Exponential utility maximization in small/large financial markets
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 assets. We assume the first asset is a risk-free asset with risk-free interest rate and the remaining assets are risky assets with returns modelled by an dimensional random vector . In this note, we assume that follows normal mean-variance mixture (NMVM) distribution as follows
(1) |
where is location parameter, controls the skewness, is a non-negative random variable with distribution function , is a symmetric and positive definite matrix of real numbers, is a dimensional Gaussian random vector with identity co-variance matrix in , and is independent from the mixing distribution .
In this paper we use the following notations. For any vectors and in , where the superscript stands for the transpose of a vector, denotes the scalar product of the vectors and , and denotes the Euclidean norm of the vector . We sometimes use the short hand notation for (1), where . denotes the set of real numbers and denotes the set of non-negative real-numbers. Following the same notations of [13], denotes the family of infinitely divisible random variables on , denotes the set of self-decomposable random variables on , and denotes the class of generalized gamma convolutions (GGCs) on that will be introduced later. The Laplace transformation of any distribution is denoted by . A gamma random variable with density function is denoted by .
A prominent example of the NMVM models is generalized hyperbolic (GH) distributions, where the mixing distribution follows generalized inverse Gaussian (GIG) distribution denoted as . The probability density function of a GIG distribution, denoted by , takes the following form
(2) |
where denotes the modified Bessel function of third kind with index and the allowed parameter ranges for in (2) are (i) if , (ii) if , (iii) if . Here the case in (i) or the case in (ii) above need to be understood in limiting cases of (2) and in these special cases we have
(3) |
where denotes the Gamma function. Here is the density function of a Gamma distribution and is the density function of a inverse Gamma distribution .
The GH distribution in dimension is denoted by and it satisfies . The parameter ranges of this distribution is and (i’) if , (ii’) if , (iii’) if . The class of GH distributions include two popular models in finance: if we have normal inverse Gaussian distribution which is denoted by and when we have the class of hyperbolic distributions denoted by . As in the case of the GIG distributions, the case in (i’) above and the case or in (iii’) above needs to be understood as limiting cases of the GH distributions. If in case (i’) above then
(4) |
where denotes weak convergence of distributions and represents the class of variance gamma distributions. If and as well as in case (iii’) above we have the shifted distributions with degrees of freedom
(5) |
If and , we have the following that shows that the Normal random vectors are limiting cases of the GH distributions
(6) |
where is the dirac function that equals to when 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 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 is a GGC, without translation term, if there exists a positive Radon measure on such that
(7) |
with
(8) |
The measure is called Thorin’s measure associated with . 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
(9) |
where is a deterministic function with and is a standard Gamma process with Lévy measure .
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 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 . In our model (1) the mixing distribution can be any distribution in . In fact, can be any non-negative random variable.
Given an initial endowment , the investor must determine the portfolio weights x on the risky assets such that the expected utility of the next period wealth is maximized. The wealth that corresponds to portfolio weight on the risky assets is given by
(10) |
and the investor’s problem is
(11) |
for some domain of the portfolio set . Note here that represents the portfolio weights on the risky assets and is the proportion of the initial wealth invested on the risk free asset. The portfolio weights on risky assets are allowed to be any vector in .
The main goal of this paper is to discuss the solution of the problem (11) for exponential utility function 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 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
(12) |
and when the investment opportunity set consists of the above stated assets. Below we obtain an expression that relates to the Laplace transformation of the mixing distribution as in (14) below. First observe that we have
(13) |
Lemma 2.1.
For any portfolio such that is finite we have
(14) |
where is the Laplace transformation of .
Proof.
Remark 2.2.
If in our model (1), from (14) we have
Since is a strictly decreasing function, the expected utility maximization problem becomes the maximization problem of the quadratic function in this case. Especially, if the risk-free interest rate is zero and our model (1) is such that the location parameter 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 . Also we assume that 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 satisfies the following relation
(15) |
where 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 : (i) if the function 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 numerically. This problem is not trivial if the dimension is a large number, i.e., for large . To overcome these problems, in this paper we take different approach and obtain in near closed form: to calculate 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 and a quadratic function of the portfolio . For convenience, we introduce the following notations
(16) |
Then the relation (14) becomes
(17) |
Therefore we have the following obvious relation
(18) |
for any domain 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 and with the mixing distribution . Then for any we have
To see this, assume that there is a such that is finite. Then by Lemma 2.1 we have
For any we have as is positive definite by the assumption of the model (1). Now it is well known that when we have whenever . Therefore whenever and this contradicts with the finiteness assumption of made above. Thus we have whenever . Therefore the problem (11) does not have a solution when the domain does not include the zero vector in it. But if , then is the optimal portfolio and . This case corresponds to investing all the initial wealth on the risk-free asset as an optimal portfolio. We remark here that since by Jensen’s inequality we have
From this relation it is difficult to see that is the expected utility optimizing portfolio when . But with the assistance of Lemma 2.1 above it becomes trivial to determine that 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 , the zero portfolio is an optimal portfolio as when one has always. It is obvious that, in this case, the function is not differentiable at . Therefore we call 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 , if for all we set and if for some and for some , we let be the real number such that
(19) |
We call the critical value (CV) (we use the acronym CV-L from now on, where 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 under Laplace transformation in this paper. Observe that since is non-negative random variable we always have .
Remark 2.6.
In the above definition 2.5, the value of at is not specified. Both of the cases and are possible. For example if , then and clearly . If is a Gamma distribution, then . In this case and we have .
Below we define some domains for the portfolio set.
(20) |
Remark 2.7.
Our main objective in this section is to find closed form solution for the optimal portfolio for the problem
(21) |
The following relations are easy to see
(22) |
if and
(23) |
if . Observe here that if , then is a nonempty set as the zero vector is in it. If , then the set is nonempty as 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 at some constant level and optimize the quadratic term in (14). More specifically, we solve the following optimization problem
(24) |
first and plug in the solution, which we denote by , into the expression (14) so that the utility maximization problem becomes an optimization problem of a function of one variable .
Lemma 2.8.
Proof.
Define . Let be the solution to the problem (24) with replaced by (here the solution is unique as is positive definite by assumption). By the optimality of , we have . Since is a decreasing function we have . Since we have . This shows that . But is optimal for (11) with . Therefore we should have . This implies and this in turn implies again due to . The uniqueness of the optimization point for (24) then implies . ∎
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 . Now consider the case of example 2.4. In the setting of this example, consider the utility maximization problem (11). Since , as explained in the Example 2.4, the vector is the solution for the optimization problem (11). Now let be the optimal solution for the problem (24) with (which means ). Then we should have . But if , then can not be optimal solution for (11). Therefore we should have . The uniqueness of the optimal solution for (24) with then implies .
Definition 2.10.
Remark 2.11.
Remark 2.12.
Consider the optimization problem (11). From Lemma 2.8, any optimal portfolio is a solution for the quadratic optimization problem (24) with for some fixed . If is irregular, then . The optimality and uniqueness (on the hyperplane ) of implies that we have for all on the hyperplane . Therefore we have for all on the hyperplane . 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 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 is not a trivial issue. We first prove following Lemma.
Lemma 2.13.
For any real number , when , the maximizing point of is given by
(25) |
and we have
(26) |
where
(27) |
Proof.
We form the Lagrangian with the Lagrangian parameter . Denoting the maximizing point by , the first order condition gives
(28) |
We plug into and obtain
(29) |
From this we find as follows
For the rest of the paper, as in [3], for convenience, we use the following notations
(32) |
We first observe that due to the assumption in Remark 2.2 and the assumption on positive definiteness of . With these notations we have
(33) |
From the relation (33), we express as a function of as follows
(34) |
We define the following function
(35) |
and we define , where is the IN of . If , the is understood to be equal to . Note here that as is non-negative random variable. Therefore is well defined. If , is finite iff and this translates into: is finite iff . If , is finite iff and this translates into: is finite iff .
Next we prove the following Lemma that relates to .
Lemma 2.14.
Proof.
Note that . The stated conditions on in the Lemma insures that is finite. Since for any with by the definition of (the optimizing point) and also since is a decreasing function of we have
(37) |
for any with . We plug the in (34) into the expression of and obtain
(38) |
∎
Remark 2.15.
The above Lemma 2.14 shows that the function achieves its unique (as the solution for (24) is unique in a hyperplane) minimum value on the hyperplane at and its minimum value is given by with in (33). For any , we can let be such that . Let be the optimal solution of (24) with replaced by . From Lemma 2.13, we have . If , then . If , then .
Theorem 2.16.
Proof.
First we show that if is a solution for (21), then is given by (39). By Lemma 2.8, is a solution for the optimization problem (24) with some . By Lemma 2.13, takes the following form
with . Again by Lemma 2.13 we have (see (33))
Since is a solution for (21) we have and this implies if is finite and if (note that implies due to the assumption in Remark 2.2 and ). The expression of above then implies (note here that for the case , we can’t have as is finite as explained above).
Now we need to show . From Lemma 2.14, we have . Take any (including the case ). Let be such that (see Remark 2.15). Let be the solution for (24) with replaced by . By Lemma 2.13 we have . Since , we have if is finite and if . Therefore either or . Then by Lemma 2.14 we have . Since is the optimal portfolio it is the minimizing point for the function (see (18) for this). Therefore we have . This implies . Since is arbitrary, we conclude that .
Next we show that any portfolio of the form (39) is an optimal portfolio for (21). Fix an arbitrary . Then if is finite and if . Let be such that and let be the solution of (24) with replaced by . By Lemma 2.13, we have
and . The condition on above implies if is finite and if . Therefore either or . By Lemma 2.14 we have which is a finite number. To show is an optimal portfolio we need to show for any that is finite (note that either or it is finite). Fix an arbitrary with . Let . Let be the solution of (24) with replaced by . Since we either have or . This means that . By Lemma 2.13 we have , where is given by (33) with replaced by . Therefore we have if is finite and if . By the definition of , we have . Therefore we have . ∎
Proposition 2.17.
Proof.
Let be a regular solution. By Lemma 2.8, is a solution for the optimization problem (24) with some . By Lemma 2.13, takes the following form
with . Again by Lemma 2.13 we have (see (33))
Since is regular we have . From this we conclude . From Lemma 2.14, we have . Note that . Now we show that . Take any . Let be such that (see Remark 2.15). Let be the solution for (24) with replaced by . By Lemma 2.13 we have . Since , we have . Therefore . Then by Lemma 2.14 we have . Since is the optimal portfolio it is the minimizing point for the function (see (18) for this). Therefore we have . This implies . Since is arbitrary, we conclude that . ∎
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 and it is unique. Here we would like to check that this optimal portfolio can also be derived from (39). To see this, note that in this example . Therefore we have and . Observe that . Also for any we have as the CV-L of is . Therefore has only one element . Then (39) gives as the only optimal solution. Observe that in fact in this example we have and therefore . Thus .
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 and its Laplace transformation which is hidden in the function . 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 th economy, there are risky assets and one riskless asset. The return vector of the risky assets in the th economy satisfies (1). A consumer with exponential utility maximizes her expected utility based on the assets in each th economy. Our main concern in this section is to investigate if the optimal expected utility of the consumer converges to a limit as 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 th 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 th economy for all . Return on the bank account is where is the risk-free interest rate. For simplicity we assume henceforth. For , is the return on the “market portfolio”, which may be thought of as an investment into an index. For , let the return on risky asset be given by
(43) |
Here the are assumed independent standard Gaussian, is a positive random variable, independent of the , , , , are constants. The classical APM corresponds to . We refer to [26] for further discussions on that model.
We consider investment strategies in finite market segments. A strategy investing in the first assets is a sequence of numbers . For simplicity, we assume initial capital and also that every asset has price at time . Self-financing imposes so a strategy is, in fact, described by which can be arbitrary real numbers. The return on the portfolio is thus
noting also that is assumed.
For utility maximization to be well-posed, one should assume a certain arbitrage-free property for the market. Notice that a probability is a martingale measure for the first assets (that is, for all ) provided that
(44) |
and, for each ,
(45) |
Now notice that, in fact, the set of such coincides with the set of
where are arbitrary real numbers. We denote by the set of all -tuples . It is more convenient to use this “-parametrization” in the sequel.
Assumption 3.1.
There are such that .
Let us define , . The next assumption is similar in spirit to the no-arbitrage condition derived in [26], see also [25].
Assumption 3.2.
We stipulate
Fact. If is standard normal then and , for all . Notice also that, for all ,
(46) |
Let us now define
Clearly, and for . Then defined by will be a martingale measure for the first assets. Indeed,
and
It follows from (46) and from Assumption 3.2 that hence exists almost surely and in , and this is a martingale measure for all the assets, that is, for all . Note also that .
Using the previous sections, we may find such that
If we wish to find (asymptotically) optimal strategies for this large financial market then we also need to verify that as .
Let us introduce
which is a Hilbert space with the norm . We may and will identify each with for all . Also define .
Proof.
It follows from Lemma 3.6 below that there is such that . Define now . It is clear that and for all . Hence it remains to establish .
Noting that almost surely, it suffices to show that . This follows from
where 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 . We leave this for further discussions.
Lemma 3.5.
There exists , such that for all with holds.
Proof.
We follow closely the proof of Proposition 3.2 in [7], see also
[6].
We argue by contradiction.
Assume that for all , there is with and
.
Clearly, in probability as .
We claim that . By the Cauchy-Schwarz inequality
However,
(47) |
for some standard normal . This implies , and hence our claim.
Since by the martingale measure property of , we also get that . It follows that hence goes to zero -a.s. (along a subsequence) and, as is equivalent to , -a.s. Using that , is uniformly -integrable by (47), we get An auxiliary calculation gives
a contradiction showing our lemma. ∎
Lemma 3.6.
There is such that .
Proof.
There are , such that . If we had then (taking a subsequence still denoted by ), , . By Lemma 3.5,
and this implies , which contradicts .
Then necessarily and the Banach-Saks theorem implies that convex combinations of the converge to some (in the norm of ). By Fatou’s lemma,
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 defined in (35). In this section, we first study some properties of this function. Then we present some examples.
Let and denote the moment generating function (MGF) and the cumulant generating function (KGF) of the mixing distribution respectively. We have the following obvious relation
Therefore the minimizing points of in (40) can also be found by using the MGF or KGF of . In the following Lemma we state some properties of the function .
Lemma 4.1.
Consider the model (1) with a non-trivial mixing distribution . Let denote the CV-L of and is defined as in Section 2. Let the function be defined by (35). Assume our model (1) is such that either or which insures and hence is a non-empty open interval. Then we have the following.
-
a)
The function is infinitely differentiable on . If is finite and or if , we have
(48) When is finite and we have and . When is finite and we have .
-
b)
The function is strictly increasing on when is finite. It is strictly increasing on when . We have which implies the in (39) can not be zero under the stated conditions in the Lemma.
-
c)
The function is strictly convex on the open interval when is finite and or when . is strictly convex on when is finite and .
Proof.
a) It is sufficient to prove that the function is infinitely differentiable when . This function is a composition of two functions and . So it is sufficient to prove the infinite differentiability of in the corresponding domain. If is ’th order differentiable then we would have . To justify the change of the order of derivative with expectation for this we need to show . Let us look at the case first. In this case we have in . Thus all the moments of are finite. This implies for any positive integer and all . If , then . Therefore when , the infinite differentiability of follows. Now let us look at the case . In this case and for any we have . Therefore it is sufficient to prove infinite differentiability of on . Fix an arbitrary positive integer . When we have for any positive number . For sufficiently large , we have and is a bounded random variable. Thus for any positive integer when . This shows that is infinitely differentiable when also.
When is finite and when from the left-hand-side or when from the right-hand-side, the function decreasingly converges to (in some neighborhood of ). Then the monotone convergence theorem gives the claim (48). Now assume which happens when the mixing distribution is a bounded non-trivial random variable. The result is clear as both and go to . The limit is less clear as and in this case. But since with positive probability, we have a positive number with . We have the following
(49) |
for all with . Then, since the right-hand-side of (49) goes to when , the claim follows. The remaining property of in part a) above is obvious by the definition of .
b) For any we have
(50) |
Observe that always (in both cases and ). Therefore always exists and from (50) we see that . Now since is a strictly decreasing function we have . Therefore is finite and when . At , we have and clearly we have for all . At , we have which is either or finite. When it is finite we have for all also.
c) Define for any real number and for all . We have and for any . Therefore is a strictly convex function for any fixed . Therefore we have
for any and for all for each fixed . This strict inequality also holds when . Also observe that when is finite and or when , for we have and . When is finite and , for all we have and . We take expectation to the above inequality when and obtain . This shows the strict convexity of 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 . In this case, if the optimal portfolio for the problem (21) is irregular then the in (39) satisfy . This means that is the minimizing point of in . As is a strictly convex function on as shown in the above Lemma 4.1, we conclude that is a strictly increasing, strictly convex function on . In comparision, when the solution for (21) is regular, then the corresponding is strictly convex but not strictly increasing on .
Example 4.3.
Assume the mixing distribution in our model (1) takes finitely many values with corresponding probabilities . Then in (1) is a mixture of Normal random vectors
(51) |
In this case, the function takes the following form
(52) |
From part c) of the above Lemma 4.1 we know that the function is strictly convex on . Thus the solution for the optimization problem (21) is unique and this unique solution is given by (41) with . Now, assume with probability one instead. Then and in this case it is easy to see that
The minimizing point of this function is and so . Then, from (39), the optimal portfolio is given by
Note here that since we assumed , the 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 is given by GIG models. First assume , the inverse Gaussian distribution. In this case we have by the definition of inverse Gaussian random variable. From Proposition 9 of [10] we have and therefore . In this case, the CV-L is and . If , as discussed in the Example 2.4, the optimal solution for (21) is . In this case, this solution is an irregular solution. Note that in this case and therefore . If , then and in this case the in (39) is given by (due to Lemma 4.1). Note that either by using the fact or by using the property (A. 8) in [10] directly, one can easily check that when . Therefore . In this case, it is not clear if (the solution is irregular) or (the solution is regular).
Corollary 4.5.
Proof.
Observe that in this case and therefore . Then . Therefore in (39) is . As also by assumption, we have by (39). It is clear that this solution is unique. If , then the Laplace transformation of is finite in and this would imply that all the moments of is finite. Therefore infinity of one of the moments of imply . ∎
Example 4.6.
(Stable distributions) Lets look at the case of 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 follows stable distribution with parameters , , , and we write if its characteristic function is given by
(53) |
When , a stable distribution is a Normal distribution. When , for all . Therefore for the mixing distributions , the corresponding CV-L is . Thus when and when in the model (1), the optimization problem (21) has a unique solution . This means that when the mixing distribution in (1) is equal to the absolute value of a stable distribution with and when , 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 . 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 is not the minimizing point of the function under the condition that or . For this unique minimizing point of the first order condition (50) can equivalently be written as
(54) |
A change of variable , which gives due to by Lemma 4.1, then gives
(55) |
From this we can conclude that if is a regular solution for (21), then with 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 . Consider for example the case in the model (1). As discussed in Example 4.3 above, in this case we have . Then and it is clear that the equation has a unique solution . This implies which then shows is the minimizing point of .
A positive random variable is a GGC with generating pair if
(56) |
If is a GGC with generating pair , then . So if the solution for (21) is regular, then the defined above satisfy the following equation
where is the CV-L of the GGC random variable .
Now consider the case of positive -stable random variables . Here we took (see lemma 1.1 of [24]). After normalization these mixing distributions have the Laplace transformation (see Proposition 1 of [4] and also see [28]). Thus we have . Assume the problem (21) has regular solution (a necessary condition for this is , see Corollary 4.5). Let with in (41). Then and it satisfies the following equation due to (55)
We square both sides of this equation and obtain
As discussed earlier, if this equation has a unique solution then it is .
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 in the model (1) only. Namely we don’t need to know the probability density function of 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 in (1) there is a corresponding Lévy process
(57) |
with and as long as (note that if then also from Lemma 2.5 of [10]). In the model (57), where is an dimensional standard Brownian motion independent from and is a subordinator (a non-negative Lévy process with increasing sample paths). We denote the Lévy measure of this subordinator by and its Laplace transformation by
(58) |
where with a constant . As stated in Proposition 2.3 of [16], the function is continuous, nondecreasing, nonnegative, and convex. At each time point we have
(59) |
Now consider a market with risky assets with price process and one risk-free asset with price process . Assume the log return process , where has the dynamics as in (57). The log return in the risk-free asset is . An exponential utility maximizer wants to determine the optimal portfolio at each time point based on the log return vector of risky assets with components and the log-return of the risk-free asset in the time horizon . Assume the time increment is . Then we have
(60) |
and from our Theorem 2.16 the exponential utility maximizer’s optimal portfolio at time is
(61) |
where is his (initial) wealth that he invests on the assets for the period and in (61) is given by in the corresponding domain . Here
(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 in their equation (21), where is the mean-rate of return on the stock under the statistical probability measure, , and with being a Brownian motion with drift and volatility . Here the gamma process has mean rate and variance rate (note here that with our notation for gamma random variables in this paper). The increment of this process has the Laplace transformation
(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 . The log returns of these two assets in the time horizon is given by
An exponential utility maximizer with utility function and wealth at time wants to decide on the optimal proportion on the risky asset of his wealth for the period . His acceptable set for is given by
(64) |
as in this case. The corresponding expressions for in (32) are given by
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
(65) |
where with given by (35). Here . Next, we calculate explicitly. We have and from this we get . The first order condition for the minimizing point of gives . This gives two solutions . But since needs to be negative due to Lemma 4.1, we take . We then plug this into (39) and obtain
(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 of being a NMVM model. However, despite this restrictive condition on , 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.