Next Article in Journal
OpenCNN: A Winograd Minimal Filtering Algorithm Implementation in CUDA
Next Article in Special Issue
Mastering the Body and Tail Shape of a Distribution
Previous Article in Journal
Regular and Intra-Regular Semigroups in Terms of m-Polar Fuzzy Environment
Previous Article in Special Issue
A New Flexible Family of Continuous Distributions: The Additive Odd-G Family
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Conditions for the Existence of Absolutely Optimal Portfolios

by
Marius Rădulescu
1,*,
Constanta Zoie Rădulescu
2 and
Gheorghiță Zbăganu
1
1
“Gheorghe Mihoc-Caius Iacob” Institute of Mathematical Statistics and Applied Mathematics of the Romanian Academy, 050711 Bucharest, Romania
2
National Institute for Research and Development in Informatics, 011455 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Mathematics 2021, 9(17), 2032; https://doi.org/10.3390/math9172032
Submission received: 17 July 2021 / Revised: 17 August 2021 / Accepted: 22 August 2021 / Published: 24 August 2021
(This article belongs to the Special Issue Probability, Statistics and Their Applications 2021)

Abstract

:
Let Δn be the n-dimensional simplex, ξ = (ξ1, ξ2,…, ξn) be an n-dimensional random vector, and U be a set of utility functions. A vector x* Δn is a U -absolutely optimal portfolio if E u ξ T x * E u ξ T x for every x Δn and u U . In this paper, we investigate the following problem: For what random vectors, ξ, do U -absolutely optimal portfolios exist? If U 2 is the set of concave utility functions, we find necessary and sufficient conditions on the distribution of the random vector, ξ, in order that it admits a U 2 -absolutely optimal portfolio. The main result is the following: If x0 is a portfolio having all its entries positive, then x0 is an absolutely optimal portfolio if and only if all the conditional expectations of ξ i , given the return of portfolio x0, are the same. We prove that if ξ is bounded below then CARA-absolutely optimal portfolios are also U 2 -absolutely optimal portfolios. The classical case when the random vector ξ is normal is analyzed. We make a complete investigation of the simplest case of a bi-dimensional random vector ξ = (ξ1, ξ2). We give a complete characterization and we build two dimensional distributions that are absolutely continuous and admit U 2 -absolutely optimal portfolios.

1. Introduction. Defining the Problem

Let (Ω,K,P) be a fixed probability space and ξ: Ω → ℝn be an n–dimensional random vector. The random vector ξ = (ξ1, ξ2, …, ξn) will be called the vector of return rates (or the financial market). Let S be a positive number. A portfolio of sum S is a vector x + n with the property that x1 + … + xn = S. The set of all portfolios of sum S will be denoted by Δn(S). A standard portfolio is a portfolio of sum 1, that is, an element of standard simplex Δn(1).
We shall also use the set Pn−1(S) = {xn−1|xi ≥ 0 ∀ i = 1, …, n − 1 and x1 + … + xn−1S}. Instead of Pn−1(1), we shall write Pn−1. We shall denote by e the n-dimensional vector with all entries equal to one, that is e = 1 , 1 , , 1 . The return of portfolio x is the random variable
Y(x) = xTξ = x1ξ1 + … + xnξn
A utility function is a continuous non-decreasing function u: I, where I is an interval. Usually, the interval is either [0,∞) or the whole real line. A utility function of the form u(x) = erx, xI (with r > 0) or u(x) = − e−rx, xI (r > 0) is called a CARA-utility.
The set of CARA-utilities will be denoted with U 3 .
The main problem in the portfolio theory is to find “the best” portfolio. How do you compare two portfolios in order to decide which of them is better? In the framework of expected utility theory, any decision maker has a utility function U: I → ℝ and he tries to maximize the expected utility of a portfolio: that is, he tries to maximize
V x , ξ ; u = E u x T ξ
for all portfolios x ∈ Δn(S), provided that x T ξ I.
From the point of view of expected utility theory, a decision maker with utility u prefers the portfolio y to portfolio x if
V x , ξ ; u V y , ξ ; u E u x T ξ E u y T ξ
To avoid trivialities, we shall also suppose that the relation (1.2) makes sense; more precisely, we shall always assume that we deal with utilities having the property u(xTξ) ∈ L1(Ω, K,P) for any x + n .
A decision maker is called risk-avoiding if the utility is concave. The reason for this is that if u is concave then, by Jensen’s inequality, Eu(Y) ≤ u(EY) for any random variable Y. The meaning of this is that such a decision maker always prefers a sure amount of money to any random one having the same expected value.
We shall assume in the sequel that all the decision makers are risk-avoiding—after all, they will not be interested in portfolio theory otherwise.
Let U 2 (S, ξ) denote the set of all concave utilities, u having the property that u (xTξ) makes sense and, moreover, u (xTξ) ∈ L1(Ω, K,P) for all x ∈ Δn(S).
Let ξ be a fixed random vector from n and u U 2 (S, ξ) be a fixed utility. Consider the problem:
P(S): max{Eu(xTξ): x ∈ Δn(S)}
The meaning of the above optimization problem is that that any rational decision maker wants to maximize his expected utility. Notice that the mapping x Eu(xTξ), x ∈ Δn(S) is concave and continuous. The concavity is obvious and the continuity comes from Lebesgue’s domination principle: if xnx, the sequence of random variables (u(xnTξ))n converges to u(xTξ) and is dominated by the random variable
maxi(|u(S ξi)|) ∈ L1K,P)
As Δn(S) is a compact set and the mapping h is continuous, the problem P(S) always has solutions. Moreover, the set of solutions is a closed convex subset from Δn (S). This is called the set of optimal portfolios and is denoted by Opt(ξ,S;u). A portfolio x ∈ Δn (S) is called an equal weight portfolio if all its entries are equal, that is x = S n , .... , S n .
Our study was motivated by the following well known fact:
Theorem 1.
If ξi, I = 1, 2, …, n are independent and identically distributed (i.i.d.) random variables, ξ = (ξ1, ξ2,…, ξn) then the equal weight portfolio is optimal for the problem P(S):
x o = S n , .... , S n   Opt ( ξ , S ; u )   for   every   u     U 2 ( S ,   ξ )
In other words, any risk-avoiding decision maker agrees that the equal weight portfolio is optimal, provided that the random variables (ξi)1≤i≤n are i.i.d. Put another way, if the return of the assets has the same distribution and they are independent, then the investor’s choice to put the same amount of money in each asset is optimal from the point of view of the expected utility of the return.
In [1], Samuelson generalized the above result. He showed that the hypothesis “ξi, I = 1, 2, …, n are i.i.d. random variables” may be replaced by a more general condition:
Denote by Fξ the distribution of the random vector ξ, that is Fξ (B) = P(ξB). Call Fξ symmetrical if Fξoσ = Fξ for any permutation σ of the set {1, 2,…, n}. Here, ξ σ means the vector (ξσ(1), ξσ(2),…, ξσ(n)).
Samuelson [1] showed that if the distribution of ξ is symmetrical, then the equal weight portfolio is optimal for the problem P(S).
Hadar and Russell [2] gave an alternative proof of Samuelson’s result based on stochastic dominance. In [3], Tamir found a more general condition than that belonging to Samuelson. He showed that, if Fξoσ = Fξ for σ = (2, 3,…, n, 1), then the equal weight portfolio is optimal for the problem P(S). Thus, it is enough that F ξ 1 , ξ 2 , , ξ n = F ξ 2 , , ξ n , ξ 1 in order that the equal weight portfolio be optimal no matter the concave utility, u. In other words, the equal weight portfolio is absolutely optimal.
Another direction of research was initiated in Deng and Li [4]. They found necessary and sufficient conditions for the optimality of equal weight portfolios for the minimum variance problem.
An interesting problem is to find larger classes of matrices A with the following properties:
(i)
F ξ A x = F ξ x for every x n ;
(ii)
The equal weight portfolio is optimal for the problem P(S);
Tamir’s result [3] for cyclically symmetry of F ξ is the most general condition we know so far on F ξ .
Definition 1.
We state that xo is an absolutely optimal portfolio with sum S if xo ∈ Opt(ξ,S;u) for every utility u ∈ U 2 (S, ξ). We state that ξ (or better F ξ ) has the absolute optimal portfolio property (AOP) if an absolutely optimal portfolio does exist.
In the following, we shall write ‘simply absolute portfolio’ instead of ‘absolutely optimal portfolio’.
We investigate the following problem: For what random vectors, ξ, do U -absolute portfolios exist?
In the second section, we recall some facts about stochastic dominance relations and we prove some auxiliary results.
In Section 3, we solve the problem, if U = U 2 , what is the set of concave utility functions; we find necessary and sufficient conditions in order that F ξ has the AOP property, i.e., F ξ admits an absolute portfolio. The main result is the following: If x0 is a portfolio having all its entries positive, then x0 is an absolute portfolio if and only if all the conditional expectations of ξ i , given the return of portfolio x0, are the same.
In Section 4, we prove that, if ξ is bounded below, then CARA-absolute portfolios are also U 2 -absolute portfolios. In the case when ξi are rates of returns of assets in a financial market, then all ξi are bounded below, hence the above assertion holds.
In Section 5, the classical case when the random vector ξ is normal is analyzed.
We make a complete investigation of the simplest case of a bi-dimensional random vector ξ = (ξ1, ξ2) in Section 6. We give a complete characterization and we build two dimensional distributions that are absolutely continuous and admit U 2 -absolute portfolios.
The AOP property is connected with the stochastic ordering; actually, this means that the set of probability distributions of the random variables xTξ, xPn(S) has a maximum in the increasing concave order.

2. Stochastic Dominance

The increasing concave order, denoted by “icv”, is extensively used in economics, where it is called second-order stochastic dominance: see Belzunce [5], Levy [6], Sriboonchita [7], or Shaked and Shantikumar [8]. Firstly, let us recall the definition of the increasing concave order.
Definition 2.
Let X,Y be two integrable random variables. We state that X is dominated by Y in the increasing concave order if:
E u X   E u Y   for   every   concave   utility   u   such   that   u X , u Y L 1
We shall denote this domination relation by “Xicv Y”. If the inequality (4) holds for any concave function, u, not necessarily non-decreasing, we say that X is concavely dominated by Y and write “Xcv Y”. It is known that Xcv Y if, and only if, Xicv Y and EX = EY. Notice that the relation “≺icv” is an order relation between the probability distributions F and G of X and Y and can be better stated as:
F icv G u d F u d G   for   every   u U 2   such   that   u d F , u d G <
In this case, we can restate (3) as:
Proposition 1.
Let ξ = (ξ1, ξ2,…, ξn) be a random vector with i.i.d. integrable entries and xo = S n , .... , S n with S > 0. Then xTξ icv (xo)Tξ for every x ∈ Δn(S).
Proof. 
Let X = (xo)Tξ =  S ξ 1 + ξ 2 + + ξ n n . We confirm that E(ξi|X) = E(ξ1|X). Indeed, E(ξi|X) is a random variable of the form gi(X) having the property that E(ξih(X)) = E(gi(X)h(X)) for every bounded measurable function h:ℝ→ℝ. As ξi are i.i.d.,
E ( ξ i h ( X ) ) = x i h ( S n ( x 1 + + x n ) ) d F n x 1 , x 2 , , x n = x 1 h ( S n ( x 1 + + x n ) ) d F n x 1 , x 2   , ,   x n = E ( ξ 1 h ( X ) )
The last equality holds because of the equality dFn(xσ(1),…,xσ(n)) = dFn(x1,…,xn).
It follows that E(ξi|X) = E(ξ1|X) ∀ i = 2 ,…, n.
Consequently X = E(X|X) = E( S ξ 1 + ξ 2 + + ξ n n |X) = S n (E(ξ1|X) + E(ξ2|X) + … + E(ξn|X)) =  S n n⋅E(ξ1|X) = SE(ξ1|X). It follows that E(ξ1|X) = X/S.
Let x ∈ Δn(S) be arbitrary and v(x) = u(x/S). Then, by Jensen’s inequality, Eu(xTξ) = E(E(u(xTξ)|X)) = E E u ( S x 1 ξ 1 + + x n ξ n S X ) = E E v x 1 ξ 1 + + x n ξ n S X ) E v ( E x 1 ξ 1 + + x n ξ n S X ) = E(v(E(ξ1|X))) = E(v(X/S)) = Eu(X).
Therefore, Eu(xTξ) ≤ Eu(X) = Eu((xo)Tξ); hence, xo is indeed optimal. □
Remark. 
Taking into account the fact that ξi have the same expectation (recall that they are identically distributed) we could say even more: xTξ cv (xo)Tξ for every x ∈ Δn(S). Moreover, from the above proof, one can easily see that the assumption that the entries ξi were i.i.d. was not essential. What we really used was the fact that the distribution F⊗n is symmetrical (A random vector ξ = (ξ1, ξ2,…, ξn) is said to be symmetrically distributed if all its permutations (ξσ(1),…,ξσ(n)) have the same distribution as ξ). Thus, the result from Proposition 1 holds for every symmetrically distributed random vector.
It is important to keep in mind the following simple properties:
Proposition 2.
(i)
If X icv Y then EX ≤ EY;
(ii)
If X cv Y then Var(X) ≥ Var(Y);
(iii)
If f:ℝ →ℝ is non-decreasing and concave, then X icv Yf(X) icv f(Y);
(iv)
(Invariance with respect to convolution): If Xj icvYj, j = 1,2 and (X1, X2) are independent and (Y1, Y2) are independent, then X1 + X2 icv Y1 + Y2;
(v)
(Invariance with respect to mixture): If FjicvGj, j = 1,2 and 0 ≤ λ ≤ 1, then (1 − λ)F1 + λF2 icv (1 − λ)G1 + λG2;
(vi)
(Invariance with respect to weak convergence): If Fn icvGnn, Supp(Fn) ⊆ [0, ∞), ∀ n. Supp(Gn) ⊆ [0, ∞) ∀ n and FnF, GnG then F icv G.
Proof. 
(i)
Take the utility u(x) = x;
(ii)
Take the concave function f(x) = − x2;
(iii)
Notice that if u,f are non-decreasing and concave, then uf is also non-decreasing concave;
(iv)
See, for instance [5,6,7,8];
(v)
Obvious;
(vi)
We know that u d F n u d G n for any concave utility u:[0,∞) → ℝ and we want to check that u d F u d G . We claim that u d F n u d F , u d G n u d G . The reason for this is that:
u d F u d F n u d F min ( u , M ) d F + min ( u , M ) d F min ( u , M ) d F n = I 1 ( M ) + I 2 ( M , n )
Let ε > 0. Then there exists Mε > 0, such that I1(Mε) < ε/2. The function f = min(u,Mε) is continuous and bounded; hence, I2(Mε,n) → 0 as n . Thus, for n that is great enough, u d F u d F n < ε. Therefore, u d F n u d F , u d G n u d G .
Thus, our problem becomes:
Let ξ = (ξ1, ξ2,…, ξn) be a random vector, S > 0, Fx = P○(xTξ)−1 be the distribution of the random variable xTξ, and let D(ξ) = {P○(xTξ)−1|x ∈ Δn(S)}. When does the family D(ξ) have the greatest element with respect to the order “icv”?
Or, even better.
Let F be a probability distribution on n, Lx: n be the linear mappings Lx(y) = xTy, and Fx = Fo(Lx)−1 and D(F) = {Fx|x ∈ Δn(S)}. When does D(F) admit the greatest element with respect to the order “icv”?
Then Proposition 1 states that, when F = Gn with G a probability distribution on the real line, then D(F) admits the greatest element. Or, more generally, that if F is symmetric, it also has the greatest element. □
A portfolio xo ∈ Δn(S), with the property that F x o is the greatest element of D(F), will be called an absolutely optimal portfolio of sum S associated to F. We chose this name because all risk-avoiding decision makers will agree that this maximal element is the best among all portfolios of sum S.
We restate the Definition 1. in stochastic order terms:
Definition 3.
We state that the random vector ξ has the absolute optimum property (AOP) for sum S if the family D(ξ) has greatest element with respect to the order “icv”. If an absolute optimal portfolio corresponding to this greatest element is xo Δn(S), we will write that ξ has the property AO (xo) or that ξ ∈ AO (xo). In terms of distributions, we state that F ∈ AO (xo). Precisely:
ξ ∈ AO (xo) ⇔ xTξ icv (xo)Tξ for every x ∈ Δn(S)
Or, in terms of distributions:
F     AO   ( x o )     F L x 1 icv F L x o 1   for   every   x     Δ n ( S )
where Lx is the linear form Lx (y) = xT y, y ∈ ℝn.
Then, for a given S > 0, the class AO of distributions on ℝn is defined by:
A O = x o Δ n S A O x o

3. Properties of Class AO

At this point, we do not know if the AO class contains other distributions on ℝn beside the symmetric ones. However, the following properties of the class AO are immediate. We shall state them both in terms of random vectors and in terms of distributions.
Proposition 3.
Invariance Properties. Let S > 0 be fixed.
(i)
If ξAO (xo), a ∈ ℝ, λ > 0 then ae + λξAO (xo). Here e = (1, 1,…, 1) ∈ ℝn (invariance with respect to scaling);
(ii)
Ifξ1, ξ2are independent and both belong to AO (xo), then ξ1 + ξ2AO (xo) (invariance with respect to convolutions);
(iii)
Ifξ1, ξ2 are independent, AK is independent on both of them and ξ1, ξ2AO (xo), then the random vector ξ = ξ 1 1 A + ξ 2 1 A c also belongs to AO (xo) (invariance with respect to mixtures);
(iv)
If (ξn)n is a sequence of non-negative random vectors and ξnξ (in distribution), then ξnAO(xo) ∀ n implies ξAO(xo) (invariance with respect to weak convergence of nonnegative distributions);
(v)
Let ξAO(xo). Then xo/S is absolutely optimal among all the portfolios of sum 1, or the same thing in terms of distributions.
(i)
FAO(xo), a ∈ ℝ, λ > 0 ⇒ δa1  Fo(hλ)−1AO(xo); here, h λ is the homothety h λ x = λ x ;
(ii)
F1, F2 AO(xo) ⇒ F1F2AO(xo);
(iii)
F1, F2AO(xo), 0 ≤ λ ≤ 1 ⇒ (1 − λ)F1 + λF2AO(xo);
(iv)
Supp(Fn) ⊆ + n n, FnF, FnAO(xo) ∀ nFAO(xo);
(v)
Suppose thatxo ∈ Δn(S) and FAO(xo), then FAO(xo/S) as well. This means that when dealing with distributions of class AO we always can assume that S = 1.
Proof. 
(i)
The assumption is that xTξ icv (xo)Tξx ∈ Δn(S). Notice that xT(ae + λξ) = aS + λxTξ and (xo)T(ae + λξ) = aS + λ(xo)Tξ. Let X = xTξ and Xo = (xo)Tξ. We know that X icv Xo. Then aS + λX icv aS + λXo, because of Proposition 2., the function xaS + λx is increasingly concave.
(ii)
Now we know that xTξ1 icv (xo)Tξ1 ∀ x ∈ Δn(S) and that xTξ2 icv (xo)Tξ2. By Proposition 2. (iv), we see that xT(ξ1 + ξ2) icv (xo)T(ξ1 + ξ2).
(iii)
Let Fj = Po(ξj)−1 with j = 1,2 and 1 − λ = P(A). Then the distribution of ξ is (1 − λ)F1 + λF2. We assumed that FjAO(xo) ⇔ F j L x 1 icv F j L x o 1 ∀ x ∈ Δn(S), j = 1,2. Then, by Proposition 2. (v), we have 1 λ F 1 L x 1 + λ F 2 L x 1 icv 1 λ F 1 L x o 1 + λ F 2 L x o 1 or 1 λ F 1 + λ F 2 L x 1 icv 1 λ F 1 + λ F 2 L x o 1 for every x ∈ Δn(S), hence (1 − λ)F1 + λF2AO(xo).
(iv)
A consequence of Proposition 2 (v).
(v)
Obvious: xTξ icv (xo)Tξx ∈ Δn(S) ⇒ xTξ/S icv (xo)Tξ/Sx ∈ Δn(S) ⇔ yTξ icv (xo/S)Tξx ∈ Δn(1). □
Important Remark
Based on the above proposition, namely on point (v), it suffices to assume that S = 1. In the sequel we shall agree that always S = 1. Thus, Δn will mean Δn(1). Now we give a necessary and sufficient condition in order that a distribution on ℝn, F be in the class AO(xo), provided that x i 0 > 0 for all i = 1, …, n.
Proposition 4.
Let xo ∈ Δn be such that x i 0 > 0 ∀ i =1,…,n. Let also X 0 = (xo)Tξ. Then
ξ     A O   ( x o )     E ( ξ 1 | X 0 ) = E ( ξ 2 | X 0 ) = = E ( ξ n | X 0 ) = X 0
Proof. 
The easy part is “⇐”. Indeed, let x ∈ Δn and let u U 2 (1, ξ). Then
E u ( x T ξ ) = E [ E u ( x T ξ | X 0 ) ]     E [ u ( E ( x T ξ | X 0 ) ] = E [ u ( x 1 E ( ξ 1 | X 0 ) + x 2 E ( ξ 2 | X 0 ) + + x n E ( ξ n | X 0 ) ) ] = E u ( x 1 X 0 + x 2 X 0 + + x n X 0 ) = E u ( X 0 ) = E u ( ( x o ) T ξ ) ;
hence, xo is an absolute optimum. We prove now “⇒”. Let u  U 2 (1, ξ). Consider the function hu: Pn−1 → ℝ defined by:
hu(x1,…,xn−1) = Eu(x1ξ1 + … + xn−1ξn−1 + (1 − x1 − … − xn−1n)
Suppose that u is differentiable, then hu is also differentiable. Let g = Grad(hu). Thus,
g i x 1 , , x n 1 = h u x i x 1 , , x n 1 = E [ ξ 1 ξ n u x 1 ξ 1 + + x n 1 ξ n 1 + 1 x 1 x n 1 ξ n
Let yo = x 1 0 , x 2 0 , , x n 1 0 Pn−1. We know that hu attains its maximum at yo and yo is an interior point in Pn−1. Then:
g ( y o ) = 0     E [ ( ξ i ξ n ) u ( X 0 ) ] = 0     i = 1 , , n 1     u     U 2 ( 1 , ξ )
Recall that u is non-decreasing and concave; hence, u’ is decreasing, continuous, and non-negative. Conversely, any function g:ℝ → [0,∞) continuous and decreasing is the derivative of some concave utility. Taking a sequence of this kind of functions that monotonically converges to 1(−∞,a], we arrive at the conclusion that:
E [ ( ξ i ξ n ) ;   X 0     a ] = 0   for   every   i = 1 , , n 1   and   a    
Denote by B (ℝ) the σ-algebra of the Borel sets on the real line.
Let C = {A B (ℝ): E((ξi − ξn)1A( X 0 ) = 0}. According to (9), C contains all the intervals (−∞, a]. Moreover, it fulfills the following conditions:
(i)
ℝ ∈ C;
(ii)
ACAcC;
(iii)
AnC and (An)n are disjoint ⇒ n A n C.
Property (i) follows if we let a → ∞. Property (ii) results from the following series of equalities: 0 = E(ξi − ξn) − E((ξi − ξn)1A( X 0 ) = E((ξi − ξn)(1 − 1A)( X 0 ). Property (iii) follows from the Lebesgue’s domination principle.
Such a family of sets is called a Dynkin − system or a U-system. SoC is a U-system that contains the intervals (−∞, a]. It then contains the U-system generated by these intervals, but it is standard knowledge that this U-system is the Borel σ-algebra on the real line, B (ℝ). Thus, C contains all the Borelian sets. The conclusion is that:
E [ ( ξ i ξ n )   1 A ( X 0 ) ] = 0     I = 1   , ,   n 1     A   B   ( )
By a standard argument (if (10) holds for indicators, then it holds for simple functions and it holds for positive measurable functions). It follows that:
E [ ξ i   f ( X 0 ) ] = E [ ξ n   f ( X 0 ) ]     i = 1 , , n 1     f :       measurable   bounded
The conclusion is that E(ξi| X 0 ) = E(ξn| X 0 ). Indeed, E(ξi| X 0 ) is a random variable of the form φi( X 0 ) with the property that E(φi( X 0 )f( X 0 )) = E(ξif( X 0 )) for any f that is measurable and bounded. Hence, E(φi( X 0 )f( X 0 )) = E(φn( X 0 )f( X 0 )) for any f that is measurable and bounded; therefore, φi( X 0 ) = φn( X 0 ). We claim that, in this case, E(ξi| X 0 ) = X 0 . Indeed,
X 0 = E ( X 0 | X 0 ) = E ( x o 1 ξ 1 + + x o n ξ n | X 0 ) = x 1 0 E ( ξ 1 | X 0 ) + + x n 0 E ( ξ n | X 0 ) = ( x 1 0 + + x n 0 ) E ( ξ n | X 0 ) = E ( ξ n | X 0 )
Hence:
E ( ξ i | X 0 ) = E ( X n | X 0 ) = X 0
As a consequence, we find a necessary condition in order that ξ may have the property AO.
Corollary 1.
If ξAO (xo) and x i 0 > 0 ∀ I = 1,…,n, then Eξ1 = Eξ2 =…= Eξn.
Proof. 
According to Proposition 3.2, E(ξ1| X 0 ) = E(ξ2| X 0 ) =…= E(ξn| X 0 ). If we average the preceding sequence of equalities, we obtain the conclusion of the Corollary. □
What if ξAO (xo) but some of the components of xo are equal to 0? Is that possible?
In order to answer this question, let us suppose that the components of ξ are ordered in such a way that EX1 ≤ EX2 ≤…≤ EXn. It is possible that some of the expectations coincide. In general, we shall use the convention:
μ1 ≤ μ2 ≤…≤ μk < μk+1 = μk+2 =…= μn
Denote by INC(k), the set of all random vectors ξ = (ξ1, ξ2,…, ξn) with the above property.
For k = 0, we imply that all the expectations coincide and for k = n − 1 imply that μn−1 < μn. Then Corollary 1. states that, if ξ has an absolute optimal portfolio with all the components positive, then ξ ∈ INC(0).
Proposition 5.
Let ξ ∈ INC(k) with k ≥ 1.
If ξ has an absolute optimal portfolio xo, then x 1 0 =…= x k 0 = 0. Moreover, Var((xo)Tξ) ≤ Var(xTξ) ∀ x ∈ Δn.
In words: If an absolute optimal portfolio does exist, it should both maximize the expectation and minimize the variance.
Proof. 
Let X 0 = (xo)Tξ. Suppose, ad absurdum, that xoi > 0 for some ik. Then E X 0 < EXn; hence, X 0 cannot be optimal, according to Proposition 2. (i), which states that EXn ≤ E X 0 . Therefore, the possible absolute optima must be of the form xo =   0 , 0 , , x k + 1 0 , , x n 0 . All the portfolios x with xj = 0 ∀ jk have the property that E(xTξ) = μn = max1≤jn μj. The absolute optimal portfolio xo, provided that it does exist at all, should dominate them in the “icv” order. As it has the same expectation, μn, it should dominate them in the concave order “cv”; according to Proposition 2 (ii), Var((xo)Tξ) ≤ Var(xTξ) for all such x. If another portfolio, say y ∈ Δn, would exist such that Var((xo)Tξ) > Var(yTξ), then (xo)Tξ could not dominate yTξ in the “icv” order. □
Let us denote Pn−1(k) = {yPn−1: y1 =…= yk = 0, yj > 0 ∀ j > k }. Note that Pn−1(0) is the interior of Pn−1. Let Δn,k = {x ∈ Δn: x1 =…= xk = 0, xj > 0 ∀ j > k }. Note that Δn,0 = {x ∈ Δn: xj > 0 ∀ j}. We want to find necessary and sufficient conditions in order that ξAO (xo), xo ∈ Δn,k. We shall use the following particular case of Kuhn–Tucker conditions, which probably is well known:
Lemma 1.
Let f: Pn−1 → ℝ be concave and differentiable and let g = Grad(f). Let a Pn−1 (k). Then:
f(a) ≥ f(y) ∀ yPn−1gj(a) ≤ 0 ∀ jk, gj(a) = 0 ∀ j > k
Proof. 
“⇒”: Let a,yPn−1 and φa,y: [0,1] → ℝ be defined as φa,y(t) = f((1 − t)a + ty), t ∈ [0,1]. Then, φa,y is differentiable, concave, and its maximum is at t = 0. Therefore, it is non-increasing, hence (φa,y)’ (0) ≤ 0. As aPn−1(k), and we see that:
( φ a , y )   ( 0 ) = j = 1 n 1 y j a j g j a = j = 1 k y j g j a + j = k + 1 n 1 y j a j g j a     0   y     P n 1
If we choose yPn−1(k), the condition becomes j = k + 1 n 1 y j a j g j a ≤ 0, which of course implies gj(a) = 0 ∀ j > k (we can choose y = a ± εej with ε > 0 small enough). Thus, we get:
j = 1 k y j g j a     0     y P n 1     g j ( a )     0     j     k
Conversely, if gj(a) ≤ 0 ∀ jk, gj(a) = 0 ∀ j > k, it is obvious that (φa,y)’ (0) ≤ 0; as φa,y is concave, this fact implies that φa,y is non-increasing, hence φa,y (0) ≥ φa,y (1) ⇔ f(a) ≥ f(y). □
Proposition 6.
Let xo ∈ Δn,k and Xo = (xo)Tξ. Then ξAO (xo) ⇔ E(ξj/ X 0 ≤ a) ≤ E(ξn/ X 0 ≤ a) ∀ j ≤ k, a ∈ ℝ and:
E ( ξ j | X 0 ) = X 0     j   >   k
Proof. 
Let u be a concave differentiable utility, zPn−1(k), defined as:
z = 0 , 0 , , x k + 1 0 , , x n 1 0   and   h u ( y ) = E u ( y 1 ξ 1 + . . . + y n 1 ξ n 1 + ( 1 y 1   . . . y n 1 ) ξ n )
If ξAO (xo), then the concave differentiable function, hu, attains its maximum at z. Let g = Grad(hu). According to Lemma 1, gj(z) ≤ 0 ∀ jk, gj(z) = 0 ∀ j > k or:
E [ ( ξ j ξ n ) u ( X 0 ) ]     0     j     k , E [ ( ξ j ξ n ) u ( X 0 ) ] = 0     j   >   k     u     U 2   differentiable
In the same way as we did in the proof of Proposition 4., we can take u’ to be 1(−∞, a]. Thus, (14) becomes:
E [ ( ξ j ξ n ) 1 ( , a ] ( X 0 ) ]     0     j     k , E [ ( ξ j ξ n )   1 ( , a ]   ( X 0 ) ] = 0     j   >   k     a    
which implies (13), by the same argument used in the proof of Proposition 4.:
E ( ξ j | X 0 ) = E ( ξ n | X 0 )     j > k
which implies that:
X o = E ( X o X o ) = j = k + 1 n x j o E ( ξ j X o ) = j = k + 1 n x j o E ( ξ n X o ) = E ( ξ n X o )
Hence, E(ξj| X o ) = X o for every j > k.
Conversely, if the conditions from the right hand of (13) hold, then E(ξiu’(Xo)) ≤ E(ξnu’( X o )) ∀ jk and E(ξiu’( X 0 )) = E(ξnu’( X 0 )) ∀ j > k for any u concave that is differentiable and non-decreasing. By Lemma 1., the function hu attains its maximum at xo for any differentiable concave utility. Thus Eu((xo)Tξ) ≥ Eu((x)Tξ) for any x ∈ Δn, u differentiable concave that is non-decreasing. However, the restriction that u be differentiable is not a serious one: any utility is the uniform limit of differentiable ones.
Corollary 2.
Let xo ∈ Δn,k. A sufficient condition in order that ξ ∈ AO (xo) is
E [ ( ξ j ξ n ) | X 0 ]     0     j     k , E [ ( ξ j ξ n ) | X 0 ) ] = 0     j   >   k
Proof. 
Obviously, E[(ξj − ξn)| X 0 ] ≤ 0 implies E(ξiu’( X 0 )) ≤ E(ξnu’( X 0 )) ∀ u differentiable, concave, and non-decreasing, such that the integral makes sense. □

4. The Class CARAAO

In this section, we shall deal only with short tailed random vectors. A random vector ξ = (ξ1,...,ξn) is short tailed if the moment generating functions, mi(t) = E[exp(tξi)], are finite in a neighborhood of 0. If we consider only CARA utilities of the form u(x) = − e−rx, we can give the following.
Definition 4.
Let xo ∈ Δn and ξ be a random vector. We say that ξ has the property CARAAO(xo) (and write ξ ∈ CARAAO(xo)) if:
E e r ( x o ) T ξ E e r x T ξ
for any r > 0 such that both the expectations are finite. The union of all the classes CARAAO(xo) is defined to be the class CARAAO.
It is obvious that, if ξAO(xo), then ξCARAAO(xo)
The fact is that, in some cases, the classes CARAAO and AO coincide. Some conditions under which the equality of classes holds are presented in the following proposition.
Proposition 7.
Suppose that ξ is bounded below, in the sense that there exists m ∈ ℝ such that ξi ≥ m a.s. ∀ i = 1,...,n. Then, CARAAO(xo) = AO(xo) for every xo ∈ Δn,0. Or, in words: if all the components of xo are positive, then the classes CARAAO(xo) and AO(xo) coincide.
Proof. 
Let xo ∈ Δn,0 and yo = x 1 0 , x 2 0 , , x n 1 0 ∈ Pn−1. Let h: Pn−1 → ℝ defined by
h x = e r x 1 ξ 1 + + x n 1 ξ n 1 + 1 x 1 x n 1 ξ n ,   x     P n 1
The function h is concave and attains its maximum at yo, which is an interior point of the compact Pn−1. Let X 0 = (xo)Tξ. Then:
(Grad h)(yo) = 0 ⇔ E[(Xi − Xn)e−rξ] = 0 ∀ r ≥ 0, i = 1,..,n − 1
Let Z = e X 0 e –m a.s. The condition (Grad h)(yo) = 0 becomes:
E ( ξ j Z r ) = E ( ξ n Z r )     r   >   0     j = 1 , . . . , n 1
Which further implies:
E(ξjf(Z)) = E(ξnf(Z)) for any polynomial function f, ∀ j = 1,...,n − 1
As the polynomials are dense in C([0.e−m]), equality (19) holds for any bounded continuous function, f. The standard procedure is: Approximate the indicators 1(a,b] with continuous functions and check that (19) implies the equality E(ξj 1(a,b](Z)) = E(ξn 1(a,b](Z)) ∀ j = 1,...,na < b ∈ ℝ. The conclusion is that E(ξj|Z) = E(ξn|Z) for any j = 1,...,n − 1. As Z and X 0 generate the same σ-algebra; we see that condition (5) is satisfied. The proof is finalized by applying Proposition 4. □
Remark. 
In the case when ξI are rates of returns of assets in a financial market, we have ξi  1 for all i. The hypothesis that ξi are bounded below from the preceding proposition is verified.
Example. 
Suppose that ξj ~ Gamma(aj,aj), j = 1,...,n with aj > 0 and (ξj)1≤j≤n are independent. Then ξ ∈ CARAAO (xo) with:
x o = a 1 S , a 2 S , , a n S , S = j = 1 n a j
Indeed, we have to minimize the function:
h ( x 1 , . . . , x n 1 ) = ln [ E e r x 1 ξ 1 + + x n 1 ξ n 1 + 1 x 1 x n 1 ξ n ]   with   x   P n 1
As (ξj)j are independent,
h ( x ) = a 1 l n a 1 a 1 + r x 1   + a 2 l n a 2 a 2 + r x 2 + . . . + a n l n a n a n + s x n   with   x n = 1 x 1   . . .   x n 1
After some calculus, one obtains the gradient g = Grad(h). Its components are:
g i ( x ) = r 2 a i x n a n x i a i + r x i a n + r x n   thus Grad ( h ) = 0     x i = a 1 a n x n
Therefore, xo = a 1 S , a 2 S , , a n S . From the above proposition, xo is an absolute optimal portfolio.
Remark. 
Even for n = 2, it is not true that if ξ1 and ξ2 are independent and Eξ1 = Eξ2, then (ξ12) has the property AO. Suppose that ξ1 ~ Uniform(−a,a) and ξ2 ~ Uniform(−b,b) for somea,b > 0, a ≠ b. Then the minimum point of the function
h x = E e r x ξ 1 r 1 x ξ 2 = s i n h ( a r x ) s i n h ( b r 1 x a b r 2 x 1 x ,   0     x     1
depends strongly on r; hence, ξ cannot belong to CARAAO.
Counterexample. 
Proposition 7 says that if the CARA-absolute optimal portfolio is in the interior of the simplex Δn, then in many cases it is an absolutely optimal portfolio as well. However, if xo is not an interior point, if it belongs to the face Δn,k, k ≥ 1 of the simplex, that fails to be true. Let x1 = (0;1), x2 = (0;3), x3 = (1;3), and x4 = (3;2) four points in the plane, and suppose that (ξ12) ~ x 1 x 2 x 3 x 4 α γ δ β . Then ξ1 ~ 0 1 3 α + γ δ β , ξ2 ~ 1 2 3 α β γ + δ .
The condition that xo = (0;1) is the absolute optimal portfolio is that:
E(ξ12a) ≤ E(ξ12a) ∀ a
One can easily check that xo = (0;1) is the absolute optimal portfolio if and only if β ≤ α; xo = (0;1) is an absolute CARA-optimal portfolio if and only if [ β 2 α ≤ 1 and 3γ + 2δ ≥ β − α] or [ β 2 α   > 1 and 3γ + 2δ ≥ β 2 α 2 ]. ξ1 icv ξ2 ⇔ Eξ1 ≤ Eξ2 ⇔ 3γ + 2δ ≥ β − α.
Thus, there exists no implications that “x0 is an absolute CARA-optimal portfolio, ⇒ x0 is an absolute optimal portfolio” or even “ξ1icv ξ2xo = (0;1) is an absolute CARA-optimal portfolio”, as we thought.

5. The Normal Case

This is the classic case. The literature concerning it is extensive. The unidimensional normal distributions are very convenient because the “icv” order is easy to establish:
Normal (μ,σ2) icv Normal (μ’,σ’2) ⇔ μ ≤ μ’, σ ≥ σ’.
Moreover, if ξ ~ Normal (μ,C) is a n-dimensional random vector and x ∈ Δn(S) is a portfolio of sum S, then xTξ ~ Normal(xTμ, xTCx). Therefore, it is very easy to compare two portfolios x and y:
xTξicv yTξxTμyTμ and xTCxyTCy
Because these relations are homogeneous, the sum S does not matter; hence, we always can restrict ourselves to consider only portfolios from Δn.
We can now answer the question: when does ξAO provided that ξ ~ Normal (μ,C)?
Assuming the convention INC(k), we agree that μ1 ≤ μ2 ≤…≤ μk < μk+1 = μk+2 =…= μn.
Proposition 8.
Let ξ ~ Normal (μ,C). Then ξAO if the Pareto problem
maximize   x T μ minimize   x T C x x T · e = 1 ,   x 0 ,
admits at least solution.
Proof. 
Obvious from (20). □
Proposition 9.
Let ξ ~ Normal (μ,C) ∩ INC(k). Then
(i)
If k = 0, then ξAO;
(i)
If k ≥ 1 then ξAO if at least one solution xo of the problem
  m i n i m i z e   x T C x x T · e = 1 ,   x 0 ,  
has the property that x j 0 = 0 ∀ jk. A sufficient condition for that to happen is that
ci,jcn,jj = k + 1,...,ni = 1,...,k
Proof. 
(i)
There is nothing to prove: all the portfolios have the same expectation.
(ii)
(Replace xn with 1 − x1x2 − … − xn−1 and let f(x) = Var(xTCx). Then:
f x = 1 i , j n 1 c i , j x i x j + 2 1 i n 1 c i , n x i 1 x 1 x n 1 + c n , n 1 x 1 x n 1 2
Its gradient g = Grad(f) has the components:
g i x = j = 1 n c i , j c n , j x j , 1 i n 1
We claim that the inequalities (23) imply that f attains a minimum at points of the form xo = (0,..., 0, xk+1,..., xn−1). Notice first that (23) implies that gi(x) ≥ 0 ∀ 1 ≤ ik for points of form x = (0,..., 0, xk+1,..., xn−1), because g i x = j = k + 1 n c i , j c n , j x j . In order to prove our claim, let us consider the convex function h: [0,∞) → ℝ, h(s) = f(sx1,...,sxk, xn-k+1,...,xn−1).
Note that h’(s) = i = 1 k x i g i s x 1 , , s x k , x k + 1 , x n 1 ; hence:
h ( 0 ) = i = 1 k x i g i 0 , , 0 , x k + 1 , x n 1     0
However, h is convex, hence its derivative is non-decreasing, hence it is non-negative, therefore h is non-decreasing itself. It follows that:
h(0) ≤ h(1) ⇔ f(0,..., 0, xk+1,..., xn−1) ≤ f(x1,..., xk, xk+1,..., xn−1)
In conclusion, f can attain its minimum only in points of the form (0,...,0,xk+1,...,xn−1). g i x = j = 1 n c i , j c n , j x j , 1 i n 1 . □
Corollary 9.
Let C be a non-negative defined matric. Let ci,⋅ be its i’th row of C. A sufficient condition that ξ ~ N(μ,C), ξ ∈ INC(k) belong to AO is that ci,⋅cn,⋅ ∀ i = 1,2,...,k.
Example. 
Take μ = 1 2 2 and C = 6 3 2 3 2 1 2 1 2
Note that ξ ~ N(μ,C) ∈ AO(xo) for some xo from Δ3,1 because its first row is greater than the third one (each component of the first row is greater than the corresponding component of the third row).
One sees that xo = (0, ½, ½) implies X 0 ~ N(2, 3 2 2 ). This is the absolute optimum.
Remark. 
The conditions from Corollary 3. are only sufficient, not necessary. For instance, the vector ξ ~ N(μ,C) where μ = 1 a b and C = 9 0 1 0 9 1 1 1 1 , 1 ≤ ab admits xo = (0, 0, 1). Hence, X 0 ~ Normal(b,1), in spite of the fact that neither the first row nor the second one are greater than the third row of the matrix.
Remark. 
If we want the absolute optimum porfolio to be of the form xo = (0,x2,...,xn), then the necessary and sufficient condition is that c1,2 ≥ cn,2,..., c1,n ≥ cn,n. This is the same as the condition c1,⋅bn,⋅ Indeed, because ci,j = σiσjρi,j with Var(ξi) = σi2 and ρi,j is the correlation coefficient between ξi and ξj, then the inequality c1,i ≥ cn,i is equivalent to σ1ρ1,i ≥ σnρn,i ∀ i = 2,...,n; hence, σ1 ≥ σn m a x ρ n , 2 ρ 1 , 2 , ρ n , 3 ρ 1 , 3 , , ρ n , n ρ 1 , n and this clearly implies c1,1 ≥ c1,n ⇔ σ1 ≥ σnρ1,n. However, for k ≥ 2, this is not true anymore. For instance, for k = 2, the conditions “ci,j ≥ cn,j ∀ j = 3,...,n ∀ i = 1,2” and “c1,⋅cn,⋅, c2,⋅cn,⋅” are not the same; the second one is stronger.
For n = 4, an example could be C = 16 0 4 2 0 9 1 2 4 1 4 1 2 2 1 1 .
Here, σ1 = 4, σ2 = 3, σ3 = 2, σ1 = 1. The minimum point is of form xo = (0,0,α,1 − α), in spite of the fact that neither the first row nor the second one are greater than the fourth one. Indeed, the components of the gradient are:
g1(x) = 14x1 − 2x2 + 3x3 + x4, g2(x) = − 2x1 + 7x2 + x4
and has the property that g1(x) ≥ 0, g2(x) ≥ 0 for x ∈Δ4,2. The reader may check that xo = e4. The optimal portfolio is e4 = (0,0,0,1) whenever ξ~N(μ,C) with μ1 ≤ μ2 < μ3 = μ4.

6. The Case n = 2

In this section, we consider the case when the random vector ξ is bi-dimensional. This is the next simple case. Now ξ = (X,Y) and instead of writing OA(s,t) with s,t ≥ 0, s + t = 1 we write OA(s). We shall give the method of how to construct probability distributions on ℝ2 with property OA(s), s ∈ [0,1). We shall assume that ξ either has a density f or has a discrete distribution on the set of integers. The same letter f will denote the density in the first case and the probability law in the second one: f(x,y) will denote P(X = x,Y = y). We shall state the conditions in both cases.
The case s > 0. We want the absolute optimum return to be Z = sX + (1 − s)Y. According to Proposition 4, the condition is that E(X|sX + (1 − s)Y) = E(Y|sX + (1 − s)Y).
In terms of densities, the condition means that:
s x + 1 s y = a x y f x , y d x = 0 a   Absolutely   continuous   case
s x + 1 s y = a x y f x , y = 0 a   Discrete   case
Or, explicitly written:
x a s x 1 s f x , a s x 1 s d x = 0 a Absolutely   continuous   case
x x a s x 1 s f x , a s x 1 s a   Discrete   case
We shall focus on the absolute continuous case. After the substitution u = x − a in (27) one obtains:
u f a + u , a λ u d u = 0 a   with   λ = s 1 s
Let us put ρ(a) = 1 1 s f a + u , a λ u d u . Then ρ is a probability density. Let also
p a ( u ) = f a + u , a λ u 1 s ρ a
Then pa is also a probability density and (28) becomes:
u p a u d u = 0   a
The meaning of (30) is that, if Za are random variables with density pa, then EZa = 0. This is very easy to construct; take any random variables and center them. The conclusion is that we may construct as many distributions FOA(s) as follows:
(i)
Take a probability density on [0,∞), denoted by ρ;
(ii)
Take a family of densities on the real line, pa, with property (30);
(iii)
Set f a + u , a λ u = 1 s p a u ρ a or substituting a + u by x and a λ u by y, we obtain:
f ( x y ) = ( 1 s ) p s x + 1 s y ( ( 1 s ) ( x y ) ) ρ ( s x + ( 1 s ) y )   and   F = f λ 2
where λ 2 is the Lebesgue measure in the plane.
Then FOA(s).
Example 1.
Take pa = 1 2 a 1 a , a and ρ = 1[0, 1]. Then:
f ( x , y ) = ( 1 s ) p s x + 1 s y ( ( 1 s ) ( y x ) ) ρ ( s x + ( 1 s ) y ) = 1 s 2 y s y x 1 Δ x , y
where Δ is the interior of the triangle ABC with A(0,0), B(0, 1 1 s ) , C( 1 + 2 s 2 s , 1 4 s 2 4 s 1 s ).
Notice, as a particular case, that if s = ½, then f(x,y) = 1 2 x + y 1 Δ x , y with A(0,0), B(0,2), C(2,0); and now the distribution F is symmetric. The reason for this is that all the densities, pa, are symmetrical. The general form of a density of a distribution FOA(½) is:
f ( x , y ) = 1 2 p x + y 2 x y 2 ρ x + y 2
which, of course, may not be symmetric.
A sufficient condition that the equality (28) holds is that the density, f, satisfies the condition:
f(a + x,aλx) = f(ax, a + λx) for any a,x
This condition is similar to the symmetry. We can construct many such densities, starting with a symmetric density. It is enough to take a symmetric density, call it g, (meaning that g(x,y) = g(y,x) ∀ x,y !). The reader can check that:
f ( x , y ) = 2 1 s g 2 s 1 1 s x + 2 y , x 1 s
is a density that satisfies (32). For s = ½ we discover again the symmetric densities because (33) becomes f(x,y) = 4g(2y,2x).
Example 2.
g(x,y) = 1[0, 1]×[0, 1](x,y) ⇒ f(x,y) = 2 1 s 1D(x,y) where D = x , y 2 : 0 x 1 s , 0 2 s 1 x + 2 2 s y 1 s .
This is the uniform distribution in the interior of the parallelogram ABCD with A(0,0), B(t, (s − t)/2), C(t,s), D(0, ½). Here t = 1 − s. If g would be the uniform density in the interior of the unity circle, then f would be the density for the uniform density in the interior of some ellipse, etc.
The case s = 0. The condition is that E(X; Ya) ≤ E(Y; Ya) ∀ a. A sufficient condition is that E(X|Y) ≤ Y. Otherwise written, the pair (Y,X) should be a sub-martingale (or rather the first two terms of a sub-martingale). It is very easy to construct such distributions.

7. Conclusions

Our work was motivated by a result from mathematical folklore, which states that in the case of a financial market where the asset rates of return are i.i.d., the equal weight portfolio was an optimal portfolio for all risk-averse investors. Our attention was focused on finding necessary and sufficient conditions for the distribution of a financial market so that the equal weight portfolio is the optimal portfolio for all risk-averse investors. We generalized the previous problem by replacing the equal weight portfolio with a given portfolio. The necessary and sufficient conditions found were formulated using the conditional mean. In case the financial market has two assets, we have provided an algorithm for construction of the financial market that admits a given optimal portfolio. It is a challenge to find algorithms for the construction of probability distributions with property AOP in the case n > 2. The study in this paper could be developed considering sets of utility functions whose derivatives have a constant sign. For example, one can study existence conditions for absolute portfolios when the set of utility functions is composed from U 2 utility functions that have a non-negative third derivative. More generally, one can investigate existence conditions for absolute portfolios in the case where the set U of utility functions is composed of functions u with the property that 1 i u i   0 , i = 1, 2,…, k.

Author Contributions

Conceptualization, G.Z. and M.R.; methodology, G.Z., C.Z.R. and M.R.; validation, G.Z., M.R. and C.Z.R.; resources, M.R.; writing—original draft preparation, G.Z., M.R. and C.Z.R.; writing—review and editing, C.Z.R.; supervision, G.Z., M.R. and C.Z.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Samuelson, P.A. General proof that diversification pays. J. Financ. Quant. Anal. 1967, 2, 1–13. [Google Scholar] [CrossRef] [Green Version]
  2. Hadar, J.; Russell, W.R. Diversification of interdependent prospects. J. Econ. Theory 1974, 7, 231–240. [Google Scholar] [CrossRef]
  3. Tamir, A. Symmetry and Diversification of Interdependent Prospects, Discussion Paper 139, 1975, Preprint Repec. Available online: http://www.kellogg.northwestern.edu/research/math/papers/139.pdf (accessed on 24 August 2021).
  4. Deng, X.; Li, R. Some research on value range of equal weight portfolio risk. In Proceedings of the International Seminar on Future BioMedical Information Engineering (FBIE 2008), Wuhan, China, 18 December 2008; pp. 164–167. [Google Scholar]
  5. Belzunce, F.; Riquelme, C.M.; Mulero, J. An Introduction to Stochastic Orders; Academic Press: New York, NY, USA, 2015. [Google Scholar]
  6. Levy, H. Stochastic Dominance: Investment Decision Making under Uncertainty; Springer: New York, NY, USA, 2015. [Google Scholar]
  7. Sriboonchita, S.; Wing-Keung, W.; Sompong, D.; Hung, T.N. Stochastic Dominance and Applications to Finance, Risk and Economics; CRC Press: Boca Raton, FL, USA, 2009. [Google Scholar]
  8. Shaked, M.; Shanthikumar, J. Stochastic Orders; Springer Science & Business Media: New York, NY, USA, 2007. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rădulescu, M.; Rădulescu, C.Z.; Zbăganu, G. Conditions for the Existence of Absolutely Optimal Portfolios. Mathematics 2021, 9, 2032. https://doi.org/10.3390/math9172032

AMA Style

Rădulescu M, Rădulescu CZ, Zbăganu G. Conditions for the Existence of Absolutely Optimal Portfolios. Mathematics. 2021; 9(17):2032. https://doi.org/10.3390/math9172032

Chicago/Turabian Style

Rădulescu, Marius, Constanta Zoie Rădulescu, and Gheorghiță Zbăganu. 2021. "Conditions for the Existence of Absolutely Optimal Portfolios" Mathematics 9, no. 17: 2032. https://doi.org/10.3390/math9172032

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop