A discussion of stochastic dominance and mean-risk optimal portfolio problems based on mean-variance-mixture models
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 dimensional random vector has normal mean-variance mixture distribution with a mixing distribution if for any given the random vector satisfies , where are constant vectors, is a constant and positive-definite matrix, and is an dimensional normal random vector with mean and co-variance matrix . It is well known that such random vector satisfies
(1) |
where and is an dimensional normal random vector with mean equal to zero vector in and with co-variance matrix equal to the identity matrix in . The mixing distribution in (1) is independent from .
There are quite popular models within (1) in financial modeling. If follows a Generalized Inverse Gaussian (GIG) distribution, the distribution of 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 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 risky assets and we assume that the return vector of these risky assets satisfy (1). The portfolio set is given by and for each portfolio the corresponding portfolio return is given by , where denotes a transpose. Our main goal of this paper is to study the following problem
(2) |
for any given expected return level . In the optimization problem (2), is a risk measure and is the dimensional vector of ones. In our note we will provide closed form solution to the problem (2) for risk measures that satisfy certain conditions that will be discussed in Section 3 below. As one of the conditions on we requite that is law-invariant, i.e., as long as . Since
where is a standard normal random variable independent from , we have
We also require that 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 . We denote by the space of all finite valued random variables and by the space of random variables with finite moments. denotes the space of bounded random variables. For vectors and we denote by the scalar product of them and by the Euclidean norm of . As in [10], for each positive integer we denote by the space of all dimensional random vectors with all the components belong to . In this paper, unless otherwise stated we assume that the mixing distribution satisfies . We also assume that has probability density function . This latter assumption on is in fact is not necessary for our results of this paper. However, we impose this assumption to keep our calculations simple. We use to denote the density function of standard normal random variable and by 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 and . The value-at-risk (VaR) is defined for any random variable in . For any , denote by the left-continuous inverse of the cumulative distribution function of . For each the is defined as . The conditional value-at-risk (CVaR) is defined on as for any . In our note we are able to provide closed form solution for (2) when while an analytical solution for (2) when seems difficult. The main reason for this is the risk measure 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 space for any , 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 convergence for , see Proposition 4.4 in [11] for this. As mentioned earlier, the risk measure VaR is well defined on 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 and we consider risk measures for the problem (2). Denoting by the domain of , we call proper if for all and . A risk measure on 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., . A convex risk measure is called coherent if it is positive homogeneous, i.e., . From Corollary 2.3 of [17] any finite convex risk measure on is continuous in . The risk measure is a coherent risk measure on and hence it is continuous on . This property of is useful in the proof of the Proposition 2.14 below.
Another property of that is essential for our discussions in this section is its consistency with the second order stochastic dominance property. We say that is consistent with the second order stochastic dominance (SSD-consistent) property if for any two the relation implies that . This property of 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 is law-invariant if whenever . 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 is finite-valued, law-invariant, and convex risk measure. To see this, note first that the map is a continuous function as long as as is continuous map on as explained above. On the other hand, for any integrable random variable one has and hence , see Appendix B of [14] and Corollary 5.1 of [12] for similar arguments. Therefore on the set we have which implies
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
(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 denote the mean vector and let denote the co-variance matrix of . For each , the solution of (3) is given by
(4) |
where
Remark 2.2.
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 with mean vector and co-variance matrix we introduce the following expression
(5) |
where
Remark 2.3.
The right-hand-side of (5) only involves the mean vector and the co-variance matrix of the random vector . Therefore as long as and have the same mean vectors and co-variance matrices. Here we used the random variable in the definition of to indicate that the portfolio is the frontier portfolio under the mean-variance criteria with return vector . Note that the above expression (5) is well defined as long as and 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.
Proof.
Remark 2.5.
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 for return vectors as in (1), can be obtained by solving a Markowitz mean-variance optimal portfolio problem with an appropriately adjusted return vector as in the Theorem 2.4 above.
The risk measure for each fixed 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.
Remark 2.8.
We remark that our Theorem 2.4 above can be applied to simplify the calculations of the optimal values of 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 in (1), is a Normal random vector. We denote this Normal random vector by (taking ) for notations simplicity in the following discussions. Observe that with all the other model parameters and are fixed, the expression (5) becomes a function of the expected return level . Therefore for convenience we denote for in expression (5). We write this in a different form as
where
With these notations we have , where
(7) |
From [27] (see also equation (2) of [19]), for a Normal random variable we have
(8) |
where is the quantile of the standard Normal random variable . For any return level define
where is given by (7). Now, consider the following domain
for some fixed real number . Define . Observe that the domain has the property that for any given , any portfolio with lies in , i.e., . On this domain, we clearly have
(9) |
where . 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 in (4) of [28] as it is done in Theorem 2 of the same paper.
Remark 2.9.
Since the risk measure is not SSD-consistent, a similar result as in Corollary 2.7 for is difficult to construct. The following relation is well known.
where is referred to as the expected shortfall at level . Clearly we have (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 . It is not clear if the frontier portfolios under are close (in the Euclidean norm) to the frontier portfolios under due to the relation .
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 we have
(10) |
Define
It is easy to check that
Therefore is an increasing function of and also of . We use these facts in the proof of the following Lemma.
Lemma 2.10.
Let be any random variable that satisfies for some finite numbers and . Denote and consider the NMVM models and , where are any real numbers and . Then we have if and only if and .
Proof.
If then follows from Theorem 1 of [26] and follows from Lemma 3.13 above. To show the other direction we need to show for any real number . From (10) we have
We have
The above inequality follows because the two terms
and
are positive numbers as the function is an increasing function in the argument as explained in the paragraph preceding to this Lemma. We need to show that the term
is a positive number also. By middle value Theorem we have
(11) |
where is between and when and also is between and when . Observe that without the terms and in (11) we would have
Now, define . Then is an increasing or decreasing linear function of depending on or . Therefore either on and on or on and on . If on and on , then and therefore . This means that the function is multiplied by a larger valued function when it is positive valued in compared to it is multiplied by a smaller valued function while it takes negative values in . This shows that we have
This in turn implies . On the other-hand if on and on then and therefore which implies
This shows . This completes the proof. ∎
Remark 2.11.
We should mention that the assumption on being bounded mixing random variable in the above Lemma 2.10 is needed as it guarantees that the functions in the proof of this Lemma are bounded functions. Since , 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 and when
we have if and only if .
Before we state our next result we first recall few definitions. Recall that a random variable second order stochastically dominates another random variable , i.e., , iff for all . The dual of SSD is the risk-seeking stochastic dominance (RSSD) relation which is defined as for all . Since we have it is clear that the relation is equivalent to the relation . The risk measure is consistent with RSSD for any in the sense that (See page 174 of [5] for this)
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 we have
Next we state the following result.
Proposition 2.14.
Fix any positive integer . Assume and consider the following two NMVM models and , where are any real numbers and . Then the following condition
(12) |
is sufficient for .
Proof.
We divide the proof into two cases. First assume . Define each positive integer . Then by the dominated convergence theorem we have . Therefore there exists a positive integer such that we have
for all . Define and . Observe that and in . From Lemma 2.10 we have for all . This implies
(13) |
for each fixed . Since for each fixed the risk measure is continuous on , by taking limits to both sides of (13) as we obtain
for each . Then follows from Lemma 2.13 above.
Now assume . Define as above and let . Since we have as . Define the following two NMVM models and . We clearly have . Therefore from Lemma 2.10 we have for all . Also observe that and in . Now by following the same arguments as in the case above we obtain . This completes the proof. ∎
Remark 2.15.
Our above Proposition 2.14 shows that the condition (12) is sufficient for . It is not clear if also implies (12) when the mixing distribution is in for some finite positive integer . In comparison, if then by Lemma 2.10, the condition implies (12). If for all then from our Lemma 3.13 in Section 3 below the condition implies and for any positively valued bounded Borel function .
Remark 2.16.
As an application of our results above we present the following example.
Example 2.17.
Let be the return vector of risky assets and assume that
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], has the following NMVM representation
where . Consider the problem (6) for for any fixed real number and for any fixed . Our Theorem 2.4 shows that this problem has closed form solution and it is given by as in (5) . Note that the mixing distribution is and we assume here that the parameters are such that for some positive integer , 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.
where is any subset of the set of probability measures on , 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 , the times differentiable utility functions with alternating signs of derivatives for each , we say that a scalar random variable order stochastically dominates another scaler random variable if for all . We use the notation 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 denote by the right-continuous distribution function of it. For each integer define the following functions
recursively. We call the kth order cumulative distribution function of and we use the short-hand notation to denote it from now on. For any two random variables and , the relation can equivalently be defined by the following condition
(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 . 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 for all positive integers . 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 (here denotes the set of positive integers from now on) and for any random variable define for all , where denotes the positive part of . For any positive integer and any two , the paper [26] in their Proposition 1 shows that the kSD is equivalent to
(15) |
According to the Proposition 6 of [26], the function is an increasing convex function with and for all . It was also shown in the same Proposition that is a right asymptotic line of the function , i.e.,
For the remainder of this section, we calculate for some classes of random variables explicitly for any . Clearly these results are related to the kSD property through (14).
Normal random variables: When , we denote by . We have
(16) |
where is the standard normal random variable. We show the following simple Lemma.
Lemma 3.1.
For any , we have
(17) |
where and .
Proof.
We use induction. When , we have
where we have used which follows from L’Hopital’s rule. Assume (17) is true for and we show that it is also true for . To this end, we first integrate both sides of (17) and then apply (17). We obtain
(18) |
where we have used which follows from multiple applications of L’Hopital’s rule. From equation (18) we obtain
and this completes the proof. ∎
Remark 3.2.
Next, by using (17), we can obtain the following expression for
Lemma 3.3.
For any we have
(19) |
where is a ’th order polynomial that satisfies
and is a ’th order polynomial that satisfies
Proof.
Lemma 3.4.
Proof.
Remark 3.5.
Observe that and hence for all . We have
and
In the case that , we denote by . By using (16) , we can easily obtain
By letting , one can easily show that it satisfies the following equation
with the initial conditions
Elliptical random variables: When , where is a positive random variable independent from , we denote by . From (16), we have
Lemma 3.6.
For each , when we have
(21) |
Proof.
Remark 3.7.
NMVM random variables: When , we denote by (taking the first letter in hyperbolic distributions). From (16), we have
(24) |
By direct calculation and by using (17), we obtain the following result
Lemma 3.8.
For each positive integer , when we have
(25) |
Proof.
3.2 High-order SD
As stated earlier the condition (14), which needs to be checked for all the real numbers , 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 and , then if an only if and , 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 order SD relation is a weaker condition than the order SD relation for any as . Therefore it is not immediately clear that if also implies the relation for 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 is equivalent to and for each , 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 ) relations on elliptical random variables.
Before we proceed with our calculations, we first recall a relevant result from the paper [26]. With defined as in sub-section 3.1 above, the value of at is called central semideviation of and it is denoted by (again using the same notation in the paper [26]). That is we have
Proposition 4 of [26] shows that is a convex function, i.e., for any and any . Corollary 2 of the same paper shows that if then
(26) |
for all as long as . 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 and be two elliptical random variables with and is any positive random variable with for all positive integers . Then for each , if and only if and .
Proof.
The relation and implies for each follows from Proposition 3.6 above. To see the other direction, note that and and so the relation follows from Theorem 1 of [26]. To see the other relation , note that the central semi-deviations of and are and . Corollary 2 of [26] implies
(27) |
Now, since is an unbounded random variable we have . Dividing both sides of (27) by and letting we obtain . This completes the proof. ∎
Remark 3.11.
The significant part of this Proposition 3.10 is that the relation implies and for any fixed as long as for all , 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 denote any Borel function from to . Denote
For any real numbers define and . We first prove the following Lemma.
Lemma 3.12.
Consider the model (1) and assume that for all positive integers . Then for any bounded Borel function , we have
(28) |
Proof.
Observe that are bounded random variables. By using for any real numbers and the triangle inequality for norms, we have
From this it follows that
(29) |
Since are bounded random variables, we have and since are unbounded random variables we have . Therefore from (29) we conclude that
(30) |
Following the same idea, we have
From this we obtain
(31) |
Since are unbounded random variables we have and therefore as . Therefore from (31) we conclude that
(32) |
Next, for any bounded Borel function and any real numbers and , let and . The following Proposition is our main result in this Section.
Proposition 3.13.
Consider the model (1) and assume that for all positive integers . Then for any bounded Borel function and for each , the relation implies and .
Proof.
Since and , the relation follows from Theorem 1 of [26]. To show , we use Corollary 2 of the same paper [26]. First observe that for any integer the central semi-deviations of order are equal to
Denote and and observe that Lemma 3.12 implies . Also since and are unbounded random variables we have and . Corollary 2 of [26] implies
(33) |
for any . Now dividing both sides of (33) by and letting we obtain . This shows that . ∎
Now for any real numbers and any , let and . We have the following Corollary.
Corollary 3.14.
For any positive valued and bounded Borel function and for each , we have
Proof.
We remark here that in our calculations in this paper we have used the probability density functions of the mixing distributions . 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 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
where for all in the support of 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 . 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 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.