her/his wealth into the risk-free asset.

Remark 4.7.

Stable distributions are infinitely divisible. The characteristic functions (53) of the stable laws can be obtained directly from their Lévy-Khintchine representations. The generelized central limit theorem states that stable laws are the only non-trivial limits of normalized sums of independent identically distributed random variables. As such they were proposed to model many empirical (heavy tails, skewness etc.) financial phenomenons in the past. The heavy tailedness of them is related with the CV-L of them being s^=0\hat{s}=0. The above example 4.6 shows that time-changed Brownian motion models with stable subordinators (the ones with Elliptical marginal distributions) always give the trivial portfolio, investing everything on the risk-free asset, as the optimal portfolio for an exponential utility maximizer.

As pointed out in Remark 4.2, our Lemma 4.1 shows that the solution for the problem (21) is unique. Part b) of this Lemma shows that θ=0\theta=0 is not the minimizing point of the function Q(θ)Q(\theta) under the condition that 𝒜0\mathcal{A}\neq 0 or s^0\hat{s}\neq 0. For this unique minimizing point θ0\theta\neq 0 of Q(θ)Q(\theta) the first order condition (50) can equivalently be written as

Z(𝒜2𝒞2θ2)Z(𝒜2𝒞2θ2)=1θ.\frac{\mathcal{L}_{Z}^{\prime}(\frac{\mathcal{A}}{2}-\frac{\mathcal{C}}{2}\theta^{2})}{\mathcal{L}_{Z}(\frac{\mathcal{A}}{2}-\frac{\mathcal{C}}{2}\theta^{2})}=\frac{1}{\theta}. (54)

A change of variable η=𝒜/2(𝒞/2)θ2\eta=\mathcal{A}/2-(\mathcal{C}/2)\theta^{2}, which gives θ=(𝒜2β)/𝒞\theta=-\sqrt{(\mathcal{A}-2\beta)/\mathcal{C}} due to θ<0\theta<0 by Lemma 4.1, then gives

Z(β)Z(β)=𝒞/(𝒜2β),s^<β<𝒜/2.\frac{\mathcal{L}_{Z}^{\prime}(\beta)}{\mathcal{L}_{Z}(\beta)}=-\sqrt{\mathcal{C}/(\mathcal{A}-2\beta)},\;\;\hat{s}<\beta<\mathcal{A}/2. (55)

From this we can conclude that if xx^{\star} is a regular solution for (21), then βmin=:𝒜/2(𝒞/2)qmin2\beta_{min}=:\mathcal{A}/2-(\mathcal{C}/2)q_{min}^{2} with qminq_{min} in (41) satisfies the relation (55). This observation is useful if it can be confirmed that the solution for the equation (55) is unique. Then this unique solution equals to βmin\beta_{min}. Consider for example the case Z=1Z=1 in the model (1). As discussed in Example 4.3 above, in this case we have Z(s)=es\mathcal{L}_{Z}(s)=e^{-s}. Then Z(β)/Z(β)=1\mathcal{L}_{Z}^{\prime}(\beta)/\mathcal{L}_{Z}(\beta)=-1 and it is clear that the equation 1=𝒞/(𝒜2β)1=\sqrt{\mathcal{C}/(\mathcal{A}-2\beta)} has a unique solution β=𝒜/2𝒞/2\beta=\mathcal{A}/2-\mathcal{C}/2. This implies qmin2=1q_{min}^{2}=1 which then shows qmin=1q_{min}=-1 is the minimizing point of Q(θ)Q(\theta).

A positive random variable ZZ is a GGC with generating pair (τ,ν)(\tau,\nu) if

Z(s)=EesZ=eτ0ln(1+sz)ν(dz).\mathcal{L}_{Z}(s)=Ee^{-sZ}=e^{-\tau-\int_{0}^{\infty}\ln(1+\frac{s}{z})\nu(dz)}. (56)

If ZZ is a GGC with generating pair (τ,ν)(\tau,\nu), then Z(β)Z(β)=τ0+1tβν(dt)\frac{\mathcal{L}_{Z}^{\prime}(\beta)}{\mathcal{L}_{Z}(\beta)}=-\tau-\int_{0}^{+\infty}\frac{1}{t-\beta}\nu(dt). So if the solution for (21) is regular, then the βmin\beta_{min} defined above satisfy the following equation

τ|s^|+1tβν(dt)=𝒞/(𝒜2β),-\tau-\int_{|\hat{s}|}^{+\infty}\frac{1}{t-\beta}\nu(dt)=-\sqrt{\mathcal{C}/(\mathcal{A}-2\beta)},

where s^\hat{s} is the CV-L of the GGC random variable ZZ.

Now consider the case of positive α\alpha-stable random variables Z=S(α,1,σ,u),0<α<1,u>0Z=S(\alpha,1,\sigma,u),0<\alpha<1,u>0. Here we took β=1\beta=1 (see lemma 1.1 of [24]). After normalization these mixing distributions have the Laplace transformation Z(s)=esα\mathcal{L}_{Z}(s)=e^{-s^{\alpha}} (see Proposition 1 of [4] and also see [28]). Thus we have Z(s)/Z(s)=sαlns\mathcal{L}_{Z}^{\prime}(s)/\mathcal{L}_{Z}(s)=-s^{\alpha}\ln s. Assume the problem (21) has regular solution (a necessary condition for this is γ0\gamma\neq 0, see Corollary 4.5). Let βmin=𝒜/2(𝒞/2)qmin2\beta_{min}=\mathcal{A}/2-(\mathcal{C}/2)q_{min}^{2} with qminq_{min} in (41). Then 0<βmin<𝒜/20<\beta_{min}<\mathcal{A}/2 and it satisfies the following equation due to (55)

βαlnβ=𝒞/(𝒜2β).\beta^{\alpha}\ln\beta=\sqrt{\mathcal{C}/(\mathcal{A}-2\beta)}.

We square both sides of this equation and obtain

𝒜β2α(lnβ)22β2α+1(lnβ)2=𝒞.\mathcal{A}\beta^{2\alpha}(\ln\beta)^{2}-2\beta^{2\alpha+1}(\ln\beta)^{2}=\mathcal{C}.

As discussed earlier, if this equation has a unique solution β\beta then it is βmin\beta_{min}.

Remark 4.8.

We should mention here that the formula (39) for the optimal portfolio for the problem (21) is related with the Laplace transformation of the mixing distribution ZZ in the model (1) only. Namely we don’t need to know the probability density function of ZZ to find the optimal portfolio for the optimization problem (21). The relation (55) gives a convenient approach to locate the unique optimal portfolio as discussed earlier.

Next, we discuss the applications of our results in continuous time financial modelling. First we recall the Lemma 2.6 of [10] here. According to this Lemma, for each model F=Nd(μ+γz,zΣ)GF=N_{d}(\mu+\gamma z,z\Sigma)\circ G in (1) there is a corresponding Lévy process

Yt=μt+γτt+B¯τt,Y_{t}=\mu t+\gamma\tau_{t}+\bar{B}_{\tau_{t}}, (57)

with Law(Y1)=FLaw(Y_{1})=F and Law(τ1)=GLaw(\tau_{1})=G as long as G𝒥G\in\mathcal{J} (note that if G𝒥G\in\mathcal{J} then X𝒥X\in\mathcal{J} also from Lemma 2.5 of [10]). In the model (57), (B¯t)t0=(ABt)t0(\bar{B}_{t})_{t\geq 0}=(AB_{t})_{t\geq 0} where BtB_{t} is an nn-dimensional standard Brownian motion independent from (τt)t0(\tau_{t})_{t\geq 0} and (τt)t0(\tau_{t})_{t\geq 0} is a subordinator (a non-negative Lévy process with increasing sample paths). We denote the Lévy measure of this subordinator by ρ\rho and its Laplace transformation by

τt(s)=etΨ(s),\mathcal{L}_{\tau_{t}}(s)=e^{-t\Psi(s)}, (58)

where Ψ(s)=bs+0(1esy)ρ(dy)\Psi(s)=bs+\int_{0}^{\infty}(1-e^{-sy})\rho(dy) with a constant b0b\geq 0. As stated in Proposition 2.3 of [16], the function Ψ(s)\Psi(s) is continuous, nondecreasing, nonnegative, and convex. At each time point t>0t>0 we have

Yt=𝑑μt+γτt+τtANd.Y_{t}\overset{d}{=}\mu t+\gamma\tau_{t}+\sqrt{\tau_{t}}AN_{d}. (59)

Now consider a market with nn risky assets with price process StdS_{t}\in\mathbb{R}^{d} and one risk-free asset with price process Bt=etrfB_{t}=e^{tr_{f}}. Assume the log return process Yt=(Yt(1),Yt(2),,Yt(d))Y_{t}=(Y_{t}^{(1)},Y_{t}^{(2)},\cdots,Y_{t}^{(d)}), where Yt(i)=ln(St(i)/S0(i))Y_{t}^{(i)}=\ln(S_{t}^{(i)}/S_{0}^{(i)}) has the dynamics as in (57). The log return in the risk-free asset is ln(Bt/B0)=rft\ln(B_{t}/B_{0})=r_{f}t. An exponential utility maximizer wants to determine the optimal portfolio at each time point tt based on the log return vector of risky assets RdR\in\mathbb{R}^{d} with components R(i)=ln(St+(i)/St(i))R^{(i)}=\ln(S_{t+\triangle}^{(i)}/S_{t}^{(i)}) and the log-return of the risk-free asset R(0)=ln(Bt+/Bt)=rfR^{(0)}=\ln(B_{t+\triangle}/B_{t})=\triangle r_{f} in the time horizon [t,t+][t,t+\triangle]. Assume the time increment is =1\triangle=1. Then we have

R=𝑑μ+γτ1+τ1ANd,R\overset{d}{=}\mu+\gamma\tau_{1}+\sqrt{\tau_{1}}AN_{d}, (60)

and from our Theorem 2.16 the exponential utility maximizer’s optimal portfolio at time tt is

xt=1aW0(t)[Σ1γqmin(t)Σ1(μ1rf)],x_{t}^{\star}=\frac{1}{aW_{0}^{(t)}}\Big{[}\Sigma^{-1}\gamma-q_{min}^{(t)}\Sigma^{-1}(\mu-\textbf{1}r_{f})\Big{]}, (61)

where W0(t)W_{0}^{(t)} is his (initial) wealth that he invests on the n+1n+1 assets for the period [t,t+][t,t+\triangle] and qmin(t)q_{min}^{(t)} in (61) is given by qmin(t)=argminθΘQ(θ)q_{min}^{(t)}=argmin_{\theta\in\Theta}Q(\theta) in the corresponding domain θ\theta. Here

Q(θ)=eCθΨ(12Aθ22C),Q(\theta)=e^{C\theta-\Psi(\frac{1}{2}A-\frac{\theta^{2}}{2}C)}, (62)

due to (58).

Example 4.9.

(Variance-gamma model) Consider the financial market that was discussed in the paper [21]. The stock price is given by S(t)=S(0)emt+X(t;σS,νS,θS)+ωStS(t)=S(0)e^{mt+X(t;\;\sigma_{S},\;\nu_{S},\;\theta_{S})+\omega_{S}t} in their equation (21), where mm is the mean-rate of return on the stock under the statistical probability measure, ωS=1νSln(1θSνSσS2νS/2)\omega_{S}=\frac{1}{\nu_{S}}\ln(1-\theta_{S}\nu_{S}-\sigma_{S}^{2}\nu_{S}/2), and X(t;σS,νS,θS)=b(γ(t;1,νS);θS,σS)X(t;\sigma_{S},\nu_{S},\theta_{S})=b(\gamma(t;1,\nu_{S});\theta_{S},\sigma_{S}) with b(t;θ,σ)=θt+σW(t)b(t;\theta,\sigma)=\theta t+\sigma W(t) being a Brownian motion with drift θ\theta and volatility σ\sigma. Here the gamma process γ(t;μ,ν)\gamma(t;\mu,\nu) has mean rate μ\mu and variance rate ν\nu (note here that γ(t;μ,ν)G(μ2/ν,ν/μ)\gamma(t;\mu,\nu)\sim G(\mu^{2}/\nu,\nu/\mu) with our notation for gamma random variables in this paper). The increment g0=:γ(t+1;1,νS)γ(t;1,νS)=𝑑γ(1;1,νS)g_{0}=:\gamma(t+1;1,\nu_{S})-\gamma(t;1,\nu_{S})\overset{d}{=}\gamma(1;1,\nu_{S}) of this process has the Laplace transformation

g0(s)=(11+sνS)1νS,\mathcal{L}_{g_{0}}(s)=(\frac{1}{1+s\nu_{S}})^{\frac{1}{\nu_{S}}}, (63)

which can be seen also from the characteristic function expression in (3) of [21] for gamma processes. The risk-free asset in this financial market is given by Bt=B0etrfB_{t}=B_{0}e^{tr_{f}}. The log returns of these two assets in the time horizon [t,t+1][t,t+1] is given by

R=:ln(S(t+1)/S(t))=𝑑m+ωS+θSγ(1;1,ν)+σSγ(1;1,νS)N(0,1),R0=:ln(Bt+1/Bt)=rf.\begin{split}R=:&\ln(S(t+1)/S(t))\overset{d}{=}m+\omega_{S}+\theta_{S}\gamma(1;1,\nu)+\sigma_{S}\sqrt{\gamma(1;1,\nu_{S})}N(0,1),\\ R^{0}=:&\ln(B_{t+1}/B_{t})=r_{f}.\end{split}

An exponential utility maximizer with utility function u(x)=eax,a>0,u(x)=-e^{-ax},a>0, and wealth W0(t)W_{0}^{(t)} at time tt wants to decide on the optimal proportion xx^{\star} on the risky asset of his wealth for the period [t,t+1][t,t+1]. His acceptable set for xx^{\star} is given by

Sa={x:aW0(t)θSxa2(W0(t))22σS2x2>1νS},S_{a}=\{x\in\mathbb{R}:aW_{0}^{(t)}\theta_{S}x-\frac{a^{2}(W_{0}^{(t)})^{2}}{2}\sigma^{2}_{S}x^{2}>-\frac{1}{\nu_{S}}\}, (64)

as s^=1νS\hat{s}=-\frac{1}{\nu_{S}} in this case. The corresponding expressions for 𝒜,,𝒞\mathcal{A},\mathcal{B},\mathcal{C} in (32) are given by

𝒜=(θSσS)2,𝒞=(m+ωSrfσS)2,=θS(m+ωSrf)σS2.\mathcal{A}=(\frac{\theta_{S}}{\sigma_{S}})^{2},\mathcal{C}=(\frac{m+\omega_{S}-r_{f}}{\sigma_{S}})^{2},\mathcal{B}=\frac{\theta_{S}(m+\omega_{S}-r_{f})}{\sigma^{2}_{S}}.

Since the mixing distribution is a gamma random variable, the solution for the corresponding problem (21) is regular. Our Theorem 2.16 shows that the optimal portfolio is given by

x=1aW0[1σS2θSqmin1σS2(m+ωSrf)].x^{\star}=\frac{1}{aW_{0}}[\frac{1}{\sigma^{2}_{S}}\theta_{S}-q_{min}\frac{1}{\sigma^{2}_{S}}(m+\omega_{S}-r_{f})]. (65)

where qmin=argminθ(θ^,θ^)Q(θ)q_{min}=argmin_{\theta\in(-\hat{\theta},\hat{\theta})}Q(\theta) with Q(θ)Q(\theta) given by (35). Here θ^=𝒜+2/νS𝒞\hat{\theta}=\sqrt{\frac{\mathcal{A}+2/\nu_{S}}{\mathcal{C}}}. Next, we calculate qminq_{min} explicitly. We have Q(θ)=e𝒞θg0(𝒜/2(𝒞/2)θ2)Q(\theta)=e^{\mathcal{C}\theta}\mathcal{L}_{g_{0}}(\mathcal{A}/2-(\mathcal{C}/2)\theta^{2}) and from this we get lnQ(θ)=Cθ1vSln(1+A2vSC2vSθ2)\ln Q(\theta)=C\theta-\frac{1}{v_{S}}\ln(1+\frac{A}{2}v_{S}-\frac{C}{2}v_{S}\theta^{2}). The first order condition for the minimizing point of lnQ(θ)\ln Q(\theta) gives (θ+1𝒞νS)2=1+𝒞νS(2+𝒜νS)𝒞2νS2(\theta+\frac{1}{\mathcal{C}\nu_{S}})^{2}=\frac{1+\mathcal{C}\nu_{S}(2+\mathcal{A}\nu_{S})}{\mathcal{C}^{2}\nu_{S}^{2}}. This gives two solutions θ=1𝒞νS±1𝒞νS1+𝒞νS(2+𝒜νS)\theta=-\frac{1}{\mathcal{C}\nu_{S}}\pm\frac{1}{\mathcal{C}\nu_{S}}\sqrt{1+\mathcal{C}\nu_{S}(2+\mathcal{A}\nu_{S})}. But since θ\theta needs to be negative due to Lemma 4.1, we take qmin=θ=1𝒞νS1𝒞νS1+𝒞νS(2+𝒜νS)q_{min}=\theta=-\frac{1}{\mathcal{C}\nu_{S}}-\frac{1}{\mathcal{C}\nu_{S}}\sqrt{1+\mathcal{C}\nu_{S}(2+\mathcal{A}\nu_{S})}. We then plug this into (39) and obtain

x=1aW0(t)σS2[θS+m+ωSrf𝒞νS+m+ωSrf𝒞νS1+𝒞νS(2+𝒜νS)].x^{\star}=\frac{1}{aW_{0}^{(t)}\sigma^{2}_{S}}\Big{[}\theta_{S}+\frac{m+\omega_{S}-r_{f}}{\mathcal{C}\nu_{S}}+\frac{m+\omega_{S}-r_{f}}{\mathcal{C}\nu_{S}}\sqrt{1+\mathcal{C}\nu_{S}(2+\mathcal{A}\nu_{S})}\Big{]}. (66)

Therefore in this case we have closed form expression for the optimal portfolio. We should mention that one can use similar calculations to obtain closed form expression for optimal portfolio in a market where risky assets are modelled by multi-dimensional variance gamma (MVG) model, see [20] for the details of MVG models.

Remark 4.10.

Price processes with log-returns of the type (57) has been quite popular in financial literature in the past. Such models include inverse Gaussian Lévy processes, hyperbomic Lévy motions, variance gamma models, and CGYM models and all of these models were shown to fit empirical data quite well, see [5, 9, 27, 8, 19] and the references therein for this. In fact, every semimartingale can be written as a time change of Brownian motion, see [23] for this. This means that all the Lévy processes are time change of Brownian motion. In all these cases, if the time changing subordinator is independent from the Brownian motion then our Theorem 2.16 is applicable in principle. However, it is not easy to find the time-change used for general semimartinagles. Recently the paper [19] obtained the time change used for the CGMY model and Meixner processes. Our results in this paper can be applied to such processes to determine optimal portfolios for an exponential utility maximizer in a market where single or multiple risky asset dynamics follow such models.

5 Conclusion

The main result of this paper is Theorem 2.16 where we show that the problem of locating the optimal portfolio for (11) when the utility function is exponential boils down to finding the minimum point of a real valued function on the real-line, improving the Theorem 1 of [3] for the case of GH models and in the mean time extending it from the class of GH models to the general class of NMVM models. Our Theorem 3.3 shows that optimal exponential utility in small markets converge to the overall best exponential utility in the large financial market. While optimal portfolio problems under expected utility criteria for exponential utility functions have been discussed extensively in the past financial literature, an explicit solution of the optimal portfolio as in Theorem 2.16 above seems to be new. This is partly due to the condition we impose on the return vector XX of being a NMVM model. However, despite this restrictive condition on XX, asset price dynamics with NMVM distributions in their log returns often show up in financial literature like exponential variance gamma and exponential generalized hyperbolic Lévy motions.

References

  • Bingham and Kiesel [2001] N. H. Bingham and R. Kiesel. Modelling asset returns with hyperbolic distributions. In Return Distributions in Finance, pages 1–20. Elsevier, 2001.
  • Birge and Bedoya [2016] J. R. Birge and L. C. Bedoya. Portfolio optimization under a generalized hyperbolic skewed t distri- bution and exponential utility. Quantitative finance, 16:1019–1036, 2016.
  • Birge and Bedoya [2020] J. R. Birge and L. C. Bedoya. Portfolio optimization under generalized hyperbolic distribution: Optimal allocation, performance and tail behavior. Quantitative Finance, 21(2), 2020.
  • Bondesson [1988] L. Bondesson. A remarkable property of generalized gamma convolutions. Probability Theory and Related Fields, 78:321–333, 1988.
  • Car et al. [2002] P. Car, H. Geman, D. Madan, and M. Yor. The fine structure of asset returns: An empirical investigation. Journal of Business, 75:305–332, 2002.
  • Carassus and Rásonyi [2020] L. Carassus and M. Rásonyi. Risk-neutral pricing for arbitrage pricing theory. Journal of Optimization Theory and Applications, 186(1):248–263, 2020.
  • Carassus and Rásonyi [2021] L. Carassus and M. Rásonyi. From small markets to big markets. Banach Center Publications, 122(41–52), 2021.
  • Eberlein and Keller [1995] E. Eberlein and U. Keller. Hyperbolic distributions in finance. Bernoulli, pages 281–299, 1995.
  • Geman [2002] H. Geman. Pure jump lévy process for asset price modelling. Journal of Bank and Finance, 26(2):936–1297, 2002.
  • Hammerstein [2010] EAv. Hammerstein. Generalized hyperbolic distributions: theory and applications to CDO pricing. PhD thesis, Citeseer, 2010.
  • Hellmich and Kassberger [2011] M. Hellmich and S. Kassberger. Efficient and robust portfolio optimization in the multivariate generalized hyperbolic framework. Quantitative Finance, 11(10):1503–1516, 2011.
  • Hu and Kercheval [2010] W. Hu and A. N. Kercheval. Portfolio optimization for student t and skewed t returns. Quantitative Finance, 10(1):91–105, 2010.
  • James et al. [2008] L. F. James, B. Boynette, and M. Yor. Generalized gamma convolutions, dirichlet means, thorin measures, with explicit examples. Probability Survey, 5:346–415, 2008.
  • Karsten [1999] P. Karsten. The generalized hyperbolic model: Estimation, financial derivatives, and risk measures. PhD thesis, Universit¨at Freiburg i. Br., Freiburg i. Br, 1999.
  • Kwak and Pirvu [2018] M. Kwak and T.A. Pirvu. Cumulative prospect theory with generalized hyperbolic skewed t distribution. SIAM Journal on Financial Mathematics, (9):54–89, 2018.
  • Lalley [2007] S. P. Lalley. Lévy processes, stable processes, and subordinators. 2007.
  • Madan and McPhail [2000] D. Madan and G. McPhail. Investing in skews. Journal of Risk Finance,, 2:10–18, 2000.
  • Madan and Yen [2008] D. Madan and J. Yen. Asset Allocation with Multivariate Non-Gaussian Returns, volume Chapter 23 in Handbooks in OR MS, J.R. Birge and V. Linetsky. Elsevier, 2008.
  • Madan and Yor [2008] D. Madan and M. Yor. Representing the cgmy and meixner lévy processes as time changed brownian motions. Journal of computational finance, 12(1):27–47, 2008.
  • Madan and Seneta [1990] D. B. 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. B. Madan, P. Carr, and E. C. Chang. The variance gamma process and option pricing. Review of Finance, 2(1):79–105, 1998.
  • Mencía and Sentana [2005] J. Mencía and E. Sentana. Estimation and testing of dynamic models with generalized hyperbolic innovations. Econometrics eJournal, 2005.
  • Monroe [1978] I. Monroe. Process that can be embedded in brownian motion. Annals of Probability, 6:42–56, 1978.
  • Nolan [2020] J. P. Nolan. Univariate stable distributions. Models for heavy tailed data. 2020.
  • Rásonyi [2004] M. Rásonyi. Arbitrage pricing theory and risk-neutral measures. Decisions in Economics and Finance, 27(109–123), 2004.
  • Ross. [1976] S. A. Ross. The arbitrage theory of capital asset pricing. J. Econom. Theory, 13(341-360), 1976.
  • Schoutens [2003] W. Schoutens. Lévy processes in finance: pricing financial derivatives. Wiley Online Library, 2003.
  • Wissem and Thomas [2013] J. Wissem and S. Thomas. Further examples of ggc and hcm densities. Bernoulli, 19(5A):1818–1838, 2013.
  • Zakamouline and Koekabakker [2009] V. Zakamouline and S. Koekabakker. Portfolio performance evaluation with generalized sharpe ratios: Beyond the mean and variance. Journal of Banking and Finance, (33):1242–1254, 2009.