Content uploaded by Zhenyu Cui
Author content
All content in this area was uploaded by Zhenyu Cui on Oct 13, 2021
Content may be subject to copyright.
A general approach for solving behavioral
optimal investment problems with non-concave
utility functions and probability weighting∗
Xiuchun Bi †Lvning Yuan ‡Zhenyu Cui §Jiacheng Fan ¶
Shuguang Zhang k
Abstract In this paper, we consider the optimal investment problem with both probability distor-
tion/weighting and general non-concave utility functions with possibly finite number of inflection points,
and propose a general approach for solving this problem. Existing literature has shown the equivalent
relationships (strong duality) between the concavified problem and the original one by either assuming
the presence of probability weighting or the non-concavity of utility functions, but not both. In this
paper, we combine both features and propose a step-wise relaxation Lagrange method to handle the
optimization problems under general non-concave utility functions and probability distortion functions.
The necessary and sufficient conditions on eliminating the duality gap for the Lagrange method have
been provided under this circumstance. We have applied this solution method to solve several rep-
resentative examples in mathematical behavioral finance: the CPT model which has inverse S-shaped
probability distortion and a S-shaped utility function (i.e. one inflection point), Value-at-Risk based risk
management (VAR-RM) model with probability distortion, Yarri’s dual model and the goal reaching
model. We obtain the closed-form optimal trading strategy for a special example of the CPT model,
where a “distorted” Merton line has been shown exactly. The slope of the “distorted” Merton line is
given by an inflation factor multiplied by the standard Merton ratio, and an interesting finding is that
the inflation factor is solely dependent on the probability distortion rather than the utility function.
Keywords: Non-concave Utility, Probability Distortion, Concavification, Lagrange Du-
ality, Relaxation Method.
MSC(2020): Primary: 49J55, 91B16; Secondary: 49K45, 91G80.
∗Xiuchun Bi and L.Y. contributed equally to this work.
†School of Mathematics and Statistics, Guizhou University of Finance and Economics, Guiyang Guizhou,
550025, PR China. Email: xcbi@ustc.edu.cn
‡PhD candidate at Department of Statistics and Finance, School of Management, University of Science and
Technology of China, China. Email: yln1994@mail.ustc.edu.cn
§Corresponding author. School of Business, Stevens Institute of Technology, 1 Castle Point on Hudson, United
States. Email: zcui6@stevens.edu
¶Postdoctoral Research Fellow, Chinese University of Hong Kong, Hong Kong. Email: jfan6@stevens.edu
kCo-Corresponding author. Department of Statistics and Finance, School of Management, University of Science
and Technology of China, China. Email: sgzhang@ustc.edu.cn
1
1 Introduction
Optimal portfolio choice is a classic yet ever-growing field, which concerns the fundamental ques-
tion on how to allocate financial capital among a portfolio of assets and investment products.
Starting from Merton’s seminal work (Merton,1969), the expected utility maximization has be-
come the dominant model to characterize the individual’s investment decision making process
in continuous time. The classical expected utility theory (EUT) requires the following key as-
sumptions: the utility function U(·) is strictly concave (i.e. decision maker is risk-averse); the
decision maker is sufficiently rational to objectively evaluate the true distribution of the traded
underlying assets. There exists an extensive amount of empirical literature showing that the
EUT may fail to explain a plethora of real market observations and two main approaches have
been introduced in the literature to relax the assumptions of EUT: introduction of non-concave
utility functions, and the incorporation of subjective probability distortion/weighting.
The relaxation of the strict concavity of the utility functions in the classical EUT framework
has been well documented in the literature and the nonconcavity may arise from miscellaneous
scenarios: the composition of a concave utility function with a nonconcave pay-off scheme will re-
sult in a overall non-concave utility function, which frequently arise from the financial managers’
compensation problem, see for example Bensoussan et al. (2015); Bichuch and Sturm (2014);
Carpenter (2000); He and Kou (2018). A S-shaped utility function indicating the loss aversion
(based on a reference point) and the switching of risk propensity of individual investors natu-
rally arises from practice, see representative examples in Berkelaar et al. (2004); Kahneman and
Tversky (1979). Thirdly, objective functions can be assigned to be of special types which are
nonconcave in general, such as some risk measures, e.g. Value-at-Risk (Basak and Shapiro,2001),
or a probability of reaching a target as in the goal-reaching model, see Browne (1999,2000).
Another stream of literature that look at the modification of the classical EUT framework
considers introducing probability distortion/weighting to incorporate into the decision process
the decision makers’ subjective judgments on the distribution of wealth dynamics. There have
been abundant amount of research considering the phenomena of probability distortion, see for
example, rank-dependent utility theory (RDUT) which consists of two components: the concave
utility of the outcome and the probability weighting/distortion function (Quiggin,1982,2012),
Yarri’s dual model (Yaari,1987), household portfolio under-diversification (Polkovnichenko,2005)
and the demand for “lottery” stocks (Barberis and Huang,2008). Especially, the cumulative
prospect theory (CPT) model which incorporates an inverse S-shaped probability weighting and a
S-shaped utility function (Kahneman and Tversky,1979;Tversky and Kahneman,1992), modifies
the classical EUT framework in terms of both probability distortions and nonconcave utilities.
To handle the expected utility maximization problem with constraints, a natural way is to apply
the Lagrange method to transform the original problem into a relaxed unconstrained version and
then refer to the first order condition to characterize the optimal solutions which satisfy the given
constraints. However, the Lagrange method and conventional optimization tools such as convex
optimization and dynamic programming cannot be directly applied to the optimization problem
when both probability distortion and nonconcave utility functions are present. To the best of our
knowledge, Jin and Zhou (2008) first formulated the mathematical behavioral portfolio selection
2
problem under the CPT model in continuous time and developed a splitting method to divide
the original problem into a concave part and a convex part such that two standard point-wise
optimization problems could be formulated. One could then apply the standard Lagrange method
to obtain the optimal wealth dynamics under the additional assumption of the monotonicity
for a function related to the pricing kernel and the distortion function. He and Zhou (2011a)
generalized this type of problems by including a series of behavioral finance models via the so
called “quantile formulation”1. A substantial amount of subsequent literature arise based on
this framework, see He and Zhou (2011b), Xia and Zhou (2016), He and Zhou (2016), He et al.
(2017), Xu (2016), etc. The idea of “quantile formulation” is central to the above development
since it translates the stochastic optimization problem to an equivalent functional optimization
problem with monotonicity constraint 2.
Instead of applying the Lagrange method directly, one could recover the concavity of the ob-
jective functions or constraints by applying concavification3. By change of variables, Xu (2016),
under a RDUT model, transformed a quantile optimization problem from He and Zhou (2011b)
into a Merton-type problem and solve it by Lagrange method with the help of concavifica-
tion. Thanks to the concavification, the monotonicity condition can be removed. However, the
strict concavity assumption on the utility function still needs to be maintained so as to make
sure that the generated optimal solution is a valid quantile function. Recently, Van Bilsen and
Laeven (2020) considered a dynamic consumption problem where S-shaped utility and probabil-
ity weighting coexist, while the monotonicity conditions (c.f. Propositions 3 & 4 in Van Bilsen
and Laeven (2020)) are still required to obtain the analytical results. R¨uschendorf and Vanduffel
(2020) reformulated the optimal portfolio selection problem as an optimization problem on real
functions under monotonicity conditions and obtained a unified method to solve several problems
of interest such as the Merton problem, the rank-dependent utility theory (RDUT) problem, a
general form of the CPT problem as well as Yaari’s optimal investment problem. Boudt et al.
(2020) derived explicit solutions to the optimal payoff problem for a general Yaari investor in a
variety of relevant cases. He and Jiang (2021) solved the Yaari problem under non-negativity
constraints in a slightly more general set-up.
Among all the literature we discussed above, without assumptions on the monotonicity condi-
tions, either the strict concavity of the utility functions or the non probability-distortion frame-
work need to be maintained in order to apply the Lagrange method. In this paper, our focus is to
propose a general approach for solving behavioral optimal investment problems with nonconcave
utility functions and probability weighting without the monotonicity assumption. The method
can handle more general settings and yield fully explicit solutions that are new to the literature.
1This idea of “quantile formulation” finds its pedigree in a more general “rearrangement formulation” for this
kind of optimization problems (Bernard et al.,2014a,2015b,2014b;Burgert and R¨uschendorf,2006;R¨uschendorf,
1983).
2To ensure the solutions generated by the Lagrange function is the true optimal solution to the original problem,
i.e. to be a valid quantile function, one needs to presume the monotonicity of a function related to the pricing
kernel and the distortion function.
3A recent paper Dai et al. (2019) presented solutions to non-concave utility maximization problems with
bounded portfolio constraints, which makes the concavification principle fail. However, the probability distortion
is not considered there.
3
It is of interest to combine the general version of quantile formulation with the Lagrangian relax-
ation method, hence the key question we address in this paper is how far we can go by applying
the Lagrange method to solving a general quantile optimization problem when both non-concave
utilities and probability distortions4are present.
Due to the existence of non-concavity and probability weighting, a straightforward point-wise
optimization may not admit a valid quantile function which needs to be right continuous and
non-decreasing. In order to tackle this difficulty, we propose the solution procedures as follows.
First we set up a “relaxed” form of the original problem by replacing the utility function U
and the distortion related function ϕby their concave envelopes ˆ
Uand ˆϕrespectively, and then
solve this problem pointwisely by referring to the first order condition of the corresponding
Lagrange function. The value of the concavified problem is essentially an upper bound of the
original value function since the main mathematical elements have been relaxed to their concave
envelopes. Furthermore, it is well known that the concave envelopes of non-concave functions
could be locally affine (Reichlin,2013), as a result of the non-strict concavity, hence the optimal
solutions may not be unique. Based on meticulous analysis of the optimal solutions to the relaxed
point-wise optimization problem, we characterize the set containing all possible solutions which
are formally non-deceasing and right continuous on the domain of the objective function. We
then take a bottom-up approach by substituting those solutions back to the original problem by
removing the relaxation in a step-wise way 5to see if the optimizer derived from the point-wise
optimization problem satisfies the given constraint of the unconcavified problem. If that is the
case, then there must exist optimal quantiles to the original unconcavified optimization problem.
However, to complement the result of Xu (2016) that there is no gap between the two problems
when the utility is assumed to be strictly concave, our results show that the concavification
may result in a strictly positive gap when we remove the concavity assumption on the utility
function. The existence of the gap is determined by the existence of the Lagrange multiplier. We
shall show in the following analysis that the existence of the Lagrange multiplier can be easily
justified by checking whether the initial value X0=x0belongs to some given intervals which
can be explicitly expressed in terms of the parameters of the utility function and the probability
distortion function.
To apply the solution method aforementioned to concrete behavioral finance examples, we have
considered a special case of the CPT model where the distortion functions are assumed to be the
same on both sides of the reference point and the S-shaped utility is designed to be bounded from
below. The optimal wealth process and corresponding trading strategies have been completely
solved in explicit closed-form when the initial wealth lies within the “solvable intervals” that
we mentioned above. According to the solutions, one is able to clearly observe how investors’
probability weighting and switching risk propensity influence their portfolio selections. We com-
pare the strategies to the original Merton problem where there exists neither non-concavity in
utility functions nor probability distortions. We find that the existence of probability distor-
4Notice that the classical utility maximization problem can still be attributed to the quantile formulation
where the distortion function is w(x) = x. By the nature of the problem, to find the optimal terminal wealth is
equivalent to find the optimal distribution of the terminal portfolio pay-off.
5See problems A, B, C, D in the following context.
4
tions shall fundamentally change the trading pattern of the investors while the non-concavity
of the utility function would not. The resulting optimal strategies from distorted beliefs on the
portfolio dynamics tend to exhibit higher holdings of the risky assets along with the increase
of portfolio value as compared to cases without distortions. Both strategies converge to a con-
stant level where the non-distorted strategies converge to the standard Merton line while the
distorted strategies shall converge to a constant which can be expressed as a standard Merton
ratio multiplied by an inflation factor that is strictly larger than 1. Intuitively, this means that
the presence of probability distortion increases the proportion of wealth invested in the risky
stock, as compared to the original Merton model. One possible explanation for this phenomenon
can be found from the inverse S-shaped probability distortion functions: theoretically speaking,
when the portfolio value goes to infinity, the investors are not supposed to be risk-averse anymore
since under this circumstance they would have infinite ability to afford all kinds of risks. Hence,
the over-weighting on the probability for those events which potentially generate extremely high
returns shall dominate those extremely negative events, thus an increase in the holdings of the
risky asset could be naturally expected. Similarly, to show the broad applicability of the proposed
method, we have applied the method to other examples including the utility maximization with
expected shortfall constraint (Armstrong and Brigo,2019), goal reaching model (Browne,1999)
and Yarri’s dual model (Yaari,1987), etc.
To summarize, the contributions of the paper are three-fold:
1. We propose a general approach to derive explicit closed-form solutions for optimal invest-
ment problems with both probability distortion and general non-concave utility function
with a finite number of inflection points.
2. We provide necessary and sufficient conditions on closing the duality gap for the Lagrange
method when combined with the step-wise relaxation method. The condition is dependent
on the initial wealth level, and we have explicitly identified the “solvable interval” for the
initial wealth.
3. We solved several representative examples in behavior finance using our unified approach.
In particular, we solved in explicit closed-form a special case of the CPT model in the Black-
Scholes model, and explicitly derived a “distorted Merton line”, which can be represented as
an inflation factor multiplying the standard Merton line. The inflation factor only depends
on the probability distortion, and the economic insight is that the existence of probability
distortions shall fundamentally change the trading pattern of the investors while the non-
concavity in the utility function would not.
The remainder of this paper is organized as follows. In Section 2, we introduce in details the
model formulation and key assumptions. In Section 3, the procedures on the concavification and
the main solution method are provided. The optimal solutions for the general problem and the
equivalence conditions are also investigated and listed. Following from the results in Section 3, in
Section 4a series of concrete examples are presented, and closed-form solutions of those models
with specific parameter settings are obtained. The rigorous mathematical analysis of the solution
scheme to the general problem is delegated to Section 5. We conclude the paper in Section 6with
5
a discussion of future research directions. The proofs for the main theorems and propositions
including the discussion on the well-posedness can be found in the Appendix.
2 Problem Formulation
2.1 A Continuous-Time Market
Suppose that there is a frictionless market equipped with a complete filtered probability space,
(Ω,F,P,{Ft}t≥0), where Pdenotes the physical/objective probability measure, and Ft=σ{Wτ:
0≤τ≤t}is the natural filtration, under which W:= {Wt, t ≥0}is a standard Brownian
motion. Assume that there are two assets traded continuously, with one of the assets Bbeing
risk-free (e.g. bank account or risk-free bond), and the other asset being risky (e.g. stock or
mutual fund). Given the terminal time T > 0, the dynamics of these assets on the investment
horizon [0, T ] are governed by
dSt=µtStdt +σtStdWt,
dP B
t=rtPB
tdt, 0≤t≤T,
where rtis the risk-free interest rate, µt> rtis the growth rate of the risky asset, and σtis the
volatility. All coefficients are assumed to be adapted to Ftwith
ZT
0|µt|+1
2|σt|2dt < ∞.(2.1)
Suppose that Xtis the total wealth of the agent at time t, and we require Xt≥0 for t∈[0, T ],
i.e. no bankruptcy constraint. Let πtbe the fraction/proportion of wealth invested in the risky
asset, and the remaining fraction 1 −πtwill be invested in the risk-free asset. Given the initial
wealth x0>0, the self-financing wealth process evolves according to the following stochastic
differential equation (SDE):
dXt= [rt+πt(µt−rt)]Xtdt +σtπtXtdWt.(2.2)
In order to guarantee the existence and uniqueness of the solution to (2.2), it is natural to
assume that the portfolio process πton the interval [0,T] is admissible, with its precise definition
given as follows.
Definition 2.1 A portfolio policy πt:= π(t, x)on the interval [0,T] is admissible at x:= Xt>0,
which we denote as πt∈ A(x), if it satisfies
•πtis Ft-progressively measurable;
•EhRT
0π2
tdti<∞.
Remark 2.1 Note that the subsequent analysis is not dependent on the specific stochastic process
governing the wealth process X, except that we consider a complete market setting. Our analysis
holds for a general class of continuous-time or discrete-time stochastic processes.
6
2.2 Optimization Problems
Our goal is then to choose the optimal portfolio policy πt∈ A(x), for a given current time tand
wealth level x:= Xt>0, to maximize the following objective function:
J(t, x, πt) := Z∞
0
ω(P(U(Xt,x,πt
T)> y))dy =Z∞
0
U(y)d1−ω(1 −FXt,x,πt
T(y)),(2.3)
where FXis the cumulative distribution function of X,ω(·) is the probability distortion function
which is strictly increasing and continuously differentiable. It is defined from [0,1] to [0,1] with
ω(0) = 0 and ω(1) = 1. U(·) is the utility function satisfying the following general assumption:
Assumption 2.1 1. U(·)is an upper semi-continuous and non-decreasing function mapping
from [0,+∞]to R.6
2. There exists a large positive real number M > 0, such that U=ˆ
Uon [M, ∞), where ˇ
Uis
the concave envelope of U, i.e. U(·)is ultimately a concave function on [M, ∞).7.
Remark 2.2 The restriction on the utility function U(·)is satisfied by almost all utility func-
tions considered in previous literature. In particular, the celebrated S-shape utility satisfies the
assumption with one kinked point, which is also the inflection point where the convex and concave
parts of the utility function meet.
For notation convenience, in the following, we shall denote Xt,x,πt
Tby XT. Let
v(t, x) := max
πt∈A(x)Z∞
0
U(y)d(1 −ω(1 −FXT(y))) (2.4)
be the value function. Our optimization problem can be rewritten in the following form,
v(t, x) = max
π∈A(x)Z∞
0
U(y)d(1 −ω(1 −FXT(y)))
subject to dXt= [rt+πt(µt−rt)]Xtdt +σtπtXtdWt, t ∈[0, T ],
X0=x0, Xt≥0,
(2.5)
where Xt≥0, t ∈[0, T ] is called the bankruptcy constraint. In the following, we shall denote
the initial wealth level as x0unless specified otherwise. In general, it is challenging to solve this
optimization problem due to the following two difficulties:
1. the utility function is not necessarily strictly concave, and it may not even be continuous,
which allows the possibility for the utility function to exhibit jumps.
2. there is presence of probability distortion, which destroys the time-consistency of the opti-
mization problem.
6The utility function may not necessarily be restricted to start from 0 and it can be any fixed finite constant.
One could also define the utility function on the whole real line while defining U(x) = −∞ if x < 0. Basically, in
this paper, we only consider the case that the wealth is non-negative.
7Intuitively, this means that the utility function is “eventually concave”. This is consistent with standard
utility theory, and means that the investor will not be risk-seeking when his wealth reaches a very large amount.
This technical assumption is needed to guarantee the well-posedness of the solution to the optimization problem.
The asymptotic elasticity condition AE(U) = lim supx→+∞xU0(x)/U (x)<1 is required instead in some related
literature (e.g. Bichuch and Sturm (2014)). Note that although this condition on asymptotic elasticity can imply
cases of eventually strictly concave, but it can not include the following special case : Uis eventually concave
with limx→+∞U(x)/x =k > 0.
7
To solve the dynamic stochastic optimization problem presented in (2.5), we first set up to
solve the corresponding quantile formulation of the optimization problem. As is shown in He and
Zhou (2011b), we can transform (2.5) into a functional optimization problem with the auxiliary
assumption on the pricing kernel, which is defined as
ρt= exp −Zt
0
(rs+θ2
s/2)dt −Zt
0
θsdWs,
with θt=µt−rt
σtdenoting the market price of risk.
Assumption 2.2 The pricing kernel is atomless.
Through the quantile formulation we can reduce problem (2.5) to the following functional
optimization problem:
sup
G(·)∈G Z1
0
U(G(x))ω0(1 −x)dx
subject to Z1
0
G(x)F−1
ρ(1 −x)dx =x0,
(2.6)
where F−1
ρ(·) denotes the quantile function of the pricing kernel ρ:= ρT, which is strictly in-
creasing due to Assumption 2.2. Here G(·) is the quantile function of the terminal wealth XT,
and Gis the set of all quantile functions formally defined as
G:={G(·) : (0,1) 7−→ R, increasing and right-continuous with left limits (RCLL)}.
The optimal solution X∗
Tto problem (2.5) and the optimal solution G∗(·) to the transformed
problem (2.6) satisfy the following link:
X∗
T=G∗(1 −Fρ(ρT)).(2.7)
Furthermore, applying change of variables similarly as in Xu (2016), we have
Q(x) := G(1 −ω−1(1 −x)),(2.8)
ϕ(x) := −Zω−1(1−x)
0
F−1
ρ(y)dy, x ∈[0,1].(2.9)
Then (2.6) can be transformed into:
sup
Q(·)∈G Z1
0
U(Q(x))dx,
subject to Z1
0
Q(x)ϕ0(x)dx =x0.
(2.10)
Here Q(x) is also a quantile function, which is RCLL on (0,1). ϕ(x) is differentiable and strictly
increasing on [0,1] with ϕ(0) = −E[ρ] and ϕ(1) = 0. Observe that the probability weighting
function has now been removed from the objective function in the above formulation, and only
appears in the constraint. Due to this equivalence, without loss of generality, the next section
investigates the functional optimization problem (2.10) and proposes a solution methodology
based on the concave envelopes of Uand ϕ.
8
3 Solution Method
3.1 Concave envelopes of two functions
Let ˆ
Uand ˆϕ(·) be the concave envelopes of Uand ϕrespectively. It is obvious that these two
concave envelope functions satisfy the following analytical properties:
1. They are continuous, non-decreasing and concave.
2. ˆ
U≥Uand ˆϕ≥ϕ.
3. ˆ
U(resp. ˆϕ) is locally affine on some open set of {ˆ
U=U}(resp. {ˆϕ=ϕ}).
3.2 Solution scheme
In this section, we discuss the main method to solve the functional optimization problem (2.10).
First, we shall assume that the problem (2.10) is well-posed, where well-posedness means that
the problem admits a finite maximum. The precise conditions for well-posedness are presented
in Appendix A.
Next, we propose the solution scheme through four steps of relaxation by using the Lagrangian
dual method. We will show that these four problems share a common solution set under some
technical conditions:
•Problem A:
VA(x0) = sup
Q∈G
JA(Q) = sup
Q∈G Z1
0
U(Q(x))dx, subject to Z1
0
Q(x)ϕ0(x)dx =x0.(3.1)
•Problem B:
VB(λ, x0) = sup
Q∈G
JB(Q, λ) = sup
Q∈G Z1
0
(U(Q(x)) −λQ(x)ϕ0(x))dx +λx0,(3.2)
•Problem C:
VC(λ, x0) = sup
Q∈G
JC(Q, λ) = sup
Q∈G Z1
0
(ˆ
U(Q(x)) −λϕ0(x)Q(x))dx +λx0.(3.3)
•Problem D:
VD(λ, x0) = sup
Q∈G
JD(Q, λ) = sup
Q∈G Z1
0
(ˆ
U(Q(x)) −λˆϕ0(x)Q(x))dx +λx0.(3.4)
In the above formulations, λ > 0 is a fixed constant standing for the Lagrange multiplier.
Then we shall characterize the solutions to each problem above under the given assumptions,
and provide verification results to link the solutions of the individual optimization problem to
each other. We take a bottom-up approach and start from the optimization problem D, and
then show that by restricting8the solution set by imposing certain constraints, we can recover
the solution set to the optimization problem C. Similarly, we can move froward and restrict the
solution set of problem C to obtain the solution set of problem B, until we arrive at the solution
set of problem A. Then the remaining task is to verify that the obtained solution set for problem
8Note that the solution set of problem C is a subset of the solution set of problem D due to relaxation.
9
A is not empty. We name our method the “relaxation approach” as we first relax the optimization
problems and then trace back the optimal solution set by consecutive restrictions of the solution
sets from previously solved optimization problem.
3.3 General Characterization of Optimal Solutions
In this part, we aim to find the connections between the optimal solutions to problems A-D and
give the characterizations of them.
Problem D in (3.4) can be solved by the point-wise optimization method. For a fixed x, we
consider the following problem,
Q∗(x) = arg max
Q{ˆ
U(Q(x)) −λˆϕ0(x)Q(x)}.
Consequently, the optimal solutions possess the form
Q∗(x)=(ˆ
U0)−1(λˆϕ0(x)),(3.5)
where ( ˆ
U0)−1(·) denotes the general inverse of ˆ
U0(·).
Remark 3.3 We should treat the solution form (3.5)as a composite function with respect to x.
For some special λand x < y,(ˆ
U0)−1(λˆϕ0(x)) and (ˆ
U0)−1(λˆϕ0(y)) can be totally different, even
if ˆϕ0(x) = ˆϕ0(y).
Now, we give more details for Q∗.
Proposition 3.1 The general inverse of ˆ
U0(·)may admit jump points. In particular, if ˆ
Uis
affine in some intervals, (ˆ
U0)−1must admit at least one jump point (see Figure 1and Figure 2).
Proof.
We define
(ˆ
U0)−1
−(x) = inf{y≥0|ˆ
U0(y)< x},(3.6)
(ˆ
U0)−1
+(x) = sup{y≥0|ˆ
U0(y)> x}.(3.7)
Then for any general inverse ( ˆ
U0)−1,
(ˆ
U0)−1
+(x)≤(ˆ
U0)−1(x)≤(ˆ
U0)−1
−(x).
The concavity of ˆ
Uindicates that ( ˆ
U0)−1
−(x),(ˆ
U0)−1
+(x) are left-continuous and right-continuous
functions, respectively.
Then, we can deduce that: if there exists k > 0, b > a ≥0, with
0≤(ˆ
U0)−1
+(k)≤a<b≤(ˆ
U0)−1
−(k)
or equivalently,
ˆ
U0(a+) = ˆ
U0(b−) = k,
or equivalently,
ˆ
Uis affine on interval (a, b). with slope k,
then ( ˆ
U)−1has a jump point at k.
In fact, if there exists a non-empty set (a, b)⊆ {x|ˆ
U(x)> U(x)}, then ˆ
Uis affine on [a, b] from
the definition of the concave envelope. Hence the results follow naturally. This completes the
proof.
10
Figure 1: A case of ˆ
U(·)
Figure 2: An example for non-increasing function ˆ
U0(x) and its general inverse function
(ˆ
U0)−1(x), where ( ˆ
U0)−1(·) is jumped at kand ( ˆ
U0)−1(k) can take any value in [a, b].
11
Figure 3: A multi-solution example for the situation in Lemma 3.1. Here all the solutions are optimal
to problem D while some may be optimal to problem C and problem B.
Note that the uniqueness of the optimal solution to problem D is highly related to the structure
of ˆ
U, ˆϕand λ. The jump points of ( ˆ
U0)−1play an important role in such solutions. Now, we
denote Kas the set containing all the jump points of ( ˆ
U0)−1, which are countable. We have the
following result.
Lemma 3.1 The optimal solution Q∗(x)define in (3.5)is not unique if and only if
there exists a k∈ K and at least one non-empty interval (α, β)⊆[0,1] satisfying λˆϕ0(x) = k, ∀x∈(α, β).
Proof. (ˆ
U0)−1is well-defined on R+\ K. If such a pair does not exist, which indicates
∀k∈ K,{x|λˆϕ0(x) = k}is a singleton (denoted as {y0}), then by the right continuity of Q∗(x),
we must have Q∗(y0)=(ˆ
U0)−1
−(k). Consequently, Q∗is unique.
If just one pair of k, (α, β) exists, then for any non-decreasing and right-continuous function
h: [α, β)→[( ˆ
U0)−1
+(k),(ˆ
U0)−1
−(k)], the function
Q∗(x) = ((ˆ
U0)−1
−(λˆϕ0(x)), x /∈[α, β)
h(x)x∈[α, β)
is optimal. So Q∗must not be unique.
Moreover, if there exist many pairs ki,(αi, βi), for each interval [αi, βi), then we can similarly
construct a function hiand
Q∗(x) = ((ˆ
U0)−1
−(λˆϕ0(x)), x /∈ ∪i[αi, βi)
hi(x)x∈[αi, βi)
is optimal to the problem D. This completes the proof.
Let YD,YC,YBbe respectively the optimal solution set for problems D,C,B. By studying the
connections between problems D,C,B, we have the following result.
12
Theorem 3.1
(1) For any Q∈ YD,Qis optimal to the problem C if and only if
∀(α, β)⊆ {x: ˆϕ(x)> ϕ(x)}, Q(·)is a constant function on (α, β).(3.8)
(2) For any Q∈ YC,Qis optimal to the problem B if and only if
ˆ
U(Q(x)) = U(Q(x)),∀x∈[0,1].(3.9)
Moreover, there exists a common solution for problems D,C,and B, which is given by
VD(λ, x0) = VC(λ, x0) = VB(λ, x0)
Proof. See Appendix B.1.
Remark 3.4 Figure 3represents some visualization of solutions to problems D,C,B in the in-
terval (α, β). In this special example, all functions are optimal to problem D, while Q1(x)
satisfies the second condition in Theorem 3.1 and Q2(x)satisfies the first condition. Here
c1, c2are the corresponding parameters. In particular, if (α, β)⊂ {x: ˆϕ(x)> ϕ(x)}and
(a, b)⊂ {x:ˆ
U(x)> U(x)}, then the optimal solutions can only take either aor b.
This theorem gives us the procedure on how to select the optimal solutions to problem B from
the set YD. Next, we will consider whether there exists a λso that the corresponding Q(x)∈ YB
is optimal to the problem A. By the weak duality, we must have:
VA(x0)≤inf
λ>0VB(λ, x0).
The most important task is thus to check that the strong duality
VA(x0) = inf
λ>0VB(λ, x0) (3.10)
holds, so that we can find the optimal solution to the problem A by selecting a subset from the
solution set YB.
Generally speaking, if we can find a λ∗and the corresponding Q(x)∈ YBsatisfying the
constraint: Z1
0
Q(x)ϕ0(x)dx =x0,
then the strong duality must hold and Q(x) must be an optimal solution to problem A. If not,
then the strong duality does not hold, and we cannot get the optimal solution to problem A from
the solution set YB.
In order for the readers to better understand our relaxation method and how it works, in
subsequent sections we will solve four representative behavior finance models in the literature,
obtaining closed-form or semi-closed-form solutions using our general method. In these examples,
both non-concave utility and probability distortion are present. The more general case for uand
ϕwith finitely many inflection points will be delegated to Section 5.
13
Figure 4: Shape of ϕunder different probability weighting models
4 Solutions for concrete models
Previous literature on the mathematical behavior portfolio choice considers tackling either the
first or the second challenge, i.e. the non-concave utility function or the probability distortion.
However, to the best of authors’ knowledge, there is no unified approach to handle the optimal
investment problem when both challenges are present.9Our method can solve all these combina-
tions in a unified way. Furthermore, previous literature considers the S-shape utility and the
inverse-S-shaped probability weighting as representative examples, while our proposed method
goes beyond that and can handle even more general utility functions and probability distortion
functions with a finite number of inflection points.
4.1 Example 1: A Special CPT model
In this subsection, we consider a S-shape utility function with an inverse S-shaped distortion
function. To illustrate the main ideas of our procedure, for convenience, we only consider the
case when the function ϕ(·) is also of S-shape10.
Following our procedure, we rewrite the original optimization problem as problem A:
VA(x0) = sup
Q∈G
JA(Q) = sup
Q∈G Z1
0
U(Q(y))dy,
subject to Z1
0
Q(y)ϕ0(y)dy =x0.
9The cumulative prospect theory (CPT) is one example in which both challenges appear. In the literature, the
optimal investment problem under the CPT model has been only solved for a special class of power-type S-shaped
utility functions and also for a special class of probability distortion functions, see for example sections 5.1 and
5.2 in He and Zhou (2011a).
10When the pricing kernel of the market follows log-normal distribution, almost all the inverse S-shape distortion
in He and Zhou (2016); Tversky and Fox (1995); Tversky and Kahneman (1992) will lead to a S-shape function
for ϕ, see Figure 4.
14
We introduce the Lagrange multiplier and transform problem A into its unconstrained form:
problem B
VB(λ, x0) = sup
Q∈G
JB(Q) = sup
Q∈G Z1
0
(U(Q(y)) −λϕ0(y)Q(y))dy +λx0,(4.1)
where λ > 0 is a fixed constant corresponding to the Lagrange multiplier. After replacing the
two functions U(·) and ϕ(·) by their concave envelopes step by step, we obtain
problem C
VC(λ, x0) = sup
Q∈G
JC(Q) = sup
Q∈G Z1
0
(ˆ
U(Q(y)) −λϕ0(y)Q(y))dy +λx0.(4.2)
problem D:
VD(λ, x0) = sup
Q∈G
JD(Q) = sup
Q∈G Z1
0
(ˆ
U(Q(y)) −λˆϕ0(y)Q(y))dy +λx0.(4.3)
We first solve problem D, and then go back to solve problem A through investigating the
solution sets of problem C and problem B in order. Let ξ:= inf{x > 0 : U(x) = ˆ
U(x)},
c:= inf{x > 0 : ϕ(x) = ˆϕ(x)}, then
{y≥0 : ˆ
U > U }= (0, ξ ),
{y∈[0,1] : ˆϕ>ϕ}= (0, c).(4.4)
Let λ0:= U0(ξ)
ϕ0(c). We obtain the following results on translating from problem D to problem
B:
•If λ6=λ0, then problem D can be solved by point-wise optimization. By Theorem 3.1, it
is easy to check that,
Q∗
1(λ, y) = ((U0)−1(λˆϕ(y)),if λˆϕ0(y)≤U0(ξ),
0,otherwise (4.5)
is the unique solution to problem D and is also the optimal solution to problems C and
B.
•If λ=λ0, then problem D has solutions given by:
Q∗(y) = ((U0)−1(λ0ˆϕ0(y)),if λ0ˆϕ0(y)< U0(ξ),
any non-decreasing function on [0, ξ],if λ0ˆϕ0(y) = U0(ξ).
By using Theorem 3.1, we have that
Q∗
2(y) = ((U0)−1(λ0ˆϕ0(y)),if λ0ˆϕ0(y)< U0(ξ),
0,otherwise, (4.6)
Q∗
3(y) = ((U0)−1(λ0ˆϕ0(y)),if λ0ˆϕ0(y)< U0(ξ),
ξ, if λ0ˆϕ0(y) = U0(ξ),(4.7)
are all the common solution to problems D, C, and B. In fact
Q∗
2(y) = lim
λ→λ+
0
Q∗
1(λ, y),
Q∗
3(y) = lim
λ→λ−
0
Q∗
1(λ, y).
15
Define
f(λ) = Z1
0
Q∗
1(λ, y)ϕ0(y)dy, λ 6=λ0,
xL=Z1
0
Q∗
2(y)ϕ0(y)dy =f(λ+),
xU=Z1
0
Q∗
3(y)ϕ0(y)dy =f(λ−),
f(λ0) = xLor xU.
and we have the following result.
Theorem 4.2 Assume that f(λ)<∞for any λ∈R+, then
(1) if x0/∈[xL, xU], then Q∗
1(λ∗, y)is the unique solution to problem A and VA(x0) =
VB(λ∗, x0),where λ∗is the unique solution to f(λ) = x0;
(2) if x0=xL, then Q∗
2(y)is the unique solution to problem A and VA(xL) = VB(λ0, xL);
(3) if x0=xU, then Q∗
3(y)is the unique solution to problem A and VA(xU) = VB(λ0, xU);
(4) if x0∈(xL, xU), then the strong duality does not hold, and our method fails to find the
optimal solution.
Proof. The proof follows straightforwardly from the strictly decreasing property of f(λ).
We next give an explicit numerical example. Let the utility function be
U(x) = ((x−η)γ,if x≥η,
−k(η−x)γ,if 0 ≤x < η, (4.8)
where ηis the reference point of the agent, 0 < γ < 1 is the risk aversion factor and k > 1 (see
Figure 5), then Uis a S-shape utility function.
For convenience, we consider a particular probability distortion function from He and Zhou
(2016) and Jin and Zhou (2008) (see Figure 7):
w(x) = (k1Φ(Φ−1(x) + a1),if x≤b,
1−k2Φ(a2−Φ−1(x)),if x > b, (4.9)
where a1≥0, a2≥0, b ∈[0,1], and
k1=exp{(a1+a2)Φ−1(b) + a2
1/2}
exp{(a1+a2)Φ−1(b) + a2
1/2}Φ(Φ−1(b) + a1) + exp{a2
2/2}Φ(a2−Φ−1(b)),
k2=exp{a2
2/2}
exp{(a1+a2)Φ−1(b) + a2
1/2}Φ(Φ−1(b) + a1) + exp{a2
2/2}Φ(a2−Φ−1(b)).
Now we assume that the market parameters r(·), µ(·), σ(·) are all deterministic. Let
˜µ=−ZT
0
(r(t) + θ2
t/2)dt, ˜σ=sZT
0
θ2
tdt, (4.10)
At=−ZT
t
(rs+θ2
s/2)ds, Bt=sZT
t
θ2
sds, (4.11)
16
Figure 5: Uand its concave envelope ˆ
U, with
parameters: η= 1, k = 2, γ = 0.5
Figure 6: ϕand its concave envelope ˆϕwith
parameters: r= 0.01, µ = 0.07, σ = 0.15, T =
1, a1= 1.17, a2= 0.86, b = 0.33
then,
ϕ(x) = −exp{˜µ+ ˜σ2/2}Φ(Φ−1(w−1(1 −x)) −˜σ)
=(−exp{˜µ+ ˜σ2/2}Φ(a2−˜σ−Φ−1(x
k2)) if x < 1−w(b),
−exp{˜µ+ ˜σ2/2}Φ(Φ−1(1−x
k1)−a1−˜σ) if x≥1−w(b).
Here we assume a2>˜σto guarantee that ϕ(x) is a S-shaped function, see an illustration in
Figure 6.
Using the notations ξ, c in (4.4), At, Btin (4.11), we obtain
xL=−ηϕ(c) + k1γ
λ01
1−γ
exp (a1˜σ+a2
1/2−˜µ)γ
1−γ
+γ2(a1+ ˜σ)2
2(1 −γ)2ΦΦ−11−c
k1+γ(a1+ ˜σ)
1−γ,
xU=xL+ (ϕ(c)−ϕ(0))ξ.
(4.12)
Theorem 4.3 Assume f(λ)<+∞for any λ∈R+and a2>˜σ.
(1) If x0≤xL, then f(λ) = x0admits a unique solution λ∗, and the optimal terminal wealth
is given by
X∗
T= (η+c2ρ−c3)1{ρ≤c1},(4.13)
17
Figure 7: An example of the distortion function in (4.9) when parameters take: a1= 1.17, a2= 0.86, b =
0.33
where
c1= exp (˜σ(ln k1−a2
1/2 + ln U0(ξ)
λ∗) + a1˜µ
a1+ ˜σ),
c2=k1γ
λ∗1
1−γ
exp −1
1−γa2
1/2−a1˜µ
˜σ,
c3=1
1−γ1 + a1
˜σ.
(4.14)
The optimal wealth process is
X∗
t=ηexp{B2
t/2 + At}Φln c1/ρt−At
Bt
−Bt
+c2exp{(1 −c3)At+ (1 −c3)2B2
t/2}ρ−c3
tΦln c1/ρt−At
Bt
−(1 −c3)Bt,
and the optimal portfolio is
π∗
t=µt−rt
σ2
tX∗(t)η
Bt
exp{B2
t/2 + At}φ(ln c1/ρt−At
Bt
−Bt)
+c2c3exp (1 −c3)At+ (1 −c3)2B2
t/2ρ−c3
tΦln c1/ρt−At
Bt
−(1 −c3)Bt
+c2
Bt
exp{(1 −c3)At+ (1 −c3)2B2
t/2}ρ−c3
tφln c1/ρt−At
Bt
−(1 −c3)Bt.
Moreover,
X∗
t→+∞, π∗
t→µt−rt
σ2
t(1 −γ)1 + a1
˜σ,as ρt→0,
X∗
t→0, π∗
t→+∞,as ρt→+∞.
(4.15)
18
(2) If x0≥xU,then f(λ) = x0admits a unique root λ∗, and the optimal wealth is given by
X∗
T=η+c2ρ−c31{ρ≤c1}+c2c−c3
11{ρ>c1}(4.16)
where
c1= exp ˜σΦ−11−c
k1+ ˜µ−a1˜σ,
c2=k1γ
λ∗1
1−γ
exp −1
1−γ−a1˜µ
˜σ+a2
1/2,
c3=1
1−γ1 + a1
˜σ.
The optimal wealth process
X∗
t=ηexp{At+B2
t/2}+c2exp{(1 −c3)At+ (1 −c3)2B2
t/2}ρ−c3
tΦln(c1/ρt)−At
Bt
−Bt(1 −c3)
+c2c−c3
1exp{B2
t/2 + At}1−Φln(c1/ρt)−At
Bt
−Bt,
and the optimal portfolio
π∗
t=µt−rt
σ2
tX∗(t)c2c3exp{(1 −c3)At+ (1 −c3)2B2
t/2}ρ−c3Φln c1/ρt−At
Bt
−(1 −c3)Bt
+c2
Bt
exp{(1 −c3)At+ (1 −c3)2B2
t/2}ρ−c3φln c1/ρt−At
Bt
−(1 −c3)Bt
−c2
Bt
c−c3
1exp{B2
t/2 + At}φln c1/ρt−At
Bt
−Bt.
Moreover,
X∗
t→+∞, π∗
t→µt−rt
σ2
t(1 −γ)1 + a1
˜σ,as ρt→0,
X∗
t→(η+c2c−c3
1) exp{At+B2
t/2}, π∗(t)→0,as ρt→+∞.
(4.17)
Proof. See Appendix C.1.
One can observe that the analytical solutions given by (4.13) and (4.16) contain a lot of
interesting economic insights. First, based on the quantitative structure of the solutions, the
value of the optimal wealth, if not trivial (i.e. X∗
T6= 0), would always be obtained from the right
side of the reference point, which indicates that the investor shall drive the optimal expected
utility from the concave part of her utility function. This finding can be intuitively explained
since the “true” utility function we are using for the computation is the concavified version of the
original one, thus the optimization was conducted by nature on a concave function. Second, we
can clearly observe from the limiting forms of the optimal trading strategies (4.15) and (4.17),
that a new type of “Merton Line” has appeared in the limiting expression when the optimal wealth
evolves to infinity. When there exists probability distortion, the “Merton Line” is significantly
amplified by a positive inflation factor as compared to the original form of the classical Merton
solution. This shows that the probability distortion shall fundamentally change the trading
structure of the investment process and consistently increases the holding of risky assets when
the accumulated wealth goes to infinity. Simultaneously, when the wealth of investors lies at
a relatively low level, the investor shall significantly decrease the exposure to risky asset which
can be observed directly from the Figures 8and 9. This coincides with the intuition behind the
construction of the probability distortion functions, i.e. the low probability events have been
distorted to be overestimated by the positive prospect holding of the investors.
19
Figure 8: Optimal strategy π∗
trespect to wealth
level when parameters taking: r= 0.01, µ =
0.07, σ = 0.15, T = 1, t = 0.5, a1= 1.17, a2=
0.86, b = 0.33, k = 2, γ = 0.5, η = 1, x0= 0.2
Figure 9: Optimal strategy π∗
trespect to wealth
level when parameters taking: r= 0.01, µ =
0.07, σ = 0.15, T = 1, t = 0.5, a1= 1.17, a2=
0.86, b = 0.33, k = 2, γ = 0.5, η = 1, x0= 1.5
If we take a closer look at Figures 8and 9, then we can find that the optimal proportion being
invested in the risky asset seems to experience no change for the non-distorted case if we vary
the initial endowment x0, which can be observed in a clearer way in Figure 10. We can see from
Figure 10 on the bottom right that all the investment ratios corresponding to different values of
x0will coincide with each other once the values of X∗
tare the same for the non-distorted situation.
We can mathematically show that under the setting of our CPT framework, the distortion shall
ruin the time consistency of the individual investor’s investment strategies from Theorem 4.3
since different x0values result in different c1and c2. If there exists no distortion, then using the
notations in (4.10) and (4.11), we get the optimal terminal wealth
X∗
T= (η+Acpρ−p)1{ρ≤c},
where p=1
1−γ, A =γ
u0(ξ)p,and cis the unique solution that solves: E[ρX∗
T] = x0. It is clear
that different x0will lead to different values of c. Since log(ρ/ρt)∼N(At, B2
t) and is adapted to
Ft, the optimal wealth process can be computed as
X∗
t=ηc1Φ(dt(c/ρt)) + c2(c/ρt)pΦ(dt(c/ρt) + pBt)
with c1exp{At+B2
t/2}, c2=Aexp{(1 −p)At+ (1 −p)2B2
t/2}, dt(x) = log x−At
Bt−Bt, and the
optimal value invested in stock is
π∗
tX∗
t=µt−rt
σ2
tηc1
Bt
φ(dt(c/ρt)) + c2
Bt
(c/ρt)pφ(d(c/ρt) + pBt) + c2p(c/ρt)pΦ(dt(c/ρt) + pBt).
As a result, X∗(t) and π∗(t) are two functions with respect to c/ρt, and X∗
tis strictly increasing
with respect to c/ρt, then the same X∗(t) shall result the same c/ρt, which will also lead to the
same π∗
t.
20
Figure 10: Comparison of optimal wealth and investment strategies without probability distortion
under different levels of initial wealth
4.2 Example 2: VaR-RM model with distortion function and non-
concave utility
In this example, we consider an extension of the Value-at-Risk based risk management (VaR-RM)
model proposed in Basak and Shapiro (2001) and solve it using our new method. The model1is
sup
X≥0Z+∞
0
U(x)d(1 −w(1 −FX(x))),
s.t. E[ρX] = x0,
Hg(X)≤ −x1.
(4.18)
Here U(·) is a non-concave utility function (e.g. S-shaped utility), H(·) is a given risk measure,
and
Hg(X) = −Z1
0
G(x)m(dx),
where G(·) is the quantile function of X,mis a fixed probability measure on [0,1]. The basic
limitations of (x0, x1) is clearly analyzed in Wei (2018).
By using the quantile formulation X=G(1 −Fρ(ρ)) and change of variables, we have
Q(x) = G(1 −ω−1(1 −x)),
1To our best knowledge, there are two type of extensions to Basak and Shapiro (2001): one considers a weighted
VaR as the constraint(see Wei (2018) ); the other considers the non-concave utility with Var/ES constraint (see
Armstrong and Brigo (2019)). Here we use an example combining both non-concave utility and weighted VaR
constraint with probability distortion.
21
and we can transform the above problem into the form of problem A:
sup
G(·)∈G Z1
0
U(Q(x))dy,
s.t. Z1
0
Q(x)ϕ0(x)dx =x0,
Z1
0
Q(x)m1(dx)≥x1,
(4.19)
where
ϕ(x) = −Zw−1(1−x)
0
F−1
ρ(y)dy,
and
m1(dx) = mdx
w0(w−1(1 −x)).
To tackle this problem, we need to first use the Lagrange multiplier method to transform it into
the non-constraint form, i.e. problem B:
sup
QZ1
0
(U(Q(x)) −λ1ϕ0(x)Q(x))dx +λ2Z1
0
Q(x)m1(dx) + λ1x0−λ2x1.(4.20)
Let
ϕλ(x) := ϕ(x) + λm1([x, 1]),
where λ=λ2/λ1, then we can transform the problem into
sup
QZ1
0
(U(Q(x)) −λ1ϕ0
λ(x)Q(x))dx +λ1(x0−λx1).(4.21)
Now, we turn to solving problem (4.21) by using the concavification procedure. We shall first
solve the following, i.e. problem D
sup
QZ1
0
(ˆ
U(Q(x)) −λ1ˆϕ0
λ(x)Q(x))dx +λ1(x0−λx1).(4.22)
It is easy to analyze the optimal solution of this problem and we can then get the optimal solution
of (4.21) by referring to Theorem 3.1.
However, for the general form of U(·), m(·), it is hard for us to find the explicit conditions
under which the Lagrange duality gap will diminish. Thus, we consider a simple case: Uis a
S-shaped utility (i.e. first strictly convex then strictly concave), w(x) = x, and the risk measure
is the expected shortfall (Armstrong and Brigo,2019;Wei,2018). Then the problem is
sup
G(·)∈G Z1
0
U(G(x))dy,
s.t. Z1
0
G(x)F−1
ρ(1 −x)dx =x0,
1
αZα
0
G(x)dx ≥x1,
(4.23)
where x1, x0>0.
We denote the concave envelope of Uas ˆ
Uand define ξ= inf{x > 0 : ˆ
U(x) = U(x)},then
{x≥0 : ˆ
U > U }= (0, ξ ),
and
U0(ξ)ξ=U(ξ)−U(0).
22
We then calculate the function
ϕλ(x) = −Z1
x
F−1
ρ(1 −y)dy +λ1−x
α1{x<α}.(4.24)
It is easy to see that ϕλ(x) is locally concave on (0, α) and (α, 1), but not globally concave on
(0,1). Thus there must exist 0 < y1≤α≤y2<1 (here if λ= 0, then y1=α=y2) such that
ϕ0
λ(y1) = ϕ0
λ(y2) = ϕλ(y2)−ϕλ(y1)
y2−y1
=: k, (4.25)
where kis a function w.r.t λ. Hence we have
ˆϕ0
λ(y) =
F−1
ρ(1 −y)−λ
α, y < y1,
k, y ∈[y1, y2],
F−1
ρ(1 −y), y > y2.
By using Theorem 3.1, the optimal solution to problem (4.20) is
Gλ,λ1(y) = (0,if λ1ˆϕ0
λ(y)> U0(ξ),
(U0)−1(λ1ˆϕ0
λ(y)),if λ1ˆϕ0
λ(y)≤U0(ξ).(4.26)
Note that the expected shortfall (ES) constraint implies
λ1k≤U0(ξ).
Set
g(c) = R1
cF−1
ρ(1 −y)dy
α−c.(4.27)
It is not hard to verify that g(c) is decreasing first and then increasing on the interval [0, α), thus
g(c) must admit a unique minimizer. Let
c∗= arg min
c∈[0,α)g(c),
R=g(c∗),
(4.28)
then we have the following result.
Lemma 4.2
(1) The problem (4.23)admits no optimal solution when x0< αRx1.
(2) If we treat λ, y2, k as the function of y1, then λ(·), y2(·)are continuously decreasing, while
k(·)is continuously increasing with respect to y1. Moreover, the reasonable range of y1is
[c∗, α]and the ranges of λ, y2, k are [0, αF −1
ρ(1 −c∗)],[α, 1],[0, F −1
ρ(1 −α)], respectively.
Proof. See Appendix C.2.
Let
f(λ, λ1) := 1
αZα
0
Gλ,λ1(y)dy,
g(λ, λ1) := Z1
0
Gλ,λ1(y)F−1
ρ(1 −y)dy.
For fixed x1, λ, let λ1=f1(λ, x1) be the solution of f(λ, λ1) = x1. Set ∆(x1) = g(0, f1(0, x1)).
Lemma 4.3 Assume 0< λ1k < U0(ξ).
(1) Both fand gare continuously increasing w.r.t λwhile continuously decreasing w.r.t λ1.
(2) f1(·, x1)is continuously increasing w.r.t. λ.
23
(3) g(λ, f1(λ, x1)) is a continuous function w.r.t. λ.
Proof. See Appendix C.2.
The main tool to check whether our approach works or not, is to actually solve the following
equation system:
f(λ, λ1) = x1,
g(λ, λ1) = x0.(4.29)
Using the continuity property in Lemma 4.3, one can deduce the following results.
Theorem 4.4 Suppose g(λ, λ1)has range R+for any λ1>0.
(1) If x0>∆(x1), then the ES constraint is redundant. Problem (4.23)must admit a unique
optimal solution.
(2) If x1≥ξ(1 −c∗
α), then for any x0∈[αRx1,∆(x1)], the equation system (4.29)admits at
least one solution (λ∗, λ∗
1)and the corresponding Gλ∗,λ∗
1(y)solves problem (4.23).
(3) If 0< x1< ξ(1 −c∗
α), then λ≤¯
λ, where λ=¯
λis the unique solution to (4.25)when taking
y1=α1−x1
ξ. Let
L(x1) = min
λ∈[0,¯
λ]g(λ, f1(λ, x1)),
–if x0=αRx1, then problem (4.23)admits a unique feasible solution which is also
optimal.
–if x0∈[L(x1),∆(x1)], then the equation system (4.29)admits at least one optimal
solution (λ∗, λ∗
1)and the corresponding Gλ∗,λ∗
1(y)solves problem (4.23).
–if x0∈(αRx1, L(x1)), then the equation system (4.29)admits no solution. The strong
duality does not hold, thus our method fails to solve problem (4.23).
Proof. See Appendix C.2.
4.3 Example 3: Yaari’s dual model
We consider the Yaari’s dual model in Yaari (1987), which was also solved by quantile formulation
method in Boudt et al. (2020); He and Jiang (2021); He and Zhou (2011b). The optimization
problem is formulated as
max
XTZ∞
0
ω(P(XT> x))dx (4.30)
subject to E[ρTXT] = x0,XT≥0 and XTis FTadapted. The structure of the problem under
quantile formulation is
V(x0) = sup
Q∈G Z1
0
Q(y)dy,
s.t. Z1
0
Q(y)ϕ0(y)dy =x0.
(4.31)
24
Then, we will follow our procedure to tackle this problem. Firstly, by using the Lagrange
multiplier methods, we consider the non-constraint problem:
V(λ, x0) := sup
Q∈G
J(Q) = Z1
0
(Q(y)−λϕ0(y)Q(y))dy +λx0.(4.32)
Secondly, replacing ϕby its concave envelope ˆϕ, the problem becomes
sup
Q∈G
J1(Q) = Z1
0
(Q(y)−λϕ0(y)Q(y))dy +λx0.(4.33)
Thus, the optimal solution to this problem is:
Q∗(y) =
0,if λˆϕ0(y)>1,
any non-decreasing and right continuous function,if λˆϕ0(y) = 1,
+∞,if λˆϕ0(y)<1.
Now, we first consider the special case when there exists 0 <c<1, such that
ˆϕ(y)> ϕ(y),∀y∈(c, 1); ˆϕ(y) = ϕ(y) otherwise. (4.34)
In fact, these conditions are the Assumptions 6,7 in He and Zhou (2011b). Let λ0=1−c
−ϕ(c), and
recalling Theorem 3.1, we can get the optimal solution to problem (4.32):
(1) if λ6=λ0, then
Q∗(y) = (0, λ ˆϕ0(y)>1,
+∞, λ ˆϕ0(y)≤1.
(2) if λ=λ0, then
Q∗(y) = (0, y < c,
c1, y ∈[c, 1),(4.35)
where c1is an arbitrary non-negative number.
Thus
V(λ, x0) = (+∞,if λ < λ0,
λx0,if λ≥λ0.
The weak duality implies:
V(x0)≤inf
λ>0V(λ, x0) = λ0x0.
By setting c1=x0
−ϕ(c), the optimal solution (4.35) satisfies the constraint in the prime problem,
which indicates that the strong duality holds, or equivalently,
V(x0) = inf
λ>0V(λ, x0).
Hence
Q∗(y) = x0
−ϕ(c)1{y≥c}(4.36)
is the optimal solution to problem (4.31).
The next theorem will discuss the existence conditions of the optimal solution to (4.31).
Theorem 4.5 Let k= limy↑1ˆϕ0(y), and c= inf{y∈(0,1) : ˆϕ0(y) = k}.
(1) If c= 1, then there is no optimal solution to problem (4.31).
(2) If c < 1and k > 0, then problem (4.31)admits an optimal solution. Moreover, if ˆϕ(y)>
ϕ(y),∀y∈(c, 1), the optimal solution is unique with the same form in (4.36).
Proof. See Appendix C.3
25
4.4 Example 4: Goal Reaching Model
In this example we consider the goal reaching model put forward in Browne (1999), see also
Cvitanic et al. (2019). The original form of the optimization problem is
max
XT
P(XT≥b),(4.37)
which is equivalent to
max
XT
E[1{XT≥b}] (4.38)
subject to E[ρTXT] = x0,XT≥0 and XTis FTadapted, where bis a constant target. When
we consider probability distortion, the objective function can be transformed into the following
form:
Z+∞
0
U(y)d(1 −w(1 −FXT(y))) = Z+∞
B
d(1 −w(1 −FXT(y))) = w(1 −FXT(B)) = w(Eu(XT)),
where U(y) = 1{y≥B}, and B > 0 is the goal of the agent. Since w(·) is strictly increasing, the
problem is equivalent to the optimization problem when there is no distortion function at all.
Accordingly, the non-distorted problem has the quantile formulation form (also see He and
Zhou (2011b)):
V(x0) := sup
G∈G Z1
0
1{G(x)≥B}dx
s.t. Z1
0
G(x)F−1
ρ(1 −x)dx =x0,
(4.39)
where ρ > 0 is a continuous random variable with distribution Fρ.
Using the Lagrange multiplier methods with parameter λ > 0, we get the unconstrained
problem:
V(λ, x0) := sup
G∈G Z1
0
(1{G(x)≥B}−λG(x)F−1
ρ(1 −x))dx +λx0.(4.40)
The concave envelope of U(x) = 1{x≥B}is
ˆ
U(x) = (x
B, x < B,
1, x ≥B.
Hence, we consider the problem :
sup
G∈G Z1
0
(ˆ
U(G(x)) −λG(x)F−1
ρ(1 −x))dx +λx0,
whose optimal solution can be easily obtained as
G∗(x) = (0, λF −1
ρ(1 −x)>1
B,
B, λF −1
ρ(1 −x)≤1
B.
And such a solution is also optimal to problem (4.40). Thus,
V(λ, x0) = Fρ1
λB +λ x0−BZ1
λB
0
ydFρ(y)!.
The first order derivative is
V0
λ(λ, x0) = x0−BZ1
λB
0
ydFρ(y) = x0−BE[ρ1{λ∗Bρ≤1}],
which is a strictly decreasing function w.r.t. λ. The equation
BE[ρ1{λ∗Bρ≤1}] = x0
admits a unique solution λ∗when 0 ≤x0≤BE[ρ]. If x0reaches the lower or upper bounds,
26
then we let λ∗be respectively +∞or 0+. Hence, if 0 < x0< BEρ, then
V(x0) = inf
λ>0V(λ, x0) = V(λ∗, x0),
or equivalently, the strong duality holds. Then
G∗(x) = B1{λ∗BF −1
ρ(1−x)≤1}
is the optimal solution to problem (4.39).
5 Solutions for the general model
Previously we demonstrate that our method can solve representative behavior finance portfolio
choice problems in the literature, under our proposed unified solution framework. In this section,
we shall characterize the solutions for the general case of Uand ϕ, i.e. when the utility function
exhibit a finite number of inflection points (reference points), and the probability distortion
function can also have multiple inflection points.
5.1 Notations
First we shall introduce some notations for convenience. For a monotone function f(x), let
f(x+) := lim
y↓xf(y)
f(x−) := lim
y↑xf(y)
be the right and left limit at point x.
Recalling Proposition 3.1 and Lemma 3.1, we shall pay more attention to the situation when
YDis not a singleton. Since Kis the jump points set of ( ˆ
U0)−1(·), for every ki∈ K, we can
find the largest open interval (ai, bi) on which ˆ
Uis affine with slope ki. We call this interval the
Affine Interval of slope ki. In fact
ai= ( ˆ
U0)−1(ki+), bi= ( ˆ
U0)−1(ki−).
Conveniently, we call Kthe Slopes Set of ˆ
Uon the domain R+.
To avoid complicated computations in the construction of the optimal solution, we separate
Kinto three disjoint subsets in the following way: for every slope k∈ K, and its corresponding
affine interval (a, b), let
K1:= k∈ K|(a, b)⊆ {x∈R+|ˆ
U(x)> U(x)},
K2:= k∈ K|(a, b)⊆ {x∈ R+|ˆ
U(x) = U(x)},
K3:= K \ (K1∪ K2).
For intuitive illustrations, please see Figure 11 and Figure 12.
Similarly, we can define Las the Slope Set of ˆϕon [0,1], the corresponding largest affine
interval (αi, βi) for any li∈ L and separate Linto L1,L2and L3. Set
Λ := k
l|k∈ K, l ∈ L,
Λ0:= λ∈Λ|there exist n > 1 and different ki∈ K and li∈ L with λ=ki
li
, i = 1,2,· · · , n.
27
Figure 11: An example for the situations ki∈ K1
and kj∈ K2.
Figure 12: An example for the situation k∈ K3.
The three sets K3,L3,Λ0will make the form of the optimal solutions complex, thus we shall
discuss them in the last subsection.
5.2 Optimal solutions under the case: K3=∅,L3=∅,Λ0=∅
In the case: K3=∅,L3=∅,Λ0=∅, we can easily get YBby using Theorem 3.1.
Proposition 5.2 Using the notations defined in Section 5.1, we have that the optimal solutions
to problem B are given by
(1) If λ /∈Λ, then
Q∗
λ(x)=(ˆ
U0)−1(λˆϕ0(x)−).(5.1)
(2) If λ∈Λ, then we can find the unique pair k∈ K, l ∈ L, such that λ=k
land their
corresponding affine intervals (a, b),(α, β).
(i) if k∈ K1and l∈ L1, then
Q∗
λ(x) = ((ˆ
U0)−1(λˆϕ0(x)−), x /∈[α, β)
aor b, x ∈[α, β)(5.2)
(ii) if k∈ K1and l∈ L2, then
Q∗
λ(x) = ((ˆ
U0)−1(λˆϕ0(x)−), x /∈[α, β)
aI[α,c1)+bI[c1,β), x ∈[α, β)(5.3)
where c1is a constant in [α, β].
(iii) if k∈ K2and l∈ L1, then
Q∗
λ(x) = ((ˆ
U0)−1(λˆϕ0(x)−), x /∈[α, β)
c2, x ∈[α, β)(5.4)
28
where c2is a constant in [a, b].
(iv) if k∈ K2and l∈ L2, then
Q∗
λ(x) = ((ˆ
U0)−1(λˆϕ0(x)−), x /∈[α, β)
h(x), x ∈[α, β)(5.5)
where h: [α, β)→[a, b]is an arbitrary non-decreasing and right-continuous function.
Assumption 5.3 There exists a λ0<+∞such that
λ0= inf λ > 0 : Z1
0
(ˆ
U0)−1(λˆϕ0(x)+)ϕ0(x)dx < +∞.
This assumption is intimately related to the well-posedness of the original optimization problem.
One should notice that if this assumption is not satisfied, the original problem has to be ill-posed
while this does not mean the well-posedness would be guaranteed if this assumption is satisfied.
The detailed discussion can be found in Appendix A.
For any λ /∈Λ, Q∗
λin (5.1) is unique, thus we can define
f(λ) := Z1
0
Q∗
λ(x)ϕ0(x)dx.
As a result, f(λ) is strictly decreasing due to the fact that ( ˆ
U0)−1is non-increasing and can not
be a constant function. Such a result forces us to only consider the case λ≥λ0. Accordingly,
the optimal Lagrange multiplier λ∗, if it exists, must be a decreasing function respect to x0. In
other words, we may discuss the two cases: λ=λ0,λ>λ0, or x0≥f(λ0+), x0< f(λ0+), to get
the sufficient and necessary conditions under which there exist a unique λ∗and a corresponding
Q∗
λsatisfying f(λ∗) = x0, i.e.
Z1
0
Q∗
λ(x)ϕ0(x)dx =x0.(5.6)
The case corresponding to x0=f(λ0+) <+∞is easy to solve by letting ( ˆ
U0)−1(λˆϕ0(x−)+) be
the optimal solution. Thus, in the following we shall consider the other cases.
Remark 5.5 If f(λ0+) = +∞, then the case for x0>+∞is redundant, and one only needs to
consider x0< f(λ0+). Thus, we assume that f(λ0+) <+∞in the next two theorems.
Set
k0:= lim
x→+∞
ˆ
U0(x),
l0:= lim
x→1−ˆϕ0(x),
¯
M:= ( ˆ
U0)−1(k0+),
¯c:= ( ˆϕ0)−1(l0+).
Theorem 5.6 Assume that x0> f (λ0+).
(1) If k0= 0,¯
M < +∞, then λ0= 0 and f(λ0+) = ¯
ME[ρ]. Moreover, any quantile function
Qwhich satisfies the constraint (5.6)and Q≥¯
Mis an optimal solution to Problem A.
(2) If k0= 0,¯
M= +∞, then
(I) if l0= 0 and ¯c < 1, then λ0= +∞, the problem is ill-posed.
29
(II) if l0= 0 and ¯c= 1, then the strong duality does not hold.
(III) if l0>0, then λ0= 0 and f(0+) = +∞.
(3) If k0>0,l0= 0, then λ0= +∞, or equally, the initial problem is ill-posed.
(4) If k0>0,l0>0, then λ0=k0
l0. Thus,
(I) if ¯
M= +∞or if ¯
M < +∞and ¯c= 1, then f(λ0+) <+∞, the strong duality does
not hold.
(II) if ¯
M < +∞and ¯c < 1, then the strong duality holds. Moreover, if l0∈ L1, then the
optimal solution is unique.
Proof. See Appendix D.1.
Let
Λ1=k
l> λ0|k∈ K1, l ∈ L1,
Λ2=k
l> λ0|k∈ K2, l ∈ L2.
(5.7)
Theorem 5.7 Assume that x0< f (λ0+).
(1) If x0∈Sλ∈Λ1(f(λ+), f (λ−)), then the strong duality (3.10)does not hold.
(2) If x0/∈Sλ∈Λ1(f(λ+), f(λ−)), then the strong duality (3.10)holds. There exists a λ∗and a
Q∗
λ∗solving (5.6), and such a function is the optimal solution to Problem A. Moreover,
if x0/∈Sλ∈Λ1∪Λ2(f(λ+), f (λ−)), then the optimal solution is unique.
Proof. See Appendix D.2.
Remark 5.6 For the case of x0∈Sλ∈Λ1(f(λ+), f (λ−)), the strong duality fails and the La-
grange method can not be directly applied to this case. To fully resolve this case is a future
research direction. A possible approach11 is to utilize the rationalization approach of Bernard
et al. (2015a) and combine it with results in Section 5 of Reichlin (2013).
Corollary 5.1 Assume that x0< f (λ0+).
If the utility function Uis concave, then Sλ∈Λ1(f(λ+), f (λ−)) is empty and the optimal solu-
tions to problem A are in the forms given in Proposition 5.2.
Moreover, if Uis strictly concave, then Sλ∈Λ2(f(λ+), f (λ−)) is empty and the optimal solution
to problem A is unique with the form (5.1), which is consistent with Xu (2016).
Corollary 5.2 Assume that x0< f (λ0+).
If ϕis concave, then Sλ∈Λ1(f(λ+), f (λ−)) is empty, thus, problem A admits optimal solution
in the forms given in Proposition 5.2.
Moreover if there is no probability distortion, then Sλ∈Λ1(f(λ+), f(λ−)) is empty too. Thus,
the problem A must admit an unique solution in the form (5.1)which is consistent with Reichlin
(2013).
11We thank Stephan Sturm for this suggestion based on private communication.
30
Remark 5.7 According to Xu (2016), a new pricing kernel ˜ρwith F−1
˜ρ(1 −x) = ˆϕ0(x)is given
with which the problem can be transformed into the following form:
sup
X
E[u(X)]
s.t. E[˜ρX] = x0.
(5.8)
From Reichlin (2013), only when ˜ρis atom-less12, this problem is equal to its concavified form
for all x0. In this paper, ˜ρmay contain an atom, but we only need to limit the value of x0out of
Sλ∈Λ1(f(λ+), f (λ−)) rather than Sλ∈Λ(f(λ+), f (λ−)).
Remark 5.8 For Yaari’s dual model, U(x) = x, then U0(x) = 1 and the general inverse (non-
increasing one) is given by
(U0)−1(x) =
+∞, x < 1,
∈[0,+∞), x = 1,
0, x > 1.
•If l0= 0, then problem A is ill-posed.
•If l0>0and ¯c= 1, then the strong duality does not hold. In fact, Theorem 4.5 tells us
that we cannot attain the optimum of problem A. If l0>0and ¯c < 1, then for any c < 1
with ϕ(c) = l0, the form (4.36)is the optimal solution to problem A.
Remark 5.9 For the goal reaching case, ˆ
U0(x) = 1
B,if x∈(0, B);ˆ
U0(x) = 0,if x > B.
Similarly, we can get
(ˆ
U0)−1(x) = B1{x< 1
B}.
(ˆ
U0)−1(0) = +∞,we get f(0−) = +∞,f(0+) = BE[ρ].Thus if x0≥f(0+),then the optimal
multiplier is λ∗= 0, and any quantile Q∗(x)≥Bsatisfying R1
0Q∗(x)ϕ0(x)dx = 0 is an optimal
solution to problem A. If x0≤f(0+), then problem A admits a unique solution.
5.3 Solutions when at least one of K3,L3,Λ0is not empty
We discuss this case in the following three sub-categories.
5.3.1 K3=∅,L3=∅,Λ06=∅
For the case K3=∅,L3=∅,Λ06=∅, there exists at least one λ∈Λ0, and two sequences ki, li
with λ=ki
liand corresponding affine intervals (ai, bi),(αi, βi), i = 1,2,· · · , n.
Now, for this special λ, using the results in (5.2), we can easily get
Q∗
λ+(x) = ((ˆ
U0)−1(λˆϕ0(x)−), x /∈S[αi, βi)
ai, x ∈[αi, βi)
Q∗
λ−(x) = ((ˆ
U0)−1(λˆϕ0(x)−), x /∈S[αi, βi)
bi, x ∈[αi, βi)
12We thank Stephan Sturm for pointing out that the general case of countable number of atoms can be handled
following similar steps as in Lemma 6.3 in Bichuch and Sturm (2014).
31
and
f(λ+) = Z1
0
Q∗
λ+(x)ϕ(x)dx,
f(λ−) = Z1
0
Q∗
λ−(x)ϕ(x)dx.
The strictly decreasing property of f(λ) implies that we can only consider x0∈[f(λ+), f(λ−)].
Let Iibe the set defined through:
1. If ki∈ K1, li∈ L1,Ii={0,(bi−ai)(ϕ(βi)−ϕ(αi))}.
2. Otherwise, Ii= [0,(bi−ai)(ϕ(βi)−ϕ(αi))].
Theorem 5.8 If x0∈[f(λ+), f (λ−)], then the strong duality (3.10)holds if and only if the
equation
10
nx=x0−f(λ+)
with the N-dimension variable x, admits at least one solution with x∗∈ ⊗n
i=1Ii.
Proof. The proof follows from Theorem 5.7 and its proof is given in Appendix D.2.
Now, if we get a solution x∗= (x∗
1,· · · , x∗
n), then λ∗=λand we can construct the optimal
solution Q∗(y) to problem A in each interval [αi, βi) with Rβi
αi(Q∗(x)−ai)ϕ0(x)dx =x∗
iin the
forms of Proposition 5.2.
5.3.2 K36=∅,L36=∅,Λ0=∅
Next, we turn to the situation K36=∅,L36=∅,Λ0=∅. Similarly, we only need to consider the
case: there exists one λ, with λ=k
l, and the corresponding (a, b),(α, β) can be written as:
[a, b) =
n
[
i=1
[ai, bi),
[α, β) =
m
[
i=1
[αi, βi),
where
ai=bi−1, a1=a, bn=n, ˆ
U(ai) = U(ai), i = 2,3,· · · , n,
αi=βi−1, α1=α, βm=β, ˆϕ(αi) = ϕ(αi), i = 2,3,· · · , m,
and the sub-intervals (ai, bi) belongs to either {ˆ
U=U}or {ˆ
U > U }, and the sub-interval (αj, βj)
belongs to either {ˆϕ=ϕ}or {ˆϕ>ϕ}.
Set
H:= {x∈[a, b] : ˆ
U(x) = U(x)},
and a sequence of non-deceasing and right-continuous functions {hi}, i = 1,2,· · · , m:
hi: [αi, βi)→ H
satisfying
hi(βi−)≤hi+1(αi+1),
hiis constant if (αi, βi)⊆ { ˆϕ>ϕ}.
We denote
xi:= Zβi
αi
(hi(x)−a)ϕ0(x)dx.
32
Now, we obtain the following result.
Theorem 5.9 If λ, H,{hi}are defined above, then the optimal solution to problem B can be
written as:
Q∗
λ(x) = Q∗
λ+(x) +
m
X
i=1
(hi(x)−a)1[αi,βi).
If x0∈[f(λ+), f (λ−)], then the strong duality (3.10)holds if and only if there exists such a
function sequence {hi}and x= (x1, x2,· · · , xm)defined above solving:
10
mx=x0−f(λ+).
Proof. We utilize Theorem 3.1 and the proofs in Theorem 5.7 to obtain the results.
5.3.3 K36=∅,L36=∅,Λ06=∅
Finally, if K36=∅,L36=∅,Λ06=∅, then for λ∈Λ0and x0∈[f(λ+), f (λ−)], we can combine
the solution construction of the former two cases and Theorem 5.8, Theorem 5.9 to solve this
problem.
6 Conclusion
In this paper, we propose to combine the relaxation method with the Lagrange method, and
provide a general solution method to the optimal investment problem with the presence of both
non-concave utility functions and nonlinear probability distortion. We obtain explicit closed-
form solutions in several representative behavior finance models. In particular, we obtain a
“distorted Merton line” in a special case of the CPT model. An interesting finding is that only
the probability distortion, but not the non-concave utility function, increases the proportion of
wealth invested in the risky stock, as compared to the standard Merton model.
Note that we only require that the market is complete, i.e. there exists a unique state-price
density or pricing kernel, and do not restrict the underlying asset price dynamics. Thus the
method is general enough and can be extended to other stochastic processes beyond the Black-
Scholes model. We leave the investigations of behavior portfolio choice for alternative stochastic
models as future research directions. In addition, the extension to possibly incomplete markets
is of great interest and left as a future research direction.
References
Armstrong, J. and D. Brigo (2019). Risk managing tail-risk seekers: Var and expected shortfall vs s-shaped utility.
Journal of Banking & Finance 101, 122–135.
Barberis, N. and M. Huang (2008). Stocks as lotteries: The implications of probability weighting for security
prices. American Economic Review 98(5), 2066–2100.
Basak, S. and A. Shapiro (2001). Value-at-risk-based risk management: optimal policies and asset prices. The
Review of Financial Studies 14(2), 371–405.
Bensoussan, A., A. Cadenillas, and H. K. Koo (2015). Entrepreneurial decisions on effort and project with a
nonconcave objective function. Mathematics of Operations Research 40(4), 902–914.
33
Berkelaar, A. B., R. Kouwenberg, and T. Post (2004). Optimal portfolio choice under loss aversion. Review of
Economics and Statistics 86(4), 973–987.
Bernard, C., P. P. Boyle, and S. Vanduffel (2014a). Explicit representation of cost-efficient strategies. Finance
35(2), 5–55.
Bernard, C., J. S. Chen, and S. Vanduffel (2015a). Rationalizing investors choices. Journal of Mathematical
Economics 59, 10–23.
Bernard, C., F. Moraux, L. R¨uschendorf, and S. Vanduffel (2015b). Optimal payoffs under state-dependent
constraints. Quantitative Finance 15, 1157–1173.
Bernard, C., L. R¨uschendorf, and S. Vanduffel (2014b). Optimal claims with fixed payoff structure. Journal of
Applied Probability 51(A), 175–188.
Bichuch, M. and S. Sturm (2014). Portfolio optimization under convex incentive schemes. Finance and Stochastics
18(4), 873–915.
Boudt, K., K. Dragun, Q. Tang, and S. Vanduffel (2020). The optimal payoff for a yaari investor. Available at
SSRN 3751644 .
Browne, S. (1999). Reaching goals by a deadline: Digital options and continuous-time active portfolio management.
Advances in Applied Probability 31(2), 551–577.
Browne, S. (2000). Risk-constrained dynamic active portfolio management. Management Science 46(9), 1188–
1199.
Burgert, C. and L. R¨uschendorf (2006). On the optimal risk allocation problem. Statistics and Decisions-
International Journal Stochastic Methods and Models 24(1), 153–172.
Carpenter, J. (2000). Does option compensation increase managerial risk appetite? The journal of finance 55(5),
2311–2331.
Cvitanic, J., S. Kou, X. Wan, and K. Williams (2019). Pi portfolio management: Reaching goals while avoiding
drawdowns. Available at SSRN 3444836 .
Dai, M., S. Kou, S. Qian, and X. Wan (2019). Non-concave utility maximization without the concavification
principle. Available at SSRN 3422276 .
He, X. and Z. Jiang (2021). Optimal payoff under the generalized dual theory of choice. Operations Research
Letters 49(3), 372–376.
He, X. and S. Kou (2018). Profit sharing in hedge funds. Mathematical Finance 28(1), 50–81.
He, X., R. Kouwenberg, and X. Zhou (2017). Rank-dependent utility and risk taking in complete markets. SIAM
Journal on Financial Mathematics 8(1), 214–239.
He, X. and X. Zhou (2011a). Portfolio choice under cumulative prospect theory: An analytical treatment.
Management Science 57(2), 315–331.
He, X. and X. Zhou (2011b). Portfolio choice via quantiles. Mathematical Finance: An International Journal of
Mathematics, Statistics and Financial Economics 21(2), 203–231.
He, X. and X. Zhou (2016). Hope, fear, and aspirations. Mathematical Finance 26(1), 3–50.
Jin, H., Z. Xu, and X. Zhou (2008). A convex stochastic optimization problem arising from portfolio selection.
Mathematical Finance 18(1), 171–183.
34
Jin, H. and X. Zhou (2008). Behavioral portfolio selection in continuous time. Mathematical Finance: An
International Journal of Mathematics, Statistics and Financial Economics 18(3), 385–426.
Kahneman, D. and A. Tversky (1979). Prospect theory: An analysis of decision under risk. Econometrica 47(2),
263–292.
Merton, R. C. (1969). Lifetime portfolio selection under uncertainty: The continuous-time case. The Review of
Economics and Statistics pp. 247–257.
Polkovnichenko, V. (2005). Household portfolio diversification: A case for rank-dependent preferences. The
Review of Financial Studies 18(4), 1467–1502.
Quiggin, J. (1982). A theory of anticipated utility. Journal of Economic Behavior & Organization 3(4), 323–343.
Quiggin, J. (2012). Generalized expected utility theory: The rank-dependent model. Springer Science & Business
Media.
Reichlin, C. (2013). Utility maximization with a given pricing measure when the utility is not necessarily concave.
Mathematics and Financial Economics 7(4), 531–556.
R¨uschendorf, L. (1983). Solution of a statistical optimization problem by rearrangement methods. Metrika 30(1),
55–61.
R¨uschendorf, L. and S. Vanduffel (2020). On the construction of optimal payoffs. Decisions in Economics and
Finance 43, 129–153.
Tversky, A. and C. R. Fox (1995). Weighing risk and uncertainty. Psychological review 102(2), 269.
Tversky, A. and D. Kahneman (1992). Advances in prospect theory: Cumulative representation of uncertainty.
Journal of Risk and uncertainty 5(4), 297–323.
Van Bilsen, S. and R. J. Laeven (2020). Dynamic consumption and portfolio choice under prospect theory.
Insurance: Mathematics and Economics 91, 224–237.
Wei, P. (2018). Risk management with weighted var. Mathematical Finance 28(4), 1020–1060.
Xia, J. and X. Zhou (2016). Arrow–Debreu equilibria for rank-dependent utilities. Mathematical Finance 26(3),
558–588.
Xu, Z. (2016). A note on the quantile formulation. Mathematical Finance 26(3), 589–601.
Yaari, M. E. (1987). The dual theory of choice under risk. Econometrica 55(1), 95–115.
35
Online Supplementary Appendix
A Well-posedness
We study the well-posedness problem in the following three cases:
1. There exists one C∈R, such that U(x)< C, ∀x∈[0,∞). This type of utilities include the
utility function arising from the goal reaching problem, exponential utility, etc.
2. limx→∞ U(x)
x=k > 0. This type of utilities include the linear utility function, or the
asymptotic linear utility function.
3. limx→∞ U(x)
x= 0. This type of utilities include the CRRA utilities, etc.
Case 1: In this case, U(X)< C means E[ˆ
U(X)] < C, so the problem is well-posed.
Case 2: In this case, we give a crucial lemma which is a simple extension of the corresponding
result in He and Zhou (2011b).
Lemma A.1 Problem (2.10)is well-posed if lim inf y→1ϕ0(y)>0and ill-posed if lim supy→1ϕ0(y) =
0.
Proof. If limx→∞ U(x)
x=k > 0, then for ∀ > 0 there exist a N∈R+such that (k−)x <
U(x)<(k+)xfor ∀x>N.
On one hand, if lim supt→1ϕ0(t) = 0, then for ∀n > 0, there exist y1∈(0,1) such that
ϕ0(y)<1
nfor ∀y∈(y1,1).
Let
Q(y) = (δ, y < y2,
b, y ≥y2,
where 0 < y2<1, and δis a constant satisfying x0+δϕ(0) >0 and U(δ)>−∞, and bis a
constant to be determined. Let b=x0−δ(ϕ(y2)−ϕ(0))
ϕ(1)−ϕ(y2), then Q(y) satisfies the constraint in (2.10),
thus Q(y) is a feasible solution. For y2≥max{ϕ−1(−x0+δϕ(0)
N−δ), y1}, we have b > N , and
Z1
0
U(Q(y))dy =U(δ)y2+U(b)(1 −y2)
>min{U(δ),0}+ (k−)x0+δϕ(0) −δϕ(y2)
ϕ(1) −ϕ(y2)(1 −y2)
>min{U(δ),0}+ (k−)(x0+δϕ(0)) 1−y2
−ϕ(y2)
≥min{U(δ),0}+ (k−)(x0+δϕ(0))n.
Thus supQR1
0U(Q(y))dy → ∞, when n→ ∞.This shows that the problem (2.10) is ill-posed.
On the other hand, if lim infy→1ϕ0(y)>0, then there exist land y3∈(0,1) such that ϕ0(y)> l
for ∀y∈[y3,1]. Thus, for ∀feasible Q(y), there exists a y4∈[0,1], such that Q(y)≥N+ 1,
36
∀y > y4(If Q(y)< N + 1 for ∀y∈[0,1], then we take y4= 1). Let y0=max{y3, y4}, then we
have
Z1
0
U(Q(y))dy =Zy0
0
U(Q(y))dy +Z1
y0
U(Q(y))dy
< U(Q(y0))y0+k+
lZ1
y0
Q(y)ϕ0(y)dy
≤U(Q(y0))y0+k+
lx0
<∞.
So, problem (2.10) is well-posed.
Case 3: Rewriting problem (2.10) as
VA(x0) := sup
Q∈G Z1
0
U(Q(y))dy,
s.t. Z1
0
Q(y)ϕ0(y)dy =x0,
(A.1)
and replacing Uby its concave envelope ˆ
U, we get
ˆ
V(x0) := sup
Q∈G Z1
0
ˆ
U(Q(y))dy,
s.t. Z1
0
Q(y)ϕ0(y)dy =x0.
(A.2)
Let M0= inf{x∈[0, M ] : ˆ
U(x)> U(x)}, and here Mis the one mentioned in Assumption 2.1,
then
ˆ
U(x)−(U(M)−U(M0)) ≤U(x)≤ˆ
U(x),∀x≥0,
i.e.
ˆ
V(x0)−(U(M)−U(M0))Eρ≤VA(x0)≤ˆ
V(x0).
Consequently, we only need to consider the well-posedness problem of problem (A.2).
Let ξ= ˆϕ0(1 −Fρ(ρ)). Problem (A.2) is equivalent to
ˆ
V(x0) = sup
X
E[ˆ
U(X)],
s.t. E[ξX ] = x0.
(A.3)
Thus, the well-posedness of problems (A.1) and (A.3) are equivalent.
The solution to problem (A.3) is
X∗= ( ˆ
U0)−1(λξ),(A.4)
where the inverse refers to the general inverse and λ > 0 is a constant satisfying E[ξX∗] = x0.
Taking the above analysis into account, the following two propositions hold true in our setup.
Proposition A.1 (Theorem 3.1 in Jin et al. (2008)) If E[( ˆ
U0)−1(λξ)ξ] = +∞for any λ > 0,
then V(x0)=+∞for ∀x0>0.
37
Here, this proposition leads to the necessity of Assumption 5.3 in Section 5.
Suppose that F(·) is the cumulative distribution function of ξ, then we have
Proposition A.2 (Theorem 5.4 in Jin et al. (2008))
1. If lim infx→+∞−xˆ
U00(x)
ˆ
U0(x)>0, then the problem (A.3)is well-posed for ∀x0>0if and
only if E[ˆ
U(( ˆ
U0)−1(ξ))] <∞.
2. If lim supx→0+−xF 0(x)
F(x)<0, then the problem is well-posed for any x0>0if and only if
E[( ˆ
U0)−1(λξ)ξ]<+∞for some λ > 0.
B Proofs in Section 3.3
B.1 Proof of Theorem 3.1
Proof.
(1) First, we claim that for every Q∈ Q and λ > 0,
JD(Q, λ)≥JC(Q, λ),(B.5)
and moreover, the equality holds if and only if Qsatisfies (3.8).
In fact, recalling that ˆϕ(0) = ϕ(0),ˆϕ(1) = ϕ(1), then we have
JD(Q, λ)−JC(Q, λ) = λZ1
0
(ϕ0(x)−ˆϕ0(x))Q(x)dx
=λZ1
0
(ϕ0(x)−ˆϕ0(x))Q(x)dx −Q(0)λZ1
0
(ϕ0(x)−ˆϕ0(x))dx
=λZ1
0
(ϕ0(x)−ˆϕ0(x)) Zx
0
dQ(y)dx
=λZ1
0
( ˆϕ(x)−ϕ(x))dQ(x) (B.6)
≥0.
Thus, JD(Q, λ)≥JC(Q, λ) for any Q.
By using the continuity of ˆϕ−ϕ, we know that the set {x∈[0,1] : ˆϕ(x)> ϕ(x)}is an open
set, thus it can be written as the union of countable disjoint open intervals, i.e.
{x∈[0,1] : ˆϕ(x)> ϕ(x)}=
+∞
[
i=1
Ii,
where Iiis an open interval. Consequently, (B.6) can be rewritten as
JD(Q, λ)−JC(Q, λ) = λ
+∞
X
i=1 ZIi
( ˆϕ(x)−ϕ(x))dQ(x)
Recalling that ˆϕ(x)−ϕ(x)>0,∀x∈Ii, it is easy to show that JD(Q, λ) = JC(Q, λ) is equivalent
to ZIi
dQ(x) = 0, i = 1,2,· · · ,
38
or
Q(·) is a constant function on each Ii.
Now, for every interval (α, β)⊆ {x: ˆϕ(x)> ϕ(x)}, there must exist an i, such that (α, β)⊆Ii.
We then complete our proof of the claim.
Next, we prove the part 1 of the theorem. By taking supremum over (B.5), we have
VD(λ, x0)≥VC(λ, x0).
“If Part”: For any Q∈ YDsatisfying (3.8), the former claim implies
VD(λ, x0) = JD(Q, λ) = JC(Q, λ)≤VC(λ, x0).
We then get VC(λ, x0) = VD(λ, x0) immediately and Qis an optimal solution to problem C.
“Only If Part”: First, let
Q1(x) = ( ˆ
U0)−1
−(λˆϕ0(x)).(B.7)
Obviously, Q1(x)∈ YDsatisfies (3.8), which indicates that VC(λ, x0) = VD(λ, x0).
Now, if Q∈ YDis optimal to problem C, then Qmust satisfy (3.8), or otherwise we have
VD(λ, x0) = VC(λ, x0) = JC(Q, λ)< JD(Q, λ) = VD(λ, x0),
which leads to a contradiction.
(2) On one hand, if
U(Q(x)) = ˆ
U(Q(x)),∀y∈[0,1],
then
VB(λ, x0)≤VC(λ, x0) = JC(Q, λ)
=Z1
0
(ˆ
U(Q(y)) −λϕ0(y)Q(y))dy +λx0
=Z1
0
(U(Q(y)) −λϕ0(y)Q(y))dy +λx0
≤VB(λ, x0)
which shows that Q(·) must be the optimal solution to problem B.
On the other hand, it is easy to check that Q1defined in (B.7) satisfies:
U(Q1(x)) = ˆ
U(Q1(x)),∀y∈[0,1].
Thus we have VC(λ, x0) = VB(λ, x0). Now, if Q(·)∈ YCis also optimal to problem B, but there
exists a point y0such that ˆ
U(Q(y0)) > U(Q(y0)), then the right-continuity of U, ˆ
U, Q implies
that there exist > 0, δ > 0 such that
ˆ
U(Q(y)) −U(Q(y)) > , ∀y∈[y0, y0+δ).
39
Hence
VC(λ, x0)−VB(λ, x0) = JC(Q, λ)−JB(Q, λ)
=Z1
0
(ˆ
U(Q(y)) −U(Q(y)))dy
≥Zy0+δ
y0
(ˆ
U(Q(y)) −U(Q(y)))dy
> δ.
We arrive at a contradiction.
In conclusion, Q1is a common solution to problems D, C, B and we have VD(λ, x0) = VC(λ, x0) =
VB(λ, x0). This completes the proof.
C Proofs in Section 4
C.1 Proof of Theorem4.3
Proof.
Because of the same solution structure in the two cases in this theorem, we only give the proof
to the case 1.
If x0≤xL, Theorem 4.2 implies that f(λ) = x0admits a unique root λ∗≥λ0, then λ∗ˆϕ0(y)≥
U0(ξ)∀y∈[0, c).Thus, Q∗(y) = 0 for y < c,and
ϕ0(y) = k−1
1exp ˜µ−a1˜σ−a2
1/2+(a1+ ˜σ)Φ−11−y
k1, y ≥c,
which shows that λ∗ˆϕ0(y) = U0(ξ), y ≥chas a unique solution
y∗= 1 −k1Φ ln k1u0(ξ)
λ∗+a1˜σ+a2
1/2−˜µ
a1+ ˜σ!.
Hence, we can obtain the explicit expression:
X∗
T=Q∗(1 −w(Fρ(ρ))) = (U0)−1(λ∗ϕ0(1 −w(Fρ(ρ))))1{1−w(Fρ(ρ))≥y∗}.
Recalling that for x≤b,
ϕ0(x) = F−1
ρ(w−1(1 −x))
w0(w−1(1 −x)) , w0(x) = k1exp{−a2
1/2−a1Φ−1(x)}.
Thus, recalling the notations c1, c2and c3in (4.14), we have
X∗
T= (U0)−1λ∗ρ
w0(Fρ(ρ))1{ρ≤c1}
= η+γw0(Fρ(ρ))
λ∗ρ1
1−γ!1{ρ≤c1}
= η+k1γ
λ∗1
1−γ
exp −1
1−γa2
1/2−a1˜µ
˜σρ−1
1−γ(1+ a1
˜σ)!1{ρ≤c1}
= (η+c2ρ−c3)1{ρ≤c1}.
40
Let
Zt:= ρ
ρt
,
then Ztis independent of Ftwith distribution function Φ ln x−At
Bt, where At, Btare defined in
(4.11). As a result, we have
X∗
t=1
ρt
E[ρX∗
T|F] = ηE[Zt1{Zt≤c1/ρt}] + c2ρ−c3
tE[Z1−c3
t1{Zt≤c1/ρt}].
Recall the fact: if Y∼N(0,1), then E[eαY 1{Y≤β}] = eα2/2Φ(β−α). Then we can easily compute
X∗
t=ηexp{B2
t+A2
t}Φln c1/ρt−At
Bt
−Bt+c2exp{(1 −c3)At
+ (1 −c3)2B2
t/2}ρ−c3Φln c1/ρt−At
Bt
−(1 −c3)Bt.
If we fix t, then X∗
tis a decreasing function respect to ρt, which we note as f(ρt). We have
π∗
t=−µt−rt
σ2
tX∗(t)ρtf0(ρt)
=µt−rt
σ2
tX∗(t)η
Bt
exp{B2
t/2 + At}φln c1/ρt−At
Bt
−Bt
+c2c3exp{(1 −c3)At+ (1 −c3)2B2
t/2}ρ−c3Φln c1/ρt−At
Bt
−(1 −c3)Bt
+c2
Bt
exp{(1 −c3)At+ (1 −c3)2B2
t/2}ρ−c3φln c1/ρt−At
Bt
−(1 −c3)Bt.
This completes the proof.
C.2 Proof of Lemma 4.2,4.3 and Theorem 4.4
Proof of Lemma 4.2:
(1) We only need to consider the following minimization problem:
inf
G(·)Z1
0
G(y)F−1
ρ(1 −y)dy
s.t. 1
αZα
0
G(y)dy =x1,
(C.8)
to judge whether the feasible set is empty or not.
Since G(y) is non-decreasing in y, the optimal solution to (C.8) must satisfy:
G(y) = G(α),∀y∈[α, 1],
then we can solve (C.8) in two steps.
Firstly, we fix G(α), and solve
inf
G(·)≤G(α)Zα
0
G(y)F−1
ρ(1 −y)dy
s.t. 1
αZ1
0
G(y)dy =x1.
41
By using the Lagrange multiplier method, it turns out that
inf
G(·)≤G(α)Zα
0
G(y)(F−1
ρ(1 −y)−λ
α)dy
admits an optimal solution
G∗(y) = (0, y < c,
G(α), y ∈[c, α),
where c= 1 −Fρ(λ
α). Substituting the solution into the constraint, we obtain:
G(α) = αx1
α−c.
Secondly, we solve the initial problem (C.8) by minimizing:
inf
c∈[0,α)αx1R1
cF−1
ρ(1 −y)dy
α−c.
Recalling the definition in (4.28), we get the optimal value αRx1, and the optimal minimizer c∗.
Thus, x0≥αRx1holds, or equivalently the problem (4.23) admits no solution when x0< αRx1.
(2) Recalling the definition in (4.25), we have:
ϕλ(y2)−ϕλ(y1) = k(y2−y1).
Applying a total differential on each side, we obtain
ϕ0
λ(y2)dy2−ϕ0
λ(y1)dy1−(1 −y1
α)dλ =k(dy2−dy1)+(y2−y1)dk.
Thus dk
dλ =−1−y1
α
y2−y1
<0,if λ > 0,
and d(k+λ
α)
dλ =dk
dλ +1
α=
y2
α−1
y2−y1
>0,if λ > 0.
Since (4.24) and (4.25) imply
y1= 1 −Fρk+λ
α,
y2= 1 −Fρ(k),
we can easily conclude that both kand y1are strictly decreasing, while y2is strictly increasing
w.r.t λ. In other words, λ, y2is strictly decreasing w.r.t. y1, while kis increasing w.r.t. y1.
Moreover, the continuity property is obvious since all the functions are differentiable.
Now, we want to find the ranges for the above variables y1, y2, k, λ.λ≥0 is needed because
of the constraint of ES. One critical case is λ= 0, then y1=y2=α, k =F−1
ρ(1 −α).
The other condition ˆϕ0
λ(y)≥0 is needed in solving problem (4.21) too, otherwise the optimal
solution is G(y)=+∞. Now, we find the other critical case: there exists y < 1, s.t. ˆϕ0
λ(y) = 0.
Thus, k= 0, y2= 1 and
F−1
ρ(1 −y1) = λ
α,
Z1
y1
F−1
ρ(1 −s)ds =λ1−y1
α,
(C.9)
42
by the definitions in (4.24) and (4.25). Combining the above two equation, we have
g(y1) = F−1
ρ(1 −y1).
By comparing this equation with g0(c∗) = 0, we have that y1=c∗is the root to (C.9).
Consequently, we obtain the desired results: y1∈[c∗, α], y2∈[α, 1], λ ∈[0, αF −1
ρ(1 −c∗)] and
k∈[0, F −1
ρ(1 −α)]. This completes the proof.
Proof of Lemma 4.3:
(1) We only prove this property for function f. The monotonic side is easy to check, since
(U0)−1(·) is a non-increasing function.
For any λ, λ1satisfying λ1k < U 0(ξ), let ˆϕ0
λ(y0) = U0(ξ)
λ1i.e.
y0= 1 −Fρλ
α+U0(ξ)
λ1,(C.10)
then y0is continuous w.r.t. λor λ1and
f(λ, λ1) = 1
αZα
y0
(U0)−1(λ1ˆϕ0
λ(y))dy
=1
αZy1
y0
(U0)−1λ1F−1
ρ(1 −y)−λλ1
αdy + (U0)−1(λ1k)1−y1
α.
Here (U0)−1is continuous on (0, U0(ξ)) by the definition of U,y1is continuous w.r.t. λand y0is
continuously w.r.t. λor λ1, so it is easy to get the continuity results.
Moreover, the statements (2) (3) can be easily deduced from the statement (1). This completes
the proof.
Proof of the Theorem 4.4:
(1) If x0>∆(x1), then we can remove the ES constraint and obtain the optimal solution to
problem (4.23), i.e. let λ= 0 and we can solve g(0, λ1) = x0to get the root λ∗
1and the optimal
solution G∗
0,λ∗
1(y).
By Lemma 4.3,x0>∆(x1) is equivalent to λ∗
1< f1(0, x1), or equivalently, f(0, λ1)> x1.
Hence the ES constraint holds, which means that G∗
0,λ∗
1(y) is optimal.
(2) It is easy to check that:
g(c∗, f1(c∗, x1)) = αRx1, g(0, f1(0, x1)) = ∆(x1).
When x1≥ξ(1 −c∗
α), f1(·, x1) is well-defined on [0, c∗], thus by Lemma 4.3, when x0∈
[αRx1,∆(x1)], x0=g(λ, f1(λ, x1)) must admit at least one solution λ∗and (λ∗, f1(λ∗, x1)) is
a root to the equation system (4.29). Thus problem (4.23) must admit an optimal solution.
(3) By definition of y1, for any solution in (4.26), we have G(y1)≥ξ, thus
x1=1
αZα
0
Gλ,λ1(y)dy ≥1
αξ(α−y1).
Now, if 0 < x1< ξ(1 −c∗
α), then y1≥α(1 −x1
ξ)> c∗, which implies that λmust be restricted
in [0,¯
λ].
43
By Lemma 4.3,g(·, f1(·, x1)) must admit its maximum and its minimum, since it is continuous
on its domain [0,¯
λ]. Obviously, its maximum is ∆(x1). Now, set
L(x1) = min
λ∈[0,¯
λ]g(λ, f1(λ, x1)),
then the range of g(·, f1(·, x1)) is [L(x1),∆(x1)]. Accordingly, g(λ, f1(λ, x1)) = x0admits at least
one root when x0∈[L(x1),∆(x1)] and admits no root when x0∈(αRx1, L(x1)).
In conclusion,
•if x0=αRx1, then problem (4.23) admits a unique feasible solution which is also optimal.
•if x0∈x0∈[L(x1),∆(x1)], then problem (4.23) admits at least one optimal solution in the
form (4.26).
•if x0∈(αRx1, L(x1)), then the strong duality does not hold, and our method fails to find
the optimal solution.
This completes the proof.
C.3 Proof of Theorem 4.5
Before starting our proof, we first introduce a useful result. Fix x0, if we consider the function ϕ
in problem (4.31) as a parameter, and denote the optimal value function as W(ϕ), then we have
the following result.
Proposition C.3 If two increasing and absolutely continuous functions f1, f2satisfy: f1≤f2
and f1(0) = f2(0), f1(1) = f2(1), then
W(f1)≤W(f2).
Moreover if Q(y)is an optimal solution to problem (4.31)with the parameter f2and Q(y)is
constant on any open interval which is a subset of {y:f1(y)< f2(y)}, then Q(y)is an optimal
solution to problem (4.31)with the parameter f1and
U(f1) = U(f2).
Proof.
We let Q(y) solve the problem (4.31) with the parameter f1. By noticing that
Z1
0
Q(y)(f0
2(y)−f0
1(y))dy =Z1
0Zy
0
dQ(z)(f0
2(y)−f0
2(y))dy =Z1
0
(f1(y)−f2(y))dQ(y)≤0,
and we denote the above value as −, where ≥0. Now, let
Q1(y) = Q(y) +
−ϕ(0),
then Q1(y) is feasible to problem (4.31) with the parameter f2, and
W(f1) = Z1
0
Q(y)dy ≤Z1
0
Q1(y)dy ≤W(f2).
44
Let = 0, the equality term is obvious. Thus we complete the proof.
Proof of Theorem 4.5:By Proposition C.3,W(ϕ)≤W( ˆϕ). If k= 0, then the case 2 in
Appendix A shows that our problem is ill-posed, and the result holds naturally. Hence we just
consider k > 0.
For any feasible solution Q(y) to problem (4.31) with parameter ˆϕ,
x0=Z1
0
Q(y) ˆϕ0(y)dy ≥kZ1
0
Q(y)dy.
By taking supreme of the right hand, we get
W( ˆϕ)≤x0
k.
Thus
W(ϕ)≤x0
k.
Now, for any strictly increasing sequence {cn},cn↑1, define
fn(y) = (ˆϕ(y), y < cn,
−ˆϕ(cn)
1−cn(y−cn) + ˆϕ(cn), y ≥cn.
Then,
fn↑ˆϕ, W (fn)↑W( ˆϕ).
By using the conclusion in the former special case, we have
W(fn) = 1−cn
−ϕ(cn)x0↑x0
k
with optimal solution
Qn(y) = x0
−ϕ(cn).
Hence we get
W( ˆϕ) = x0
k.
(1) If c= 1, then ˆϕis not affine on the neighborhood of point 1, then there exists a δ > 0 such
that ˆϕ(y) = ϕ(y),∀y∈(1 −δ, 1). Consequently, there exists a N∈N∗, such that cn>1−δfor
any n > N, which indicates that Qnis also a feasible solution to problem (4.31) with parameter
ϕ. So we must have
W( ˆϕ) = lim
n→+∞W(fn)≤W(ϕ),
which leads to
W(ϕ) = x0
k.
Now, we will show that this value is unattainable.
For any feasible Qto the initial problem (4.31), there must exist an a∈(1 −δ, 1) with
Q(a) = b > 0,Ra
0Q(y)dy =: I1> 1and ˆϕ0(a)> k +2. Let I2:= R1
aQ(y)dy, then
x0=Z1
0
Q(y)ϕ0(y)dy ≥Z1
0
Q(y) ˆϕ0(y)dy
=Za
0
Q(y) ˆϕ0(y)dy +Z1
a
Q(y) ˆϕ0(y)dy
≥ˆϕ0(a)I1+kI2.
45
Hence, we have
x0
k≥I2+ˆϕ0(a)
kI1=I1+I2+ˆϕ0(a)−k
kI1>Z1
0
Q(y)dy +12
k,
which implies that the maximum x0
kcannot be attained.
(2) c < 1, then ˆϕ(c) = ϕ(c), and
Q∗(y) = x0
−ϕ(c)1{y≥c}
is an optimal solution which can reach its maximum x0
k. And if ˆϕ(y)> ϕ(y),∀y∈(c, 1), by the
conclusion in case (4.34), the optimal solution is unique. This completes the proof.
D Proofs in Section 5
D.1 Proof of Theorem 5.6
Proof.
(1) k0= 0,¯
M < +∞. For any λ > 0, ( ˆ
U0)−1(λˆϕ0(x)+) <¯
Mwhich indicates f(λ+) <+∞
for any λ > 0, we then get λ0= 0 immediately. Now, let λ= 0, then ( ˆ
U0)−1(λˆϕ0(x)+) = ¯
Mfor
every x∈[0,1], which implies f(λ0+) = ¯
ME[ρ].
The conditions on k0,¯
Mare equal to U(x) = U(¯
M),for every x∈[¯
M, +∞), which shows that
VA(x0)≤U(¯
M).
For any feasible Qwith Q(x)≥¯
Mand R1
0U(Q(x))dx =U(¯
M), Qis optimal to Problem A.
(2) k0= 0,¯
M= +∞. If l0= 0 and ¯c < 1, then, for any λ > 0, λˆϕ0(x)=0,∀x∈[¯c, 1]. As a
result,
(ˆ
U0)−1(λˆϕ0(x)+) = +∞,∀x∈[¯c, 1],
and
f(λ+)≥Z1
¯c
(ˆ
U0)−1(λˆϕ0(x)+)dx = +∞.
Thus, λ0= +∞, and Problem A is ill-posed.
If l0= 0 and ¯c= 1, then the exact value of λ0depends on the precise expressions of ˆ
Uand ˆϕ.
But the strong duality does not hold.
If l0>0, for any λ > 0, λˆϕ0(x)≥λl0, then
(ˆ
U0)−1(λˆϕ0(x)+) ≤(ˆ
U0)−1(λl0)<+∞.
Consequently,
f(λ+) ≤(ˆ
U0)−1(λl0)E[ρ]<+∞
for all λ > 0, which implies λ0= 0. Moreover, the conditions k0= 0 and ¯
M= +∞means
that, for any large n, there exist a δ > 0, such that ( ˆ
U0)−1(x)≥nif x∈(0, δ). Similarly,
there exists 0 < δ1<1 such that ˆϕ0(x)< l0+ 1 on (δ1,1). Now, if we choose λ < δ
l0+1 , we get
f(λ+) ≥(1 −δ1)n, hence we have f(0+) = +∞.
46
(3) k0>0, l0= 0. For any λ > 0, there exist 0 < δ < 1 such that ˆϕ0(x)<k0
λfor x∈(δ, 1),
hence
f(λ+) ≥Z1
δ
(ˆ
U0)−1(k0−)dx = +∞.
As a result, λ0= +∞, and Problem A is ill-posed.
(4) k0>0, l0>0, it is not hard to get λ0=k0
l0.
If ¯
M= +∞, and f(λ0+) <+∞, then there is no Q∗
λ∈ YBwhich solves the constraint (5.6),
thus the strong duality does not hold.
If ¯
M < +∞and ¯c= 1, then f(λ0+) ≤¯
ME[ρ]. For any x0> f (λ0+), there is no Q∗
λ∈ YB
solves the constraint (5.6), hence the strong duality does not hold.
If ¯
M < +∞and ¯c < 1, let
Q∗(x) = ((ˆ
U0)−1(λ0ˆϕ0(x)+), x < ¯c,
¯
M+x0−f(λ0+)
−ϕ(¯c), x ≥¯c,
then Q∗(x) solves the constraint (5.6). Moreover, if l0∈ L1, such Q∗can be determined uniquely.
This completes the proof.
D.2 Proof of Theorem 5.7
Proof. The proof can be divided into separate discussions for the following many cases that
exhaust all possibilities:
Case 1: If λ /∈Λ, firstly, we want to show that f(λ) is continuous at λ. Obviously, we can
find the explicit left or right limit of Q∗
λ. In fact, for sequence 1/n ↓0, Q∗
λ+1/n ↑(ˆ
U0)−1(λˆϕ0(·)+)
and Q∗
λ−1/n ↓(ˆ
U0)−1(λˆϕ0(·)−). Thus, by using the monotone convergence theorem, we have
lim
n→+∞f(λ+ 1/n) = f(λ+),
lim
n→+∞f(λ−1/n) = f(λ−).
Now, ( ˆ
U0)−1is jumped on K, which indicates that ( ˆ
U0)−1(λˆϕ0(·)−)>(ˆ
U0)−1(λˆϕ0(·)+) on {x:
λˆϕ0(x)∈ K},λ /∈Λ, and that {x:λˆϕ0(x)∈ K} is a countable set with measure 0. Hence we get
f(λ−)−f(λ+) = 0,
which implies that f(λ) is continuous at λ. Since Q∗
λis uniquely determined if λ /∈Λ, we get
that λ /∈Λ is equivalent to f(λ)/∈Sλ∈Λ0[f(λ+), f (λ−)].
Thus, for any x0/∈Sλ∈Λ[f(λ+), f (λ−)], we can find a unique λ∗, so that f(λ∗) = x0. Therefore,
Q∗
λ∗is the unique optimal solution to problem A.
Case 2: If λ∈Λ, then, we can find the corresponding k∈ K, l ∈ L and (a, b),(α, β). Recalling
(5.1), we can easily obtain:
Q∗
λ−(x) = ((ˆ
U0)−1(λˆϕ0(x)−), x /∈[α, β),
b, x ∈[α, β),
Q∗
λ+(x) = ((ˆ
U0)−1(λˆϕ0(x)−), x /∈[α, β),
a, x ∈[α, β).
47
Hence we have
f(λ−)−f(λ+) = Z1
0
(Q∗
λ−(x)−Q∗
λ+(x))ϕ0(x)dx
=Zβ
α
(b−a)ϕ0(x)dx
= (b−a)(ϕ(β)−ϕ(α))
>0.
Thus, f(λ) is jumped at λand λ∈Λ is equivalent to x0∈[f(λ+), f (λ−)]. We will discuss this
situation in four cases.
Case I: If k∈ K1, l ∈ L1, i.e. λ∈Λ1, then only when x0=f(λ+) or f(λ−), we can choose
Q∗
λin (5.2) which satisfies the constraint in (5.6), while x0∈(f(λ+), f(λ−)), the optimal λ∗is
not exist. Thus, the strong duality (3.10) does not hold.
Case II: If k∈ K1, l ∈ L2, then Q∗
λis given by (5.3), or equivalently,
Q∗
λ=Q∗
λ++ (b−a)1[c,β).
Thus,
Z1
0
Q∗
λ(x)ϕ0(x)dx =f(λ+) + (b−a)(ϕ(β)−ϕ(c)).
Let the right hand to be x0, then we must have
c=ϕ−1ϕ(β)−x0−f(λ+)
b−a.
Only in this situation, the optimal solution Q∗
λsatisfies the constraint (5.6). Thus, problem A
admits a unique solution.
Case III: If k∈ K2, l ∈ L1, thus Q∗
λis defined in (5.4), or equivalently,
Q∗
λ(x) = Q∗
λ++ (c−a)1[α,β).
Thus,
Z1
0
Q∗
λ(x)ϕ0(x)dx =f(λ+) + (c−a)(ϕ(β)−ϕ(α)).
Let the right hand be x0, then we can get
c=a+x0−f(λ+)
ϕ(β)−ϕ(α).
Only in this situation Q∗
λsatisfies the constraint (5.6) which gives the evidence that Problem
Aadmits a unique solution.
Case IV: If k∈ K2, l ∈ L2i.e. λ∈Λ2, then the optimal solution to problem B is described
in (5.5), or equivalently,
Q∗
λ(x) = Q∗
λ++ (h(x)−a)1[α,β).
Let h(x) defined in (5.5) satisfy:
Z1
0
(h(x)−a)ϕ0(x)dx =x0−f(λ+),
48
then Q∗
λsatisfies the constraint (5.6). This indicates that such Q∗
λis the optimal solution to
Problem A.
It is easy to show, only when x0=f(λ+) or f(λ−), h(x) is unique. Thus, if x0∈(f(λ+), f (λ−)),
then Problem A may admit infinite number of optimal solutions.
In conclusion, combining Case 1 and Case 2, we will get the desired results. This completes
the proof.
49