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

11institutetext: Department of Mathematical Sciences, Worcester Polytechnic Institute, 100 Institute Road, Worcester, MA 01609, USA; 11email: mbichuch@wpi.edu, ssturm@wpi.edu. Work partially supported by NSF grant DMS-0739195 while the authors were Postdoctoral Research Associates at Princeton University.

Portfolio Optimization under Convex Incentive Schemes

Maxim Bichuch    Stephan Sturm
(July 27, 2025)
Abstract

We consider the terminal wealth utility maximization problem from the point of view of a portfolio manager who is paid by an incentive scheme, which is given as a convex function gg of the terminal wealth. The manager’s own utility function UU is assumed to be smooth and strictly concave, however the resulting utility function UgU\circ g fails to be concave. As a consequence, the problem considered here does not fit into the classical portfolio optimization theory. Using duality theory, we prove wealth-independent existence and uniqueness of the optimal portfolio in general (incomplete) semimartingale markets as long as the unique optimizer of the dual problem has a continuous law. In many cases, this existence and uniqueness result is independent of the incentive scheme and depends only on the structure of the set of equivalent local martingale measures. As examples, we discuss (complete) one-dimensional models as well as (incomplete) lognormal mixture and popular stochastic volatility models. We also provide a detailed analysis of the case where the unique optimizer of the dual problem does not have a continuous law, leading to optimization problems whose solvability by duality methods depends on the initial wealth of the investor.

Keywords:
portfolio optimization, fund manager’s problem, incentive scheme, convex duality, delegated portfolio management
MSC:
91G10, 90C26
journal: Finance and Stochastics

JEL Subject Classification G11

1 Introduction

Whereas classical portfolio theory studies utility maximization from the point of view of an investor, whose preferences are modeled by a concave utility function, in reality, portfolio management is commonly delegated to a fund manager. To increase the efficacy of the manager, he is often paid by an incentive scheme that depends on the performance of the fund he manages. Such a scheme can be composed, for example, of a fixed fee, some percentage of the fund, plus an additional reward, which consists of one (or a combination of several) call options on the fund. As a consequence, two differences to the classical setting arise. First, the utility function, under which the optimization is carried out, does not represent the preference structure of the investor (also called the principal), but rather the manager’s (the agent’s) preference structure. Second, what is optimized under this utility function is not the terminal value of the fund itself, but rather some function of it, which depends on the specific incentive scheme.

The resulting optimization problem is, in general, no longer concave, and therefore does not fit into the classical setting first studied by Merton Merton , who used a stochastic optimal control approach. Specifically, Merton derived a Hamilton-Jacobi-Bellman (HJB) equation satisfied by the value function and found a closed form solution in the case of power utility. The drawback of this method – namely that it requires the state process to be Markov – can be overcome by using the fact that the processes dual to the portfolio processes are given via the set of equivalent local martingale measures, as pioneered by Karatzas, Lehoczky and Shreve KILS and Pliska Pliska . A thorough study of the portfolio optimization problem in a general (incomplete) semimartingale market was conducted by Kramkov and Schachermayer KramSchach1 , KramSchach2 , Bouchard, Touzi and Zeghal BTZ and others.

As pointed out, all of the above literature concentrates on the principal investor himself. The problem becomes more involved, if the investor, rather than investing himself, delegates his money to a fund manager. The agent invests on the principal’s behalf, in exchange for a fee schedule, which is based on the fund’s performance at the final time TT, and given by a function gg of the portfolio at terminal time. We assume that the agent’s utility function UU is smooth, strictly concave and has a domain bounded from below. These assumptions allow for the classical examples of power and logarithmic utility (but not utility functions defined on the whole real line such as the exponential). The fee schedule function gg is assumed to be convex and dominated by an affine function – i.e., its slope has to be bounded; without loss of generality we will assume that the maximal slope is 11. The financial reasoning for these assumptions on gg is that we expect the manager’s fees to increase as the fund’s profit increases. Therefore, gg should be convex. The fund manager’s utility, which results from his payoff, is hence a composition of the two functions, U¯:=Ug\bar{U}:=U\circ g, and may no longer be concave. Thus, the previously mentioned results are no longer applicable.

The resulting problem is not well understood; the existing literature discusses mainly the question of whether such a compensation scheme leads the portfolio manager to take excessive risk. In Ross , Ross discusses some conditions to make the agent more or less risk averse then the principal. Carpenter shows in carp the existence of the fund manager’s optimal portfolio in case of a utility function UU with constant relative risk aversion and a call option like fee schedule gg in a Brownian stock price model. In this setting, her analysis is generalized by Larsen Lars into an agency problem, where the investor optimizes the resulting payoff over piecewise affine incentive schemes, which he might choose to offer the portfolio manager.

We want to point out that there is also a different approach to portfolio optimization under incentive schemes, in which the compensation is based on high-watermarks, i.e., the running maximum of the fund. Recent references to this compensation approach include ObGua , JanSir , PanWest . In all of these papers the authors also assume a Brownian stock price model and solve the appropriate HJB equation.

In the present paper we will investigate the more fundamental problem of existence and uniqueness of an agent’s optimal investment portfolio in a general semimartingale model. As noted above, the resulting fund manager’s utility function U¯\bar{U} may not be concave. It is well-known that the solution is then to concavify U¯\bar{U}, and solve the concavified problem instead. Even though this new utility is now concave, it is not necessarily strictly concave, nor does it necessarily satisfy the usual Inada condition at zero, both of which are needed in the classical utility maximization framework. Moreover, the smoothness of the concavified function is not clear a priori. Using a dynamic programming approach via HJB equation is – at least in the straightforward way – also not possible, since the concavified utility function can (and usually will) be affine in some parts, and hence finding the optimal portfolio becomes impossible. Thus, we have effectively to weaken the utility function requirement of Kramkov and Schachermayer KramSchach1 . Our approach is to use the more general framework of Bouchard, Touzi and Zeghal BTZ and, by proving additional regularity of the concavified utility function, show the uniqueness of the dual optimizer. We are thus able to utilize the abstract framework of Bouchard, Touzi and Zeghal in a concrete setting, which is a rare feat (note, however, the exception of Seifried Seifried , who discusses capital gains taxes in a complete market).

The next step is to develop sufficient conditions, broad enough to be of interest, for the solution of the concavified problem to be also the solution of the original problem. It turns out that a necessary and sufficient condition is that the corresponding unique dual optimizer has a continuous law (i.e., the distribution of the random variable has no atoms). A similar procedure can be found in a related paper by Carassus and Pham Carassus , who consider a problem of portfolio optimization in a complete market with Brownian stock price, with a utility function created by two piecewise concave functions. We show that the condition of atomlessness holds, not only true in the classical Black-Scholes model with discounted stock price having nonzero drift, but also in two example classes of markets, independent of the initial capital of the fund and independent of the concrete incentive schemes: (1) complete one-dimensional Itô-process models (such as local volatility models), and (2) incomplete lognormal mixture and stochastic volatility models (such as the popular correlated Hull-White, Scott and Heston models).

The practical consequence of this is that the agent shuns successfully away from any part of the domain where the concavified utility function is linear . However, he does this in a smooth way. The optimal terminal wealth has no atoms except possibly at zero (meaning that the fund manager jeopardizes the fund with a positive probability), and it is zero under any linear spot of the concavified utility function.

If the assumption on the non-atomic structure of the dual optimizers fails, we are still able to give an affirmative answer, albeit only for some initial capitals. In general, the fund manager’s optimal wealth does not have to agree with the one calculated from the concavified problem, and even if it does, it does not have to be unique. As a note of caution, we present easy counterexamples that show that this method should not be implemented without proper conditions. We also give simple examples for our theorems, which conceptually present how the optimal portfolio can be explicitly calculated in a complete market setting.

The rest of this paper is organized as follows. In Section 2 we introduce the mathematical model of our delegated portfolio optimization problem and state our main results. The two following sections are devoted to examples illustrating our findings. In Section 3 we discuss in detail the case of power utility in a Black-Scholes market, highlighting the central importance of the distributional properties of the dual optimizer and investigating the problem from the point of view of the managers risk aversion. Section 4 contains several complete and incomplete market models in which our assumptions hold true. The remaining sections are devoted to the more technical side of the problem. Section 5 provides the background on general results on smooth and non-smooth duality theory and discusses how they can put to work for our needs. Section 6 contains the detailed proofs on the relationship of the conacavified and the dual problem. Section  refsec:7 draws the conclusions for the original problem and contains the proof of the main theorem. Finally, Section 8 discusses the limitations of the main theorem and provides partial results for an atomic dual optimizer. The conclusions of our exposition are summarized in Section 9.

After finishing a first version of the present paper, we have learnt of the work of Reichlin Rei , who studies the utility maximization problem for more general non-concave utility functions under a fixed pricing measure.

2 Setting and Main Results

We start by reviewing utility maximization in a general semimartingale framework and state our main results. Assume that SiS^{i}, i=1,,di=1,\ldots,d is a d-dimensional, locally bounded semimartingale on a filtered probability space (Ω,,(t)0tT,)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{0\leq t\leq T},\mathbb{P}), representing discounted stock price processes; without loss of generality we assume T=\mathcal{F}_{T}=\mathcal{F}. We focus on portfolio processes with initial capital xx and predictable and SS-integrable hedging strategies HH. The value process of such a portfolio is then given by

Xtx,H=x+0tHs𝑑Ss,0tT.X_{t}^{x,H}=x+\int_{0}^{t}H_{s}\,dS_{s},\qquad 0\leq t\leq T.

Denote by 𝒳(x)\mathcal{X}(x) the set of all nonnegative wealth processes with initial capital xx,

𝒳(x)\displaystyle\mathcal{X}(x) ={X0:Xt=Xtx,H for some predictable and S-integrable strategy H.\displaystyle=\Bigl{\{}X\geq 0\,:\,X_{t}=X_{t}^{x,H}\mbox{ for some predictable and $S$-integrable strategy }H\Bigr{.}
. for every 0tT}.\displaystyle\phantom{=\Bigl{\{}\Bigr{.}}\Bigl{.}\mbox{ for every }0\leq t\leq T\Bigr{\}}. (1)

We refer to 𝒳(x)\mathcal{X}(x) as the set of all admissible wealth processes.

We want to look at the portfolio optimization problem from the perspective of a portfolio manager, who is paid with incentives that depend on the performance of the portfolio at some future time T>0T>0. In this article we allow the incentive scheme to be a function g:00g:\mathbb{R}_{\geq 0}\to\mathbb{R}_{\geq 0}, nonconstant, nondecreasing, convex and with maximal slope c>0c>0, i.e.,

supx0g(x)c.\sup\bigcup_{x\geq 0}\partial g(x)\leq c. (2)

We note that the agent’s private capital can be absorbed into gg (if positive). To simplify the exposition, we will assume throughout this text that, without loss of generality, c=1c=1. Setting U¯:=Ug\bar{U}:=U\circ g, the portfolio manager’s utility maximization problem is

u(x):=supX𝒳(x)𝔼[U¯(XT)].u(x):=\sup_{X\in\mathcal{X}(x)}\mathbb{E}\bigl{[}\bar{U}\bigl{(}X_{T}\bigr{)}\bigr{]}. (3)
Assumption 1

To make the problem nontrivial, we assume that there exists at least some x0>0x_{0}>0 such that

supX𝒳(x0)𝔼[U(XT)]<.\sup_{X\in\mathcal{X}(x_{0})}\mathbb{E}\bigl{[}U\bigl{(}X_{T}\bigr{)}\bigr{]}<\infty.
Assumption 2

To preclude the possibility of arbitrage in the sense of ‘free lunch with vanishing risk’ (for details see the work of Delbaen and Schachermayer, DS ) we assume that the set of equivalent local martingale measures is not empty,

e={:,S is a local -martingale}.\mathcal{M}^{e}=\Bigl{\{}\mathbb{Q}\,:\,\mathbb{Q}\sim\mathbb{P},\,S\mbox{ is a local }\mathbb{Q}\mbox{-martingale}\Bigr{\}}\neq\emptyset.
Assumption 3

The fund manager’s preferences are represented by a utility function U:>0U:{\mathbb{R}_{>0}}\rightarrow\mathbb{R} (without loss of generality we assume U():=limxU(x)>0U(\infty):=\lim\limits_{x\to\infty}U(x)>0).

  • a)

    We assume that UU is strictly increasing, strictly concave and continuously differentiable on >0{\mathbb{R}_{>0}}; we extend UU continuously to 0\mathbb{R}_{\geq 0}, allowing the value -\infty at 0;

  • b)

    The utility function satisfies the Inada-conditions

    U(0):=limx0U(x)=,U():=limxU(x)=0,U^{\prime}(0):=\lim\limits_{x\rightarrow 0}U^{\prime}(x)=\infty,\qquad\qquad U^{\prime}(\infty):=\lim\limits_{x\rightarrow\infty}U^{\prime}(x)=0, (4)
  • c)

    Moreover, it satisfies the asymptotic elasticity condition

    AE(U):=lim supxxU(x)U(x)<1.AE(U):=\limsup_{x\to\infty}\frac{xU^{\prime}(x)}{U(x)}<1. (5)

These three standard assumptions of utility maximization problems (see, e.g., KramSchach1 ) will be the standing assumptions for the rest of this paper.

Before introducing the dual problem, we recall some standard notions and notation of convex analysis. A function f:𝕌[,]f\,:\,\mathbb{U}\subseteq\mathbb{R}\to[-\infty,\infty] defined on some convex domain 𝕌\mathbb{U} is called convex (respectively concave) if its epigraph (respectively hypograph)

epif:={(x,μ)𝕌×:f(x)μ},hypof:={(x,μ)𝕌×:f(x)μ},\operatorname{epi\,}f:=\bigl{\{}(x,\mu)\in\mathbb{U}\times\mathbb{R}\,:\,f(x)\leq\mu\bigr{\}},\qquad\operatorname{hypo\,}f:=\bigl{\{}(x,\mu)\in\mathbb{U}\times\mathbb{R}\,:\,f(x)\geq\mu\bigr{\}},

is a convex set. The effective domain of a convex function ff is defined as

domf:={x𝕌:f(x)<}.\operatorname{dom\,}f:=\bigl{\{}x\in\mathbb{U}\subseteq\mathbb{R}\,:\,f(x)<\infty\bigr{\}}.

Similarly, for a concave function, we define its domain as the set of points in the pre-image not mapping to -\infty. Generalizing the usual notations from utility maximization problems in an obvious way, we define, for any function ff dominated by some affine function, its convex conjugate ff^{*} and its biconjugate ff^{**} by

f(y):=supxdomf(f(x)xy),f(x):=infydomf(f(y)+xy).f^{*}(y):=\sup_{x\in\operatorname{dom\,}f}\Bigl{(}f(x)-xy\Bigr{)},\qquad f^{**}(x):=\inf_{y\in\operatorname{dom\,}f^{*}}\Bigl{(}f^{*}(y)+xy\Bigr{)}.

Note that ff^{**} is the concavification of ff, i.e., the hypograph of ff^{**} is the closed convex hull of the hypograph of ff, hypof=co(hypof)¯\operatorname{hypo\,}f^{**}=\overline{\operatorname{co\,}(\operatorname{hypo\,}f)}. We note that ff^{*} is the convex conjugate of f()-f(-\,\cdot\,) in the classical sense of convex analysis. We will use standard results of convex analysis (cf., e.g., HUL ) with the obvious modifications without further notice.

We note that the function U¯\bar{U} is not necessarily concave, placing the problem (3) outside the standard setting of utility maximization. Instead of analyzing the non-concave problem (3) directly, we will first consider the concavified problem

w(x):=supW𝒳(x)𝔼[U¯(WT)].w(x):=\sup_{W\in\mathcal{X}(x)}\mathbb{E}\bigl{[}\bar{U}^{**}\bigl{(}W_{T}\bigr{)}\bigr{]}. (6)

Similar to KramSchach1 we define the set of process dual to (2) by

𝒴(y):={Y0:Y0=y and XY is a supermartingale for all X𝒳(1)}.\mathcal{Y}(y):=\Bigl{\{}Y\geq 0\,:\,Y_{0}=y\mbox{ and }XY\mbox{ is a supermartingale for all }X\in\mathcal{X}(1)\Bigr{\}}.

It turns out then that both problems share the dual problem (see Theorem 2.1 below), i.e.,

v(y):=infY𝒴(y)𝔼[U¯(YT)].v(y):=\inf_{Y\in\mathcal{Y}(y)}\mathbb{E}\bigl{[}\bar{U}^{*}\bigl{(}Y_{T}\bigr{)}\bigr{]}. (7)

In general the concavified utility function U¯\bar{U}^{**} will be neither strictly concave nor satisfy the Inada condition at 0. Hence, we will have to rely on results for nonsmooth utility maximization (see Theorem 5.2 for more details). We will see that Assumptions 1, 2, and 3 place us in a setting where we will be able to apply this theorem.

Finally let

β:=inf{x>0:U¯(x)>}[0,).\beta:=\inf\{x>0:\bar{U}(x)>-\infty\}\in[0,\infty).

The following are the main theorems of this paper. Theorem 2.1 establishes the duality relationship between vv and ww, and relates W^T(x)\hat{W}_{T}(x) and Y^T(y)\hat{Y}_{T}(y), the optimizers of the problems (6) and (7), respectively. Theorem 2.2 provides conditions under which X^T(x)\hat{X}_{T}(x) and W^T(x)\hat{W}_{T}(x) – the optimizers of the problems (3) and (6), respectively, are the same.

Theorem 2.1

For the utility optimization problem under a convex incentive scheme gg it holds that

  • a)

    The functions uu and ww are finite on (β,)(\beta,\infty) as is vv on >0\mathbb{R}_{>0}, and v=wv=w^{*}. Moreover, vv is strictly convex on the whole domain (0,)(0,\infty) if U(0)=U(0)=-\infty. Otherwise, there exists some δ(0,]\delta\in(0,\infty] such that vv is convex on the interval (0,δ)(0,\delta) and constant U¯(0)\bar{U}^{**}(0) on [δ,)[\delta,\infty). The function ww is continuously differentiable on (β,)(\beta,\infty), and concave.

  • b)

    The optimizer Y^(y)\hat{Y}(y) of the dual problem (7) exists for every y>0y>0 and is a.s. unique on (0,(U¯)(0))\bigl{(}0,\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(0)\bigr{)}.

  • c)

    For x>βx>\beta there exists an optimizer W^(x)\hat{W}(x) of the concavified problem (6) that satisfies

    W^T(x)U¯(Y^T(y))\hat{W}_{T}(x)\in-\partial\bar{U}^{*}\bigl{(}\hat{Y}_{T}(y)\bigr{)}

    for y=w(x)y=w^{\prime}(x) such that W^(x)𝒳(x)\hat{W}(x)\in\mathcal{X}(x) and W^(x)Y^(y)\hat{W}(x)\hat{Y}(y) is a uniformly integrable martingale.

  • d)

    Additionally we have

    v(y)=infe𝔼[U¯(ydd)],v(y)=\inf_{\mathbb{Q}\in\mathcal{M}^{e}}\mathbb{E}\biggl{[}\bar{U}^{*}\biggl{(}y\frac{d\mathbb{Q}}{d\mathbb{P}}\biggr{)}\biggr{]},

    however the infimum is in general not attained in e\mathcal{M}^{e}.

Theorem 2.2

Assume that for every y(0,w(β)]y\in\bigl{(}0,w^{\prime}(\beta)\bigr{]} the terminal value of the dual optimizers Y^(y)\hat{Y}(y) has a continuous cumulative distribution function. Then

  • a)

    The optimizer W^(x)\hat{W}(x) for the concavified problem (6) is unique for every x>βx>\beta.

  • b)

    For every x>βx>\beta there exists a solution X^(x)\hat{X}(x) of the original problem (3) and this solution is unique. It coincides also with W^(x)\hat{W}(x), the solution of the concavified problem (6).

At a first glance the condition that the distribution of the dual optimizer has no atoms seems quite abstract and hard to check. Therefore, we present a sufficient condition for no atoms, in terms of equivalent local martingale measures, which can be checked more easily in some concrete models.

Proposition 1

Assume that the laws of the Radon-Nikodým derivatives ZT=dd|TZ_{T}=\frac{d\mathbb{Q}}{d\mathbb{P}}|_{\mathcal{F}_{T}}, e\mathbb{Q}\in\mathcal{M}^{e}, are uniformly absolutely continuous with respect to the Lebesgue measure λ\lambda on >0{\mathbb{R}_{>0}} (i.e., the densities dZT1dλ\frac{d\mathbb{P}\circ Z_{T}^{-1}}{d\lambda} are uniformly integrable). Then the terminal value of the optimizer Y^(y)\hat{Y}(y) of the dual problem (7) has a continuous law.

While this assumption it is quite restrictive, it is the only one we know which works in general without having a priori knowledge of the maximizer. It is in particular satisfied in the Black-Scholes model with nonzero drift. We will show in section 4 that these assumptions are also satisfied in other incomplete market models, such as lognormal mixture models. In more general models – as stochastic volatility models, see also section 4 – one can nevertheless derive the result, if one has some a priori knowledge about the optimizer, essentially depending on the measurability properties of the Sharpe ratio.

3 Examples around the Black-Scholes model

We first present our findings for power utility maximization in the Black-Scholes model with an incentive gg of call option type: g(x)=λ(xk)+g(x)=\lambda(x-k)^{+}. This setting not only allows us to connect our results to previous work carp and provide explicit solutions, but it also allows us to illustrate the degeneracy if the Sharpe ratio vanishes – producing a purely atomic Radon-Nikodým derivative. Another benefit of studying this setting is that it allows us to analyze the situation from the point of view of the (relative) risk aversion of the manager. Specifically, we address the issue of the optimal relation between the number of options and the value of the strike among all those producing the same relative risk aversion (compare this to Ross ).

Example 1

Assume that the discounted stock price is modeled by

St=exp(σWt+(μσ2/2)t),μ0,σ>0,S_{t}=\exp{\Bigl{(}\sigma W_{t}+\bigl{(}\mu-\sigma^{2}/2\bigr{)}t\Bigr{)}},\qquad\mu\geq 0,\,\,\sigma>0,

for some Brownian motion WW generating the filtration (t)(\mathcal{F}_{t}). These stock price dynamics, together with the riskless numéraire, describe a complete market. The set of all equivalent local martingale measures is hence the singleton e={}\mathcal{M}^{e}=\{\mathbb{Q}\}, where the measure \mathbb{Q} is given by the Radon-Nikodým density ZT=dd|T=exp{θWTθ22T}Z_{T}=\frac{d\mathbb{Q}}{d\mathbb{P}}|_{\mathcal{F}_{T}}=\exp{\bigl{\{}-\theta W_{T}-\frac{\theta^{2}}{2}T\bigr{\}}} with market price of risk θ:=μσ\theta:=\frac{\mu}{\sigma} and Wt:=Wt+θtW^{\mathbb{Q}}_{t}:=W_{t}+\theta t is a \mathbb{Q}-Brownian motion. Furthermore, let the incentive scheme be given by g(x)=λ(xk)+g(x)=\lambda(x-k)^{+}, k>0k>0, 0<λ<10<\lambda<1.

The portfolio manager’s utility will be given by the function be UU. We will now consider two cases, U(0)=U(0)=-\infty and U(0)>U(0)>-\infty. In the second case, we will assume without loss of generality that U(0)=0U(0)=0 and find

U¯(x)\displaystyle\bar{U}(x) ={00xk,U(λ(xk))x>k,U¯(x)={yx0xx,U(λ(xk))x>x,\displaystyle=\left\{\begin{array}[]{ll}0&0\leq x\leq k,\\ U\bigl{(}\lambda(x-k)\bigr{)}&x>k,\end{array}\right.\quad\bar{U}^{**}(x)=\left\{\begin{array}[]{ll}y_{*}x&0\leq x\leq x_{*},\\ U\bigl{(}\lambda(x-k)\bigr{)}&x>x_{*},\end{array}\right.
U¯(y)\displaystyle\bar{U}^{*}(y) ={U(y/λ)ky0<yy,0y>y,\displaystyle=\left\{\begin{array}[]{ll}U^{*}(y/\lambda)-ky&0<y\leq y_{*},\\ 0&y>y_{*},\end{array}\right.

where xx_{*} is the solution of λxU(λ(xk))=U(λ(xk))\lambda x_{*}U^{\prime}\bigl{(}\lambda(x_{*}-k)\bigr{)}=U\bigl{(}\lambda(x_{*}-k)\bigr{)}, and y=λU(λ(xk))y_{*}=\lambda U^{\prime}\bigl{(}\lambda(x_{*}-k)\bigr{)}. In the former case

U¯(x)=U¯(x)={0<xk,U(λ(xk)+)x>k,U¯(y)=U(y/λ)ky,y>0.\bar{U}^{**}(x)=\bar{U}(x)=\left\{\begin{array}[]{ll}-\infty&0<x\leq k,\\ U\bigl{(}\lambda(x-k)^{+}\bigr{)}&x>k,\end{array}\right.\quad\bar{U}^{*}(y)=U^{*}(y/\lambda)-ky,\;\;y>0.

To make our example more computationally tractable, we will focus on the power utility case U(x)=xppU(x)=\frac{x^{p}}{p} with 0<p<10<p<1. Here U(y)=1ppypp1U^{*}(y)=\frac{1-p}{p}y^{\frac{p}{p-1}}, x=k1px_{*}=\frac{k}{1-p}, and y=λp(p1pk)p1y_{*}=\lambda^{p}\bigl{(}\frac{p}{1-p}k\bigr{)}^{p-1}, and thus

U¯(x)\displaystyle\bar{U}(x) ={00xk,(λ(xk))ppx>k,U¯(x)={λpx(pk1p)p10xk1p,(λ(xk))ppx>k1p,\displaystyle=\left\{\begin{array}[]{ll}0&0\leq x\leq k,\\ \frac{(\lambda(x-k))^{p}}{p}&x>k,\end{array}\right.\quad\bar{U}^{**}(x)=\left\{\begin{array}[]{ll}\lambda^{p}x(\frac{pk}{1-p})^{p-1}&0\leq x\leq\frac{k}{1-p},\\ \frac{\bigl{(}\lambda(x-k)\bigr{)}^{p}}{p}&x>\frac{k}{1-p},\end{array}\right. (12)
U¯(y)\displaystyle\bar{U}^{*}(y) ={1pp(yλ)pp1ky0<yy,0y>y.\displaystyle=\left\{\begin{array}[]{ll}\frac{1-p}{p}\Bigl{(}\frac{y}{\lambda}\Bigr{)}^{\frac{p}{p-1}}-ky&0<y\leq y_{*},\\ 0&y>y_{*}.\end{array}\right. (15)

For the illustration in Figure 1 we assumed p=12,λ=14,k=3,U(x)=2xp=\frac{1}{2},\lambda=\frac{1}{4},k=3,U(x)=2\sqrt{x}. Then it is easily seen that x=6,y=36x_{*}=6,~y_{*}=\frac{\sqrt{3}}{6} and

U¯(x)\displaystyle\bar{U}(x) ={00x3,x3x>3,U¯(x)={36x0x6,x3x>6,\displaystyle=\left\{\begin{array}[]{ll}0&0\leq x\leq 3,\\ \sqrt{x-3}&x>3,\end{array}\right.\quad\bar{U}^{**}(x)=\left\{\begin{array}[]{ll}\frac{\sqrt{3}}{6}x&0\leq x\leq 6,\\ \sqrt{x-3}&x>6,\end{array}\right. (20)
U¯(y)\displaystyle\bar{U}^{*}(y) ={14y3y0<y36,0y>36.\displaystyle=\left\{\begin{array}[]{ll}\frac{1}{4y}-3y&0<y\leq\frac{\sqrt{3}}{6},\\ 0&y>\frac{\sqrt{3}}{6}.\end{array}\right. (23)
Refer to caption
Refer to caption
Figure 1: Left: Composed utility function U¯\bar{U} of equation (23) and its concavification U¯\bar{U}^{**}. Right: Dual utility function U¯\bar{U}^{*}.

The non-atomicity condition on the dual optimizer (here just the Radon Nikodým derivative dd\frac{d\mathbb{Q}}{d\mathbb{P}}) implies that we have to consider two distinct cases: either μ\mu is positive (equivalently, we could assume μ\mu negative), or μ=0\mu=0. The reason for this distinction lies in the fact that with θ=0\theta=0, the original measure \mathbb{P} is already the (unique) risk-neutral measure \mathbb{Q}. In other words, ZT=dd|T=1Z_{T}=\frac{d\mathbb{Q}}{d\mathbb{P}}|_{\mathcal{F}_{T}}=1. Hence, the random variable ZTZ_{T} has an atom of mass one at 11. In the other case the random variable ZT=dd|TZ_{T}=\frac{d\mathbb{Q}}{d\mathbb{P}}|_{\mathcal{F}_{T}} has a (smooth) density.

3.1 Case 1: θ>0\theta>0

We proceed first with the case in which the random variable ZTZ_{T} has a density. Tedious, but straightforward stochastic calculus reveals the following results. The dual value function is given by

v(y)\displaystyle v(y) =𝔼[U¯(yZT)]=1pp(yλ)pp1ep(1p)2θ2T2Φ(d+)kyΦ(d),\displaystyle=\mathbb{E}\bigl{[}\bar{U}^{*}(yZ_{T})\bigr{]}=\frac{1-p}{p}\Bigl{(}\frac{y}{\lambda}\Bigr{)}^{\frac{p}{p-1}}e^{\frac{p}{(1-p)^{2}}\frac{\theta^{2}T}{2}}\Phi\bigl{(}d_{+}\bigr{)}-ky\Phi\bigl{(}d_{-}\bigr{)},
d+\displaystyle d_{+} :=logylogyθT+(12+p1p)θT,\displaystyle:=\frac{\log y_{*}-\log y}{\theta\sqrt{T}}+\Bigl{(}\frac{1}{2}+\frac{p}{1-p}\Bigr{)}\theta\sqrt{T},
d\displaystyle d_{-} :=logylogyθTθT,\displaystyle:=\frac{\log y_{*}-\log y}{\theta\sqrt{T}}-\theta\sqrt{T},

where Φ\Phi is the cumulative distribution function of the normal distribution. Thus, vv is continuously differentiable and strictly concave on the whole real line. Therefore, using the fact that w=vw^{*}=v, the function ww is also strictly concave and continuously differentiable on >0{\mathbb{R}_{>0}}. For the terminal value of the optimizer we obtain (we can use almost everywhere defined derivatives since the law of Y^T(y)=ydd\hat{Y}_{T}(y)=y\frac{d\mathbb{Q}}{d\mathbb{P}} has no atoms)

X^T(x)\displaystyle\hat{X}_{T}(x) =W^T(x)=(U¯)(Y^T(w(x)))\displaystyle=\hat{W}_{T}(x)=-\bigl{(}\bar{U}^{*}\bigr{)}^{\prime}\Bigl{(}\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)}\Bigr{)}
=((λpeθWT+θ2T2w(x))11p+k)1l{WT>1θ(logw(x)logy)θT2}.\displaystyle=\left(\Biggl{(}\frac{\lambda^{p}e^{\theta W_{T}+\frac{\theta^{2}T}{2}}}{w^{\prime}(x)}\Biggr{)}^{\frac{1}{1-p}}+k\right){\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\bigl{\{}W_{T}>\frac{1}{\theta}\bigl{(}\log w^{\prime}(x)-\log y_{*}\bigr{)}-\frac{\theta T}{2}\bigr{\}}}.

To compute the optimal strategy we simply observe that

f(t,z)\displaystyle f(t,z) =12π(Tt)1θ(logw(x)logy)+θT2e(yz)22(Tt)((λpeθyθ2T2w(x))11p+k)𝑑y\displaystyle=\frac{1}{\sqrt{2\pi(T-t)}}\int_{\frac{1}{\theta}\bigl{(}\log w^{\prime}(x)-\log y_{*}\bigr{)}+\frac{\theta T}{2}}^{\infty}e^{-\frac{(y-z)^{2}}{2(T-t)}}\left(\Biggl{(}\frac{\lambda^{p}e^{\theta y-\frac{\theta^{2}T}{2}}}{w^{\prime}(x)}\Biggr{)}^{\frac{1}{1-p}}+k\right)dy

solves the (reverse) heat equation on [0,T)×[0,T)\times\mathbb{R} with terminal condition

f(T,z)=((λpeθzθ2T2w(x))11p+k)1l{z>1θ(logw(x)logy)+θT2}f(T,z)=\left(\Biggl{(}\frac{\lambda^{p}e^{\theta z-\frac{\theta^{2}T}{2}}}{w^{\prime}(x)}\Biggr{)}^{\frac{1}{1-p}}+k\right){\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\bigl{\{}z>\frac{1}{\theta}\bigl{(}\log w^{\prime}(x)-\log y_{*}\bigr{)}+\frac{\theta T}{2}\bigr{\}}}

satisfying f(T,WT)=X^T(x)f(T,W_{T}^{\mathbb{Q}})=\hat{X}_{T}(x). Then it follows from Itô’s formula that

X^T(x)=x+0Tfz(t,Wt)𝑑Wt=x+0Tfz(t,Wt+θt)σSt𝑑St.\hat{X}_{T}(x)=x+\int_{0}^{T}f_{z}(t,W_{t}^{\mathbb{Q}})dW_{t}^{\mathbb{Q}}=x+\int_{0}^{T}\frac{f_{z}(t,W_{t}+\theta t)}{\sigma S_{t}}dS_{t}.

Thus, the optimal strategy (in terms of cash invested in stock) is simply fz(t,Wt+θt)σ\frac{f_{z}(t,W_{t}+\theta t)}{\sigma}.

Explicitly, we derive

f(t,z)=\displaystyle f(t,z)= (λpw(x))11peθz1p+θ22(1p)(T+Tt1p)\displaystyle\Biggl{(}\frac{\lambda^{p}}{w^{\prime}(x)}\Biggr{)}^{\frac{1}{1-p}}e^{\frac{\theta z}{1-p}+\frac{\theta^{2}}{2(1-p)}\bigl{(}-T+\frac{T-t}{1-p}\bigr{)}}
Φ(logw(x)logyθTtθT2Tt+zTt+θTt1p)\displaystyle\phantom{=}\cdot\Phi\Biggl{(}-\frac{\log w^{\prime}(x)-\log y_{*}}{\theta\sqrt{T-t}}-\frac{\theta T}{2\sqrt{T-t}}+\frac{z}{\sqrt{T-t}}+\frac{\theta\sqrt{T-t}}{1-p}\Biggr{)}
+kΦ(logw(x)logyθTtθT2Tt+zTt).\displaystyle+k\Phi\Biggl{(}-\frac{\log w^{\prime}(x)-\log y_{*}}{\theta\sqrt{T-t}}-\frac{\theta T}{2\sqrt{T-t}}+\frac{z}{\sqrt{T-t}}\Biggr{)}.

Thus, the optimal strategy in terms of money invested in stock is given by

Ht(x)=\displaystyle H_{t}(x)= θ(1p)σ(λpw(x))11peθ(Wt+θt)1p+θ22(1p)(T+Tt1p)\displaystyle\frac{\theta}{(1-p)\sigma}\Biggl{(}\frac{\lambda^{p}}{w^{\prime}(x)}\Biggr{)}^{\frac{1}{1-p}}e^{\frac{\theta(W_{t}+\theta t)}{1-p}+\frac{\theta^{2}}{2(1-p)}\bigl{(}-T+\frac{T-t}{1-p}\bigr{)}}
Φ(logw(x)logyθTtθT2Tt+Wt+θtTt+θTt1p)\displaystyle\phantom{=}\cdot\Phi\Biggl{(}-\frac{\log w^{\prime}(x)-\log y_{*}}{\theta\sqrt{T-t}}-\frac{\theta T}{2\sqrt{T-t}}+\frac{W_{t}+\theta t}{\sqrt{T-t}}+\frac{\theta\sqrt{T-t}}{1-p}\Biggr{)}
+1σTt(λpw(x))11peθ(Wt+θt)1p+θ22(1p)(T+Tt1p)\displaystyle+\frac{1}{\sigma\sqrt{T-t}}\Biggl{(}\frac{\lambda^{p}}{w^{\prime}(x)}\Biggr{)}^{\frac{1}{1-p}}e^{\frac{\theta(W_{t}+\theta t)}{1-p}+\frac{\theta^{2}}{2(1-p)}\bigl{(}-T+\frac{T-t}{1-p}\bigr{)}}
φ(logw(x)logyθTtθT2Tt+Wt+θtTt+θTt1p)\displaystyle\phantom{=}\cdot\varphi\Biggl{(}-\frac{\log w^{\prime}(x)-\log y_{*}}{\theta\sqrt{T-t}}-\frac{\theta T}{2\sqrt{T-t}}+\frac{W_{t}+\theta t}{\sqrt{T-t}}+\frac{\theta\sqrt{T-t}}{1-p}\Biggr{)}
+kσTtφ(logw(x)logyθTtθT2Tt+Wt+θtTt).\displaystyle+\frac{k}{\sigma\sqrt{T-t}}\varphi\Biggl{(}-\frac{\log w^{\prime}(x)-\log y_{*}}{\theta\sqrt{T-t}}-\frac{\theta T}{2\sqrt{T-t}}+\frac{W_{t}+\theta t}{\sqrt{T-t}}\Biggr{)}.

We note that this can also be derived more generally by using (OconeKaratzas, , Theorem 6.2) and observing that their condition (6.7) is always satisfied in our case, as asymptotic elasticity implies the inequalities of (KramSchach1, , Lemma 6.3) (where we only have to replace the derivatives by suprema of subdifferentials).

On a more practical side, we would like to investigate the optimization problem conditional on the portfolio manager’s risk aversion. It turns out that the proper concept for this issue is to look on relative risk aversion (RRA) and to apply this also to the dual value function. Using the fact that (w)1=v(w^{\prime})^{-1}=v^{\prime}, it follows that v′′(y)=1w′′(x)v^{\prime\prime}(y)=\frac{1}{w^{\prime\prime}(x)} for y=w(x)y=w^{\prime}(x), and we can compute the relative risk aversion

RRAv(y)=yv′′(y)v(y)=w(x)xw′′(x)=1RRAw(x)=1RRAw(v(y)).RRA_{v}(y)=-\frac{yv^{\prime\prime}(y)}{v^{\prime}(y)}=-\frac{w^{\prime}(x)}{xw^{\prime\prime}(x)}=\frac{1}{RRA_{w}(x)}=\frac{1}{RRA_{w}(v^{\prime}(y))}.

Using the fact the d=d+=1yθTd_{-}^{\prime}=d_{+}^{\prime}=-\frac{1}{y\theta\sqrt{T}}, and that Φ′′(x)=xΦ(x)\Phi^{\prime\prime}(x)=-x\Phi^{\prime}(x), we obtain

RRAv(y)\displaystyle\phantom{=}RRA_{v}(y)
=λp1py1p1ep(1p)2θ2T2(Φ(d+)1p+Φ(d+)pθT+Φ(d+)θT1ppd+Φ(d+)θ2T)+kΦ(d)θT(1+dθT)λp1py1p1ep(1p)2θ2T2(Φ(d+)+1ppΦ(d+)θT)+kθTΦ(d)kΦ(d)\displaystyle=\frac{\lambda^{\frac{p}{1-p}}y^{\frac{1}{p-1}}e^{\frac{p}{(1-p)^{2}}\frac{\theta^{2}T}{2}}\Bigl{(}\frac{\Phi(d_{+})}{1-p}+\frac{\Phi^{\prime}(d_{+})}{p\theta\sqrt{T}}+\frac{\Phi^{\prime}(d_{+})}{\theta\sqrt{T}}-\frac{1-p}{p}\frac{d_{+}\Phi^{\prime}(d_{+})}{\theta^{2}T}\Bigr{)}+\frac{k\Phi^{\prime}(d_{-})}{\theta\sqrt{T}}\Bigl{(}1+\frac{d_{-}}{\theta\sqrt{T}}\Bigr{)}}{-\lambda^{\frac{p}{1-p}}y^{\frac{1}{p-1}}e^{\frac{p}{(1-p)^{2}}\frac{\theta^{2}T}{2}}\Bigl{(}\Phi(d_{+})+\frac{1-p}{p}\frac{\Phi^{\prime}(d_{+})}{\theta\sqrt{T}}\Bigr{)}+\frac{k}{\theta\sqrt{T}}\Phi^{\prime}(d_{-})-k\Phi(d_{-})}
=λp1py1p1ep(1p)2θ2T2(Φ(d+)1p+Φ(d+)(1pplogylogyθ3T32+121+pp))+kΦ(d)θTlogylogyθ2Tλp1py1p1ep(1p)2θ2T2(Φ(d+)+1ppΦ(d+)θT)+kθTΦ(d)kΦ(d).\displaystyle=\frac{\lambda^{\frac{p}{1-p}}y^{\frac{1}{p-1}}e^{\frac{p}{(1-p)^{2}}\frac{\theta^{2}T}{2}}\Bigl{(}\frac{\Phi(d_{+})}{1-p}+\Phi^{\prime}(d_{+})\bigl{(}-\frac{1-p}{p}\frac{\log y_{*}-\log y}{\theta^{3}T^{\frac{3}{2}}}+\frac{1}{2}\frac{1+p}{p}\bigr{)}\Bigr{)}+\frac{k\Phi^{\prime}(d_{-})}{\theta\sqrt{T}}\frac{\log y_{*}-\log y}{\theta^{2}T}}{-\lambda^{\frac{p}{1-p}}y^{\frac{1}{p-1}}e^{\frac{p}{(1-p)^{2}}\frac{\theta^{2}T}{2}}\Bigl{(}\Phi(d_{+})+\frac{1-p}{p}\frac{\Phi^{\prime}(d_{+})}{\theta\sqrt{T}}\Bigr{)}+\frac{k}{\theta\sqrt{T}}\Phi^{\prime}(d_{-})-k\Phi(d_{-})}.
Refer to caption
Refer to caption
Figure 2: Left: RRAvRRA_{v} for values of p=18,14,12,34p=\frac{1}{8},\frac{1}{4},\frac{1}{2},\frac{3}{4} in red, green, blue and black. Right: RRAvRRA_{v} for values of k=14,12,1,2k=\frac{1}{4},\frac{1}{2},1,2 in red, green, blue and black.

For the rest of this subsection, to highlight the dependency on kk and λ\lambda, we will write U¯k,λ(y)\bar{U}^{*}_{k,\lambda}(y), uk,λ(x)u_{k,\lambda}(x), vk,λ(x)v_{k,\lambda}(x), wk,λ(x)w_{k,\lambda}(x) for the concavified dual utility functions, the value function, its dual, and the concavified value function, respectively.

Notice that if we parametrize kk and λ\lambda by α>0\alpha>0 in the way that α(k(α),λ(α))=(ακ,α1ppl)\alpha\mapsto\bigl{(}k(\alpha),\lambda(\alpha)\bigr{)}=\Bigl{(}\alpha\kappa,\alpha^{\frac{1-p}{p}}l\Bigr{)}, then y=λp(α)(p1pk(α))p1=lp(1ppκ)1py_{*}=\lambda^{p}(\alpha)\Bigl{(}\frac{p}{1-p}k(\alpha)\Bigr{)}^{p-1}=l^{p}\bigl{(}\frac{1-p}{p\kappa}\bigr{)}^{1-p} does not depend on α\alpha. The same is true for d±d_{\pm}, and it follows that vk(α),λ(α)(y)=αvk(1),λ(1)(y)v^{k(\alpha),\lambda(\alpha)}(y)=\alpha v^{k(1),\lambda(1)}(y). This is not surprising, since the same scaling property holds for the concavified dual utility function U¯k(α),λ(α)=αU¯k(1),λ(1)\bar{U}^{*}_{k(\alpha),\lambda(\alpha)}=\alpha\bar{U}^{*}_{k(1),\lambda(1)}. Finally, we conclude that RRAvk(α),λ(α)(y)RRA_{v^{k(\alpha),\lambda(\alpha)}}(y) does not depend on α\alpha. That is, the relative risk aversion of the fund manager is does not change when his compensation scheme is scaled in the above way. Additionally, it is easily seen that

wk(α),λ(α)(x)=infy>0(vk(α),λ(α)(y)+xy)=αinfy>0(vk(1),λ(1)(y)+xαy)=αwk(1),λ(1)(xα).w^{k(\alpha),\lambda(\alpha)}(x)=\inf_{y>0}\Bigl{(}v^{k(\alpha),\lambda(\alpha)}(y)+xy\Bigr{)}=\alpha\inf_{y>0}\Bigl{(}v^{k(1),\lambda(1)}(y)+\frac{x}{\alpha}y\Bigr{)}=\alpha w^{k(1),\lambda(1)}\Bigl{(}\frac{x}{\alpha}\Bigr{)}.

Note that the family {vk(α),λ(α)(y)}α>0\bigl{\{}v^{k(\alpha),\lambda(\alpha)}(y)\bigr{\}}_{\alpha>0} includes all the functions (up to an additive constant) that have the same relative risk aversion as the original dual function vv.

This leads finally to the following questions. Among the incentive schemes with same relative risk aversion RRAw(x)RRA_{w}(x), is there is one that is optimal from the manger’s point of view? If so, how it can be characterized? It turns out that the answer to the former question is affirmative if there is some cc^{*} such that the elasticity of wk(1),λ(1)w^{k(1),\lambda(1)} is equal to one: E(wk(1),λ(1))(c)=1E(w^{k(1),\lambda(1)})(c^{*})=1. The elasticity of a utility function UU being defined as

E(U)(c):=cU(c)U(c),c>0,E(U)(c):=\frac{cU^{\prime}(c)}{U(c)},~c>0, (24)

(compare this with the definition of the asymptotic elasticity in (5)). Indeed, any solution α\alpha to αwk(α),λ(α)(x)=wk(1),λ(1)(xα)xα(wk(1),λ(1))(xα)=0\frac{\partial}{\partial\alpha}w^{k(\alpha),\lambda(\alpha)}(x)=w^{k(1),\lambda(1)}\bigl{(}\frac{x}{\alpha}\bigr{)}-\frac{x}{\alpha}(w^{k(1),\lambda(1)})^{\prime}\bigl{(}\frac{x}{\alpha}\bigr{)}=0 is precisely a solution to E(wk(1),λ(1))(xα)=1E\bigl{(}w^{k(1),\lambda(1)}\bigr{)}\bigl{(}\frac{x}{\alpha}\bigr{)}=1. Hence, an optimal solution is characterized via α=xc\alpha=\frac{x}{c^{*}}, and thus g(x)=(xc)1pp(xxc)+g(x)=\bigl{(}\frac{x}{c^{*}}\bigr{)}^{\frac{1-p}{p}}\bigl{(}x-\frac{x}{c^{*}}\bigr{)}^{+}. Moreover, because wk(α),λ(α)w^{k(\alpha),\lambda(\alpha)} is concave this is a maximum. In the case that E(wk(α),λ(α))<1E(w^{k(\alpha),\lambda(\alpha)})<1 on >0{\mathbb{R}_{>0}} there is no optimal α\alpha, as the manager’s expected utility increases as α\alpha tends to infinity since αwk(α),λ(α)>0\frac{\partial}{\partial\alpha}w^{k(\alpha),\lambda(\alpha)}>0.

3.2 Case 2: θ=0\theta=0

We now consider the second case, in which θ=0\theta=0. We remind the reader that we are still assuming that U(x)=xpp,0<p<1U(x)=\frac{x^{p}}{p},~0<p<1. Also, for future reference, note that

U¯(x)>U¯(x),x(0,x).\bar{U}^{**}(x)>\bar{U}(x),~x\in(0,x_{*}). (25)

Indeed, this holds for x(0,k]x\in(0,k], where from (15) U¯(x)=λpx(pk1p)p1>0\bar{U}^{**}(x)=\lambda^{p}x(\frac{pk}{1-p})^{p-1}>0, and for x(k,x)x\in(k,x_{*}) it is true, because (U¯)(x)<(U¯)(x)\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(x)<\bigl{(}\bar{U}\bigr{)}^{\prime}(x) and U¯(x)=U¯(x)\bar{U}^{**}(x_{*})=\bar{U}(x_{*}). As mentioned above this case is different from Case 1 as we have now e={}\mathcal{M}^{e}=\{\mathbb{P}\} with ZT=dd|T=1Z_{T}=\frac{d\mathbb{P}}{d\mathbb{P}}|_{\mathcal{F}_{T}}=1. It follows that the dual value function is given by v(y)=U¯(y)v(y)=\bar{U}^{*}(y) and thus, using the fact that w=vw^{*}=v, we have for the concavified problem w(x)=U¯(x)w(x)=\bar{U}^{**}(x).

We first consider the case when x(0,x)x\in(0,x_{*}). We have from duality that w(x)=U¯(x)w(x)=\bar{U}^{**}(x). This optimum is of course attained by the trivial strategy H0H\equiv 0 yielding the optimal wealth process W^(x)x\hat{W}(x)\equiv x for the concavified problem. However, plugging this into the original problem yields 𝔼[U¯(W^T(x))]=U¯(x)\mathbb{E}\bigl{[}\bar{U}\bigl{(}\hat{W}_{T}(x)\bigr{)}\bigr{]}=\bar{U}(x). Thus, W^T(x)\hat{W}_{T}(x) is an optimizer for the concavified problem, but yields a smaller value for the original problem. Moreover, W^T(x)\hat{W}_{T}(x) is even not an optimizer for the primal problem, as we will show that U¯(x)<u(x)=w(x)=U¯(x)\bar{U}(x)<u(x)=w(x)=\bar{U}^{**}(x).

In this example, a way around this problem can be seen by thinking in terms of investment strategies. Not only does the trivial strategy H0H\equiv 0 lead to the optimum for the concavified problem, but so does every strategy with terminal value W^T(x)\hat{W}_{T}(x) satisfying suppW^T(x)[0,x]\operatorname{supp\,}\hat{W}_{T}(x)\subseteq[0,x_{*}], since in the interval [0,x][0,x_{*}] the concavified utility function U¯\bar{U}^{**} is linear. Therefore, by the martingale property of the wealth process under \mathbb{P}, we have 𝔼[U¯(W^T(x))]=U¯(x)\mathbb{E}\bigl{[}\bar{U}^{**}\bigl{(}\hat{W}_{T}(x)\bigr{)}\bigr{]}=\bar{U}^{**}(x). However, any strategy yielding a terminal value W^T(x)\hat{W}_{T}(x), which has some support in (x,)(x_{*},\infty) is clearly not optimal by the strict concavity of the concavified utility function. Finally, a strategy that maximizes, not only the concavified problem, but also yields the same value for original problem, has to satisfy suppX^T(x)=suppW^T(x)={0,x}\operatorname{supp\,}\hat{X}_{T}(x)=\operatorname{supp\,}\hat{W}_{T}(x)=\{0,x_{*}\} since U¯<U¯\bar{U}<\bar{U}^{**} on (0,x)(0,x_{*}) by (25).

The existence of this strategy follows from a simple application of martingale representation theorem. Indeed, fix aa such that Φ(a/T)=xx\Phi(a/\sqrt{T})=\frac{x}{x_{*}}. Then, by the martingale representation theorem, the random variable x1l{WT<a}x_{*}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{W_{T}<a\}} has the representation x1l{WT<a}=x+0THs𝑑Wsx_{*}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{W_{T}<a\}}=x+\int_{0}^{T}H_{s}dW_{s}, where HL2([0,T]×Ω)H\in L^{2}([0,T]\times\Omega). Thus, x1l{WT<a}=x+0THsσSs𝑑Ssx_{*}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{W_{T}<a\}}=x+\int_{0}^{T}\frac{H_{s}}{\sigma S_{s}}dS_{s}. 111We thank an anonymous referee for pointing out this straightforward existence proof.

It turns out that one can easily explicitly construct the optimal strategy by using a strategy similar to the classical doubling strategy in the Black-Scholes model. However, contrary to the classical dubbling strategy our strategy will be admissible. Define the strategy Ht=1σStTtH_{t}=\frac{1}{\sigma S_{t}\sqrt{T-t}}, which gives rise to the value process

Xt1,H=x+0tdSsσSsTs=x+0tdWsTs.X_{t}^{1,H}=x+\int_{0}^{t}\frac{dS_{s}}{\sigma S_{s}\sqrt{T-s}}=x+\int_{0}^{t}\frac{dW_{s}}{\sqrt{T-s}}.

We note that X1,HX^{1,H} is a local martingale with quadratic variation process

X1,Ht=0tdsTs=logTTt,\Bigl{\langle}X^{1,H}\Bigr{\rangle}_{t}=\int_{0}^{t}\frac{ds}{T-s}=\log{\frac{T}{T-t}},

hence, it is a time changed Brownian motion Xt1,H=x+W~logTTtX_{t}^{1,H}=x+\tilde{W}_{\log{\frac{T}{T-t}}}. Defining now the stopping time τ:=inf{t0:Xt1,H[0,x]}\tau:=\inf\{t\geq 0\,:\,X_{t}^{1,H}\notin[0,x_{*}]\} we can see that we have for the stopped strategy Htτ=1σStτTtH_{t}^{\tau}=\frac{1}{\sigma S_{t}^{\tau}\sqrt{T-t}},

Xt1,Hτ=x+0t1l{sτ}dWsTs=x+W~logTTtτ.X_{t}^{1,H^{\tau}}=x+\int_{0}^{t}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{s\leq\tau\}}\frac{dW_{s}}{\sqrt{T-s}}=x+\tilde{W}_{\log{\frac{T}{T-t}}}^{\tau}.

Thus, the process X1,HτX^{1,H^{\tau}} hits either 0 or xx_{*} before time TT a.s. and the stopped process at terminal time, XT1,HτX_{T}^{1,H^{\tau}}, is hence almost surely concentrated on {0,x}\{0,x_{*}\}. Thus, HτH^{\tau} is indeed a strategy which yields the optimum.

This example, specifically the treatment of the case θ=0\theta=0 in section 3.2, reveals yet an other interesting fact. While the dual optimizer Y^T(y)=y\hat{Y}_{T}(y)=y is purely atomic for ever y>0y>0, for xxx\geq x_{*} it nevertheless follows that w(x)=U¯(x)=U¯(x)w(x)=\bar{U}^{**}(x)=\bar{U}(x) is reached also by the trivial strategy H0H\equiv 0. However, in this case, the solution of the concavified and the original problem coincide. This means that the condition of the atomlessness of the dual optimizer is not a necessary one, at least as one does not require an existence result which is independent of the initial capital.

4 Examples of Models in the Class of Itô Process

We now want to illustrate that Theorems 2.1 and 2.2 not only hold in Black-Scholes type markets, but also in many complete and incomplete markets, where the stock price process is given by an Itô process. First we will consider complete market models given by one-dimensional Itô processes and prove a general sufficient condition in terms of Malliavin differentiability, which can be applied, e.g., to local volatility models. Then we show that some classes of incomplete market models, such as the lognormal mixture models of Brigo and Mercurio Brigo , satisfy the conditions of Proposition 1. Additionally, we show that, under certain assumptions, stochastic volatility models satisfy directly the conditions of Theorem 2.2. We also provide examples of well-known models by Hull-White, Heston and Scott satisfying those assumptions.

Example 2

(One dimensional diffusion models): Let WW be a one-dimensional 222Generalization to the multi-dimensional case is straightforward. However, to make the exposition more tractable, we stay in one dimension. Brownian motion, defined on some probability space (Ω,,P)(\Omega,\mathcal{F},P). Denote by (tW)\bigl{(}\mathcal{F}_{t}^{W}\bigr{)} the filtration generated by the Brownian motion, augmented by all \mathbb{P}-negligible sets (as usual, we assume without loss of generality that TW=\mathcal{F}_{T}^{W}=\mathcal{F}). Additionally, let ([0,t])\mathcal{B}([0,t]) denote the Borel-σ\sigma-field on the interval [0,t][0,t]. Let the stock price process given by

dSt=μtStdt+σtStdWt,S0=s,dS_{t}=\mu_{t}S_{t}\,dt+\sigma_{t}S_{t}\,dW_{t},\qquad S_{0}=s, (26)

where μt\mu_{t} and σt\sigma_{t} are tW([0,t])\mathcal{F}_{t}^{W}\otimes\mathcal{B}([0,t])-progressive processes satisfying

𝔼[e20T|μt|𝑑t+e0Tσt2𝑑t]<andσ>0Pdt-a.e.\mathbb{E}\biggl{[}e^{2\int_{0}^{T}|\mu_{t}|\,dt}+e^{\int_{0}^{T}\sigma_{t}^{2}\,dt}\biggr{]}<\infty\qquad\mbox{and}\qquad\sigma>0\quad P\otimes dt\mbox{-a.e.}

In particular, we do not assume any Markovianity of the drift or diffusion coefficient. Moreover, let the money market account be given by

dBt=rtBtdt,B0=1dB_{t}=r_{t}B_{t}\,dt,\qquad B_{0}=1

for some progressive interest process rr satisfying 𝔼[e0T|rt|𝑑t]<\mathbb{E}\bigl{[}e^{\int_{0}^{T}|r_{t}|\,dt}\bigr{]}<\infty. Define the market price of risk θ\theta through

θtσt=μtrt.\theta_{t}\sigma_{t}=\mu_{t}-r_{t}.

To preclude arbitrage in the sense of a ’free lunch with vanishing risk’, we must assume that the market price of risk satisfies

𝔼[(0θtdWt1)T]=1,\mathbb{E}\Biggr{[}\mathcal{E}\biggl{(}-\int_{0}^{\bf\cdot}\theta_{t}\,dW^{1}_{t}\biggr{)}_{T}\Biggr{]}=1,

where (X)T:=exp(XT1/2XT)\mathcal{E}(X)_{T}:=\exp{\bigl{(}X_{T}-1/2\langle X\rangle_{T}\bigr{)}} denotes the stochastic (Doléans-Dade) exponential of the semimartingale XX. Additionally, for our results, we have to assume a little bit more regularity in terms of Malliavin differentiability (for a reference on Malliavin calculus see Nua , ENua ). We are in a one-dimensional stochastic volatility model. Hence, the underlying Hilbert space \mathcal{H} is given by L2([0,T];)L^{2}([0,T];\mathbb{R}), endowed with the canonical inner product. For p1p\geq 1 we denote by

𝔻1,p\displaystyle\mathbb{D}^{1,p} :={FLp(Ω,,P):F1,p:=(𝔼[|F|p]+𝔼[DFp])1p<}\displaystyle:=\biggl{\{}F\in L^{p}(\Omega,\mathcal{F},P)\,:\,\|F\|_{1,p}:=\Bigl{(}\mathbb{E}[|F|^{p}]+\mathbb{E}\bigl{[}\|DF\|_{\mathcal{H}}^{p}\bigr{]}\Bigr{)}^{\frac{1}{p}}<\infty\biggr{\}} (27)
={FLp(Ω,,P):F1,p:=(𝔼[|F|p]+𝔼[0T(DtF)p𝑑t])1p<}\displaystyle=\Biggl{\{}F\in L^{p}(\Omega,\mathcal{F},P)\,:\,\|F\|_{1,p}:=\biggl{(}\mathbb{E}[|F|^{p}]+\mathbb{E}\biggl{[}\int_{0}^{T}(D_{t}F)^{p}\,dt\biggr{]}\biggr{)}^{\frac{1}{p}}<\infty\Biggr{\}}

the subspace of LpL^{p} of random variables with pp-integrable Malliavin derivatives. We note that DtFD_{t}F denotes the the Malliavin derivative. Moreover, denote by 𝕃1,2\mathbb{L}^{1,2} the class of all processes uL2(Ω×[0,T])u\in L^{2}(\Omega\times[0,T]) such that ut𝔻1,2u_{t}\in\mathbb{D}^{1,2} for almost all tt such that there exists a measurable version of the two-parameter process DsutD_{s}u_{t} satisfying

𝔼0T0T(Dsut)2𝑑s𝑑t<.\mathbb{E}\int_{0}^{T}\int_{0}^{T}(D_{s}u_{t})^{2}\,ds\,dt<\infty.
Assumption 4

We assume that θ2𝕃1,2\theta^{2}\in\mathbb{L}^{1,2} and

𝔼[((0θtdWt1)T)2]<.\mathbb{E}\Biggr{[}\Biggl{(}\mathcal{E}\biggl{(}-\int_{0}^{\bf\cdot}\theta_{t}\,dW^{1}_{t}\biggr{)}_{T}\Biggr{)}^{2}\Biggr{]}<\infty. (28)

We note, in particular, that this Assumption is satisfied by local volatility models

dSt=μ(t,St)Stdt+σ(t,St)StdWt,S0=s,dS_{t}=\mu(t,S_{t})S_{t}\,dt+\sigma(t,S_{t})S_{t}\,dW_{t},\qquad S_{0}=s, (29)

as long as μ(t,s)\mu(t,s) and σ(t,s)\sigma(t,s) are nice enough. Specifically, a sufficient condition is that μ(t,s)\mu(t,s), μs(t,s)s\mu_{s}(t,s)s, σ(t,s)\sigma(t,s), and σs(t,s)s\sigma_{s}(t,s)s are bounded functions, uniformly continuous in the first component and twice continuously differentiable in the second component with bounded derivatives and that σ\sigma is uniformly bounded away from zero.

Recall that e\mathcal{M}^{e} denotes the set of all equivalent local martingale measures, in our current setting given by

e={Q:QP,B1S is a local -martingale}.\mathcal{M}^{e}=\bigl{\{}Q\,:\,Q\sim P,\,B^{-1}S\mbox{ is a local $\mathbb{Q}$-martingale}\bigr{\}}. (30)

Since we are in a complete market case, it is well-known that the set of all equivalent local martingale measures consists of a single measure

e={:dd|TW=ZT:=(0θt𝑑Wt1)T}.\mathcal{M}^{e}=\Biggl{\{}\mathbb{Q}\,:\,\left.\frac{d\mathbb{Q}}{d\mathbb{P}}\right|_{\mathcal{F}_{T}^{W}}=Z_{T}:=\mathcal{E}\Bigl{(}-\int_{0}^{\bf\cdot}\theta_{t}\,dW^{1}_{t}\Bigr{)}_{T}\Biggr{\}}. (31)
Lemma 1

If 0Tθt2𝑑t>0\int_{0}^{T}\theta_{t}^{2}\,dt>0 \mathbb{P}-a.e., then ZTZ_{T} has continuous law.

Proof

We will show more then required – asserting that under the stated conditions the random variable ZTZ_{T} has a density with respect to the Lebesgue measure. This will be done by using a Malliavin calculus-based result, which is due to Bouleau and Hirsch. For the logarithm LT:=logZtL_{T}:=\log{Z_{t}} we have

LT=0Tθs𝑑Ws120Tθs2𝑑s.L_{T}=-\int_{0}^{T}\theta_{s}\,dW_{s}-\frac{1}{2}\int_{0}^{T}\theta_{s}^{2}\,ds.

It follows that

𝔼[DLT2]=\displaystyle\mathbb{E}\Bigl{[}\|DL_{T}\|_{\mathcal{H}}^{2}\Bigr{]}= 𝔼[0T(θt+tTDtθs𝑑Ws+12tTDtθs2𝑑s)2𝑑t]\displaystyle\mathbb{E}\biggl{[}\int_{0}^{T}\biggl{(}\theta_{t}+\int_{t}^{T}D_{t}\theta_{s}\,dW_{s}+\frac{1}{2}\int_{t}^{T}D_{t}\theta_{s}^{2}\,ds\biggr{)}^{2}\,dt\biggr{]}
\displaystyle\leq 𝔼[0T3(θt2+(tTDtθs𝑑Ws)2+(12tTDtθs2𝑑s)2)𝑑t]\displaystyle\mathbb{E}\biggl{[}\int_{0}^{T}3\biggl{(}\theta_{t}^{2}+\Bigl{(}\int_{t}^{T}D_{t}\theta_{s}\,dW_{s}\Bigr{)}^{2}+\Bigl{(}\frac{1}{2}\int_{t}^{T}D_{t}\theta_{s}^{2}\,ds\Bigr{)}^{2}\biggr{)}\,dt\biggr{]}
=\displaystyle= 3𝔼[0Tθt2𝑑t]+3𝔼[0TtT(Dtθs)2𝑑s𝑑t]+34𝔼[0T(tTDtθs2𝑑s)2𝑑t]\displaystyle 3\mathbb{E}\biggl{[}\int_{0}^{T}\theta_{t}^{2}\,dt\biggr{]}+3\mathbb{E}\biggl{[}\int_{0}^{T}\int_{t}^{T}(D_{t}\theta_{s})^{2}\,ds\,dt\biggr{]}+\frac{3}{4}\mathbb{E}\biggl{[}\int_{0}^{T}\biggl{(}\int_{t}^{T}D_{t}\theta_{s}^{2}\,ds\biggr{)}^{2}\,dt\biggr{]}
\displaystyle\leq 30T𝔼[θt2]𝑑t+30T𝔼[Dθs2]𝑑s+34T20T𝔼[Dθs22]𝑑s\displaystyle 3\int_{0}^{T}\mathbb{E}[\theta_{t}^{2}]\,dt+3\int_{0}^{T}\mathbb{E}\Bigl{[}\|D\theta_{s}\|_{\mathcal{H}}^{2}\Bigr{]}\,ds+\frac{3}{4}T^{2}\int_{0}^{T}\mathbb{E}\Bigl{[}\|D\theta_{s}^{2}\|_{\mathcal{H}}^{2}\Bigr{]}\,ds
\displaystyle\leq 30Tθs1,22𝑑s+34T20Tθs21,22𝑑s30T(1T+T2θs21,2)2𝑑s<\displaystyle 3\int_{0}^{T}\|\theta_{s}\|_{1,2}^{2}\,ds+\frac{3}{4}T^{2}\int_{0}^{T}\|\theta_{s}^{2}\|_{1,2}^{2}\,ds\leq 3\int_{0}^{T}\biggl{(}\frac{1}{T}+\frac{T}{2}\|\theta_{s}^{2}\|_{1,2}\biggr{)}^{2}\,ds<\infty

by Assumption 4. Hence,

ZT1,1\displaystyle\|Z_{T}\|_{1,1} =𝔼[ZT]+𝔼[DZT]=𝔼[ZT]+𝔼[ZTDLT]=𝔼[[ZT(1+DLT)]\displaystyle=\mathbb{E}\bigl{[}Z_{T}\bigr{]}+\mathbb{E}\Bigl{[}\|DZ_{T}\|_{\mathcal{H}}\Bigr{]}=\mathbb{E}\bigl{[}Z_{T}\bigr{]}+\mathbb{E}\Bigl{[}Z_{T}\|DL_{T}\|_{\mathcal{H}}\Bigr{]}=\mathbb{E}\Bigl{[}[Z_{T}\bigl{(}1+\|DL_{T}\|_{\mathcal{H}}\bigr{)}\Bigr{]}
𝔼[ZT2]𝔼[(1+DLT)2]𝔼[ZT2]2𝔼[1+DLT2]<.\displaystyle\leq\sqrt{\mathbb{E}\bigl{[}Z_{T}^{2}\bigr{]}}\sqrt{\mathbb{E}\Bigl{[}\bigl{(}1+\|DL_{T}\|_{\mathcal{H}}\bigr{)}^{2}\Bigr{]}}\leq\sqrt{\mathbb{E}\bigl{[}Z_{T}^{2}\bigr{]}}\sqrt{2\mathbb{E}\Bigl{[}1+\|DL_{T}\|_{\mathcal{H}}^{2}\Bigr{]}}<\infty.

Following the criterium for absolute continuity (cf. (Nua, , Theorem 2.1.3)), it is therefore enough to show that

DZT>0-a.s.\|DZ_{T}\|_{\mathcal{H}}>0\qquad\mathbb{P}\mbox{-a.s.}

From

(DZT)2=ZT2(DLT)2(DZ_{T})^{2}=Z_{T}^{2}(DL_{T})^{2}

and from the fact that ZT>0Z_{T}>0 \mathbb{P}-a.s., this is equivalent to the fact that DLT>0\|DL_{T}\|_{\mathcal{H}}>0~\mathbb{P}-a.s. However, for every adapted process Ydom(δ)L2(Ω;)Y\in\operatorname{dom\,}{(\delta)}\subseteq L^{2}(\Omega;\mathcal{H}), the domain of the Skorohod integral, we have, by the definition of LTL_{T}

𝔼[Yt,DtLT]\displaystyle\mathbb{E}\Bigl{[}\bigl{\langle}Y_{t}\,,\,D_{t}L_{T}\bigr{\rangle}_{\mathcal{H}}\Bigr{]} =𝔼[LTδ(Yt)]=𝔼[LT(0TYt2𝑑Wt)]=𝔼[0θt𝑑Wt,0Yt𝑑WtT]\displaystyle=\mathbb{E}\Bigl{[}L_{T}\,\delta(Y_{t})\Bigr{]}=\mathbb{E}\biggl{[}L_{T}\biggl{(}\int_{0}^{T}Y_{t}^{2}\,dW_{t}\biggr{)}\biggr{]}=\mathbb{E}\biggl{[}\Bigl{\langle}-\int_{0}^{\cdot}\theta_{t}\,dW_{t},\int_{0}^{\cdot}Y_{t}\,dW_{t}\Bigr{\rangle}_{T}\biggr{]}
=𝔼[0θt𝑑Wt,0Yt𝑑WtT]=𝔼[0TθtYt𝑑t].\displaystyle=\mathbb{E}\biggl{[}\Bigl{\langle}-\int_{0}^{\cdot}\theta_{t}\,dW_{t},\int_{0}^{\cdot}Y_{t}\,dW_{t}\Bigr{\rangle}_{T}\biggr{]}=\mathbb{E}\biggl{[}-\int_{0}^{T}\theta_{t}Y_{t}\,dt\biggr{]}.

Thus, we conclude that DLT=0\|DL_{T}\|_{\mathcal{H}}=0~\mathbb{P}-a.s., if only if 0Tθt2𝑑t=0\int_{0}^{T}\theta_{t}^{2}\,dt=0~\mathbb{P}-a.s.∎

We turn our attention now to incomplete market models:

Example 3

(Lognormal mixture models): Similar to Example 2 let WW be a one-dimensional Brownian motion defined on some probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) and denote by (tW)\bigl{(}\mathcal{F}_{t}^{W}\bigr{)} the filtration generated by it, augmented by all \mathbb{P}-negligible sets. Let NN\in\mathbb{N} and consider a random variable XX on {1,N}\{1,\ldots N\} such that ν[X=i]=pi,i=1,,N\nu\bigl{[}X=i\bigr{]}=p_{i},~i=1,\ldots,N, for some counting measure ν\nu where pi>0p_{i}>0 and i=1Npi=1\sum_{i=1}^{N}p_{i}=1. Let the stock price process modeled on the space (Ω~,~,~)=(Ω×{1,,N},𝒫({1,,N}),ν)(\tilde{\Omega},\tilde{\mathcal{F}},\tilde{\mathbb{P}})=(\Omega\times\{1,\ldots,N\},\mathcal{F}\otimes\mathcal{P}(\{1,\ldots,N\}),\mathbb{P}\otimes\nu) by

dSt=μtStdt+σt(X)StdWt,S0=s,dS_{t}=\mu_{t}S_{t}\,dt+\sigma_{t}(X)S_{t}\,dW_{t},\qquad S_{0}=s, (32)

where μt\mu_{t} and σt(i),i=1,,N\sigma_{t}(i),~i=1,\ldots,N are deterministic functions, bounded and bounded away from zero, and satisfying σt(i)=σ0\sigma_{t}(i)=\sigma_{0} for t[0,t0]t\in[0,t_{0}] for some t0>0t_{0}>0, and σt(i)σt(j),ij\sigma_{t}(i)\neq\sigma_{t}(j),~i\neq j for all t(t0,t1)t\in(t_{0},t_{1}), for some t1>t0t_{1}>t_{0}, satisfying t1Tt_{1}\leq T. We note that the filtration tS\mathcal{F}_{t}^{S} generated by the stock price is not right continuous at t0t_{0}, and it agrees with the filtration generated by XX and tW\mathcal{F}^{W}_{t} only for t>t0t>t_{0}, whereas it is strictly smaller at tt0t\leq t_{0}. Following Brigo and Mercurio (Brigo, , Section 10.4) this SDE has a unique strong solution.

Let the money market account be given by

dBt=rtBtdt,B0=1,dB_{t}=r_{t}B_{t}\,dt,\qquad B_{0}=1,

for some bounded progressive interest process rr and, conditioned on X=iX=i, define the market price of risk θ\theta through

θi=μtrtσt(i).\theta_{i}=\frac{\mu_{t}-r_{t}}{\sigma_{t}(i)}.

Assume also that the {θi}i=1n\left\{\theta_{i}\right\}_{i=1}^{n} are linearly independent.

Lemma 2

The the set of all equivalent local martingale measures e\mathcal{M}^{e} is given by

e={~:d~d~|TS=ZTq:=i=1Nqipi(0θi𝑑Wt)T,q𝒦},\mathcal{M}^{e}=\Biggl{\{}\tilde{\mathbb{Q}}\,:\,\left.\frac{d\tilde{\mathbb{Q}}}{d\tilde{\mathbb{P}}}\right|_{\mathcal{F}_{T}^{S}}=Z_{T}^{{{\overrightarrow{q}}}}:=\sum_{i=1}^{N}\frac{q_{i}}{p_{i}}\mathcal{E}\Bigl{(}-\int_{0}^{\bf\cdot}\theta_{i}\,dW_{t}\Bigr{)}_{T},~{\overrightarrow{q}}\in\mathcal{K}\Biggr{\}}, (33)

where 𝒦:={q>0N:i=1Nqi=1}\mathcal{K}:=\left\{{\overrightarrow{q}}\in\mathbb{R}^{N}_{>0}\colon~\sum_{i=1}^{N}q_{i}=1\right\}. If 0Tθi(t)2𝑑t>0\int_{0}^{T}\theta_{i}(t)^{2}\,dt>0 \mathbb{P}-a.e. for every i=1,,Ni=1,\ldots,N, then the set {ZTq:q𝒦}\bigl{\{}Z_{T}^{{\overrightarrow{q}}}\,:\,{\overrightarrow{q}}\in\mathcal{K}\bigr{\}} has uniformly absolutely continuous distributions.

Proof

Note first that from (Amen, , Theorem 4.1) it follows that any integrable random variable ZT=d~d~|TSZ_{T}=\left.\frac{d\tilde{\mathbb{Q}}}{d\tilde{\mathbb{P}}}\right|_{\mathcal{F}_{T}^{S}} in the filtration S\mathcal{F}^{S} has the representation

ZT=Zt0++t0Tηs𝑑WsZ_{T}=Z_{t_{0}+}+\int_{t_{0}}^{T}\eta_{s}\,dW_{s} (34)

where Zt0+t0+SZ_{t_{0}+}\in\mathcal{F}_{t_{0}+}^{S} and ηs\eta_{s} an S\mathcal{F}^{S}-predictable and locally integrable process. More precisely, setting t=t+t0+ε\mathcal{F}^{\prime}_{t}=\mathcal{F}_{t+t_{0}+\varepsilon} and applying (Amen, , Theorem 4.1) yields ZT=Zt0+ε+t0+εTηs𝑑WsZ_{T}=Z_{t_{0}+\varepsilon}+\int_{t_{0}+\varepsilon}^{T}\eta_{s}\,dW_{s} and sending ε\varepsilon to zero this converges to (34) as the stochastic integrals are consistently constructed and Zt0+ε=𝔼[ZT|t0+εS]Z_{t_{0}+\varepsilon}=\mathbb{E}[Z_{T}\,|\,\mathcal{F}^{S}_{t_{0}+\varepsilon}] is a backward martingale that converges by the backward martingale convergence theorem almost surely to Zt0+:=𝔼[ZT|t0+S]Z_{t_{0}+}:=\mathbb{E}[Z_{T}\,|\,\mathcal{F}^{S}_{t_{0}+}]. Moreover, using the classical martingale representation theorem this yields

ZT\displaystyle Z_{T} =ΔZt0+Zt0+t0Tηs𝑑Ws=ΔZt0+Z0+0t0ηs𝑑Ws+t0Tηs𝑑Ws\displaystyle=\Delta Z_{t_{0}}+Z_{t_{0}}+\int_{t_{0}}^{T}\eta_{s}\,dW_{s}=\Delta Z_{t_{0}}+Z_{0}+\int_{0}^{t_{0}}\eta_{s}\,dW_{s}+\int_{t_{0}}^{T}\eta_{s}\,dW_{s}
=1+ΔZt0+0Tηs𝑑Ws\displaystyle=1+\Delta Z_{t_{0}}+\int_{0}^{T}\eta_{s}\,dW_{s}

with ηs\eta_{s} an S\mathcal{F}^{S}-predictable and locally integrable process and ΔZt0=Zt0+Zt0\Delta Z_{t_{0}}=Z_{t_{0}+}-Z_{t_{0}}. Applying Itô’s formula to logZt\log{Z_{t}} we conclude that

Zt=(1+ΔZt0Zt01l{t>t0}(s))exp(0tηsZs𝑑Ws120t(ηsZs)2𝑑s)Z_{t}=\left(1+\frac{\Delta Z_{t_{0}}}{Z_{t_{0}}}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{t>t_{0}\}}(s)\right)\exp{\Biggl{(}{\int_{0}^{t}\frac{\eta_{s}}{Z_{s}}\,dW_{s}-\frac{1}{2}\int_{0}^{t}\biggl{(}\frac{\eta_{s}}{Z_{s}}\biggr{)}^{2}\,ds\Biggr{)}}}

The no arbitrage condition requires that the discounted stock price is a local martingale under ~\tilde{\mathbb{Q}} or, equivalently, that Bt1StZtB_{t}^{-1}S_{t}Z_{t} is a local ~\tilde{\mathbb{P}}-martingale. Noting that

Bt1St\displaystyle B_{t}^{-1}S_{t} =S0+0tσs(i)Bs1Ss𝑑Ws+0t(μsrs)Bs1Ss𝑑t\displaystyle=S_{0}+\int_{0}^{t}\sigma_{s}(i)B_{s}^{-1}S_{s}\,dW_{s}+\int_{0}^{t}(\mu_{s}-r_{s})B_{s}^{-1}S_{s}\,dt
=S0exp(0tσs(i)𝑑Wt+0t(μsrs12σs2(i))𝑑s)\displaystyle=S_{0}\exp{\biggl{(}\int_{0}^{t}\sigma_{s}(i)\,dW_{t}+\int_{0}^{t}\Bigl{(}\mu_{s}-r_{s}-\frac{1}{2}\sigma_{s}^{2}(i)\Bigr{)}\,ds\biggr{)}}

implies

Bt1StZt\displaystyle B_{t}^{-1}S_{t}Z_{t} =S0(1+ΔZt0Zt01l{t>t0}(s))\displaystyle=S_{0}\cdot\left(1+\frac{\Delta Z_{t_{0}}}{Z_{t_{0}}}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{t>t_{0}\}}(s)\right)
exp(0t(σs(i)+ηsZs)dWt+0t(μsrs12σs2(i)12(ηsZs)2ds)\displaystyle\phantom{=}\cdot\exp{\Biggl{(}\int_{0}^{t}\Bigl{(}\sigma_{s}(i)+\frac{\eta_{s}}{Z_{s}}\Bigr{)}\,dW_{t}+\int_{0}^{t}\Bigl{(}\mu_{s}-r_{s}-\frac{1}{2}\sigma_{s}^{2}(i)-\frac{1}{2}\biggr{(}\frac{\eta_{s}}{Z_{s}}\biggr{)}^{2}\,ds\Biggr{)}} (35)

we have just to check under which conditions (3) is a ~\tilde{\mathbb{P}}-martingale. Note that it is enough to check when the exponential is a local martingale (checking that for 𝔼[Bt1StZt|sS]\mathbb{E}[B_{t}^{-1}S_{t}Z_{t}\,|\mathcal{F}^{S}_{s}] for ts>t0t\geq s>t_{0} and t0tst_{0}\geq t\geq s is straightforward, at t0t_{0} this follows from the continuity of the exponential). Calculating the quadratic variation of the stochastic integral and comparing it with the determinist one yields

ηsZs=rsμsσs(i)=θi(s).\frac{\eta_{s}}{Z_{s}}=\frac{r_{s}-\mu_{s}}{\sigma_{s}(i)}=\theta_{i}(s).

Moreover, as 1+ΔZt0Zt0=Zt0+Zt0σ(X)1+\frac{\Delta Z_{t_{0}}}{Z_{t_{0}}}=\frac{Z_{t_{0}+}}{Z_{t_{0}}}\in\sigma(X) is only supported on {1,N}\{1,\ldots N\} and the martingale condition forces 𝔼[1+ΔZt0Zt0|t0S]=1\mathbb{E}\bigl{[}1+\frac{\Delta Z_{t_{0}}}{Z_{t_{0}}}\,\big{|}\mathcal{F}_{t_{0}}^{S}\bigr{]}=1, we get that the representation (33) is a necessary condition on equivalent local martingale measures. It is straightforward to check that it is also sufficient.

Additionally, we have that all θi,i=1,,N\theta_{i},~i=1,\ldots,N are uniformly bounded. It follows that the densities of the normally distributed random variables 0Tθi𝑑Wt\int_{0}^{T}\theta_{i}\,dW_{t} are also uniformly bounded. Furthermore, the Gramian matrix G(θ1,,θN)G(\theta_{1},\ldots,\theta_{N}) with elements Gij=0Tθiθj𝑑t,1i,jNG_{ij}=\int_{0}^{T}\theta_{i}\theta_{j}\,dt,~1\leq i,j\leq N, has non-zero Gram determinant, by our assumption that {θi}i=1N\left\{\theta_{i}\right\}_{i=1}^{N} are linearly independent. It follows that the random vector

ν:=(0Tθ1𝑑Wt,,0TθN𝑑Wt){\overrightarrow{\nu}}:=\left(\int_{0}^{T}\theta_{1}\,dW_{t},\ldots,\int_{0}^{T}\theta_{N}\,dW_{t}\right)

is normally distributed with mean zero, and variance G(θ1,,θN)G(\theta_{1},\ldots,\theta_{N}). Hence, it has a bounded density. Finally, note that for q𝒦{\overrightarrow{q}}\in\mathcal{K} the random variable

ZTq:=i=1Nqipi(0θi𝑑Wt)TZ_{T}^{{{\overrightarrow{q}}}}:=\sum_{i=1}^{N}\frac{q_{i}}{p_{i}}\mathcal{E}\Bigl{(}-\int_{0}^{\bf\cdot}\theta_{i}\,dW_{t}\Bigr{)}_{T}

is an exponential transformation of ν{\overrightarrow{\nu}}, since the functions θi,i=1,,N\theta_{i},~i=1,\ldots,N are deterministic. Therefore, it has a density, too. This transformation is continuous as a function of q{\overrightarrow{q}} and, thus, so is the density of ZTqZ_{T}^{{{\overrightarrow{q}}}}. It follows that the density of ZTqZ_{T}^{{{\overrightarrow{q}}}} is uniformly bounded for any q{\overrightarrow{q}} in the compact closure 𝒦¯\overline{\mathcal{K}}.∎

As the family {ZTq:q>0N,i=1Nqi=1}\bigl{\{}Z_{T}^{{\overrightarrow{q}}}\,:\,{\overrightarrow{q}}\in\mathbb{R}^{N}_{>0},~\sum_{i=1}^{N}q_{i}=1\bigr{\}} has uniformly absolutely continuous distributions, we can then apply Proposition 1 to establish the distributional continuity of the dual optimizer.

Stochastic Volatility Examples

We now generalize the setting of Example 2 to encompass stochastic volatility models, dropping the assumption of market completeness. Namely, let W1W^{1} and W2W^{2} be two independent one-dimensional Brownian motions (the generalization to the multi-dimensional case is again straightforward) defined on some probability space (Ω,,P)(\Omega,\mathcal{F},P) and denote by (tW1,W2)\bigl{(}\mathcal{F}_{t}^{W^{1},W^{2}}\bigr{)} the filtration generated by them, augmented by all \mathbb{P}-negligible sets (as usual we assume without loss of generality that TW1,W2=\mathcal{F}_{T}^{W^{1},W^{2}}=\mathcal{F}). Let the stock price process given by

dSt=μtStdt+σtStdWt1,S0=s,dS_{t}=\mu_{t}S_{t}\,dt+\sigma_{t}S_{t}\,dW^{1}_{t},\qquad S_{0}=s,

where μt\mu_{t} and σt\sigma_{t} are tW1,W2([0,t])\mathcal{F}_{t}^{W^{1},W^{2}}\otimes\mathcal{B}([0,t])-progressive processes satisfying

𝔼[e20T|μt|𝑑t+e0Tσt2𝑑t]<andσ>0Pdt-a.e.\mathbb{E}\biggl{[}e^{2\int_{0}^{T}|\mu_{t}|\,dt}+e^{\int_{0}^{T}\sigma_{t}^{2}\,dt}\biggr{]}<\infty\qquad\mbox{and}\qquad\sigma>0\quad P\otimes dt\mbox{-a.e.}

In particular we again do not assume any Markovianity of the drift or diffusion coefficient. Moreover, we still assume that the money market account be given by

dBt=rtBtdt,B0=1dB_{t}=r_{t}B_{t}\,dt,\qquad B_{0}=1

for some progressive interest process rr satisfying 𝔼[e0T|rt|𝑑t]<\mathbb{E}\bigl{[}e^{\int_{0}^{T}|r_{t}|\,dt}\bigr{]}<\infty and define the market price of risk θ\theta through

θtσt=μtrt.\theta_{t}\sigma_{t}=\mu_{t}-r_{t}.

We readily adjust all of the remaining definitions of Example 2 to this framework. We note that, in this case, \mathcal{H} will be given by L2([0,T];2)L^{2}([0,T];\mathbb{R}^{2}). The definition of 𝔻1,p\mathbb{D}^{1,p} in (27) will not change. Finally, we will adjust the definition 𝕃1,2\mathbb{L}^{1,2} to be the class of all processes uL2(Ω×[0,T])u\in L^{2}(\Omega\times[0,T]) such that ut𝔻1,2u_{t}\in\mathbb{D}^{1,2} for almost all tt such that there exists a measurable version of the two-parameter process DsutD_{s}u_{t} satisfying

𝔼0T0T(Ds1ut)2+(Ds2ut)2dsdt<.\mathbb{E}\int_{0}^{T}\int_{0}^{T}(D_{s}^{1}u_{t})^{2}+(D_{s}^{2}u_{t})^{2}\,ds\,dt<\infty.

We, of course, still preclude arbitrage in the sense of a ’free lunch with vanishing risk’ by assuming that the market price of risk satisfies

𝔼[(0θtdWt1)T]=1,\mathbb{E}\Biggr{[}\mathcal{E}\biggl{(}-\int_{0}^{\bf\cdot}\theta_{t}\,dW^{1}_{t}\biggr{)}_{T}\Biggr{]}=1,

and we still assume Assumption 4 is satisfied.

Recall that e\mathcal{M}^{e}, given by (30), denotes the set of all equivalent local martingale measures. The first major change, in comparison to Example 2, is that, in our current setting, e\mathcal{M}^{e} is given by

Lemma 3

The set of all equivalent local martingale measures can be characterized as

e\displaystyle\mathcal{M}^{e} ={:dd|TW1,W2=ZTθ(λ):=(0θtdWt1+0λtdWt2)T,.\displaystyle=\Biggl{\{}\mathbb{Q}\,:\,\left.\frac{d\mathbb{Q}}{d\mathbb{P}}\right|_{\mathcal{F}_{T}^{W^{1},W^{2}}}=Z_{T}^{\theta}(\lambda):=\mathcal{E}\Bigl{(}-\int_{0}^{\bf\cdot}\theta_{t}\,dW^{1}_{t}+\int_{0}^{\bf\cdot}\lambda_{t}\,dW^{2}_{t}\Bigr{)}_{T},\Biggr{.}
.λΛ,𝔼[ZTθ(λ)]=1},\displaystyle\phantom{=\Biggl{\{}\,}\Biggl{.}\lambda\in\Lambda,\,\mathbb{E}[Z_{T}^{\theta}(\lambda)]=1\Biggr{\}}, (36)

where

Λ:={λ predictable, such that 0Tλt2𝑑t<-a.s.}.\Lambda:=\Bigl{\{}\lambda\mbox{ predictable, such that }\int_{0}^{T}\lambda_{t}^{2}\,dt<\infty\,\mathbb{P}\mbox{-a.s.}\Bigr{\}}.

Moreover, if 0Tθt2𝑑t>0\int_{0}^{T}\theta_{t}^{2}\,dt>0 \mathbb{P}-a.e., then the random variable ZTθ(0)Z_{T}^{\theta}(0) has a continuous law.

Proof

To begin, we prove the characterization of the set of equivalent local martingale measures, which follows the proof for the classical Markovian case (cf. Frey ). First, it is clear that, only under the condition 𝔼[ZTθ(λ)]=1\mathbb{E}[Z_{T}^{\theta}(\lambda)]=1, will the new measure \mathbb{Q} be a probability measure. By the martingale representation theorem, we know that we can find predictable processes η\eta, ξ\xi such that

Ztθ(λ)=dd|tW1,W2=1+0tηs𝑑Ws1+0tξs𝑑Ws2.Z_{t}^{\theta}(\lambda)=\left.\frac{d\mathbb{Q}}{d\mathbb{P}}\right|_{\mathcal{F}_{t}^{W^{1},W^{2}}}=1+\int_{0}^{t}\eta_{s}\,dW^{1}_{s}+\int_{0}^{t}\xi_{s}\,dW^{2}_{s}.

Since \mathbb{Q} and \mathbb{P} are equivalent, the density process is strictly positive and we can define its logarithm Ltθ(λ)=logZtθ(λ)L_{t}^{\theta}(\lambda)=\log{Z_{t}^{\theta}(\lambda)} which satisfies, by Itô’s formula

dLtθ(λ)=ηtZtθ(λ)dWt1+ξtZtθ(λ)dWt212((ηtZtθ(λ))2+(ξtZtθ(λ))2)dt.dL_{t}^{\theta}(\lambda)=\frac{\eta_{t}}{Z_{t}^{\theta}(\lambda)}\,dW^{1}_{t}+\frac{\xi_{t}}{Z_{t}^{\theta}(\lambda)}\,dW^{2}_{t}-\frac{1}{2}\Biggl{(}\biggl{(}\frac{\eta_{t}}{Z_{t}^{\theta}(\lambda)}\biggr{)}^{2}+\biggl{(}\frac{\xi_{t}}{Z_{t}^{\theta}(\lambda)}\biggr{)}^{2}\Biggr{)}\,dt.

Expressing now the stock-price process SS under \mathbb{Q}, we get, by Girsanov’s theorem

dBt1St\displaystyle dB_{t}^{-1}S_{t} =Bt1(σtStdWt1+(μtrt)Stdt)\displaystyle=B_{t}^{-1}\Bigl{(}\sigma_{t}S_{t}dW^{1}_{t}+(\mu_{t}-r_{t})S_{t}\,dt\Bigr{)}
=Bt1(σtStdWt+St(μtrt+σtηtZsθ(λ))dt)\displaystyle=B_{t}^{-1}\Biggl{(}\sigma_{t}S_{t}dW_{t}^{\mathbb{Q}}+S_{t}\biggl{(}\mu_{t}-r_{t}+\sigma_{t}\frac{\eta_{t}}{Z_{s}^{\theta}(\lambda)}\biggr{)}\,dt\Biggr{)} (37)

for some \mathbb{Q}-Brownian motion WW^{\mathbb{Q}} independent of W2W^{2}. Hence, the discounted stock price is a local martingale only if

ηZθ(λ)=θandξZθ(λ)=λ\frac{\eta}{Z^{\theta}(\lambda)}=-\theta\qquad\mbox{and}\qquad\frac{\xi}{Z^{\theta}(\lambda)}=\lambda

for some predictable, square-integrable process λ\lambda. On the other hand, every expression on the right hand side of (3) defines an equivalent probability measure. By (4), the stock price is a local martingale under this measure.

The second part of the assertion – that ZTθ(0)Z_{T}^{\theta}(0) has a continuous law – is a direct consequence of Lemma 1.∎

Next, we would like to apply Proposition 1 to conclude that Y^T(y)\hat{Y}_{T}(y) has a continuous law. But, this is not easy to do, since we need to satisfy the assumption of Proposition 1 that the family ZTθ(λ)Z_{T}^{\theta}(\lambda) is uniformly absolutely continuous with respect to the Lebesgue measure. Here, we propose an alternative approach, which requires a certain restriction, namely that the market price of risk does not depend on the Brownian motion driving the stochastic volatility. The classical example of this would be a constant market price of risk. While this is a shortcoming from a theoretical point of view, it does not matter much from a practical perspective. The straight-forward idea of using a constant drift does not work out nicely in many situations, as there may not exist an equivalent change of measure (e.g., in the Stein & Stein model or in the Heston model without Feller condition - c.f. HW ). A standard procedure (compare the discussion in (FPSS, , Section 2.4.2)) is exactly to assume a constant market price of risk to avoid these integrability issues.

Lemma 4

Assume that the market price of risk θttW1\theta_{t}\in\mathcal{F}_{t}^{W^{1}}. Then the infimum over all equivalent local martingale measures in the calculation of the value function of the dual problem vv is reached for λ=0\lambda=0, i.e.,

v(y)=infe𝔼[U¯(ydd)]=𝔼[U¯(yZTθ(0))].v(y)=\inf_{\mathbb{Q}\in\mathcal{M}^{e}}\mathbb{E}\biggl{[}\bar{U}^{*}\biggl{(}y\frac{d\mathbb{Q}}{d\mathbb{P}}\biggr{)}\biggr{]}=\mathbb{E}\biggl{[}\bar{U}^{*}\biggl{(}yZ_{T}^{\theta}(0)\biggr{)}\biggr{]}.
Proof

From Jensen’s inequality it follows that for any λΛ\lambda\in\Lambda

𝔼[U¯(yZTθ(λ))]=𝔼[𝔼[U¯(y(0θt𝑑Wt1)T(0λt𝑑Wt2)T)|W1]]\displaystyle\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}yZ_{T}^{\theta}(\lambda)\bigr{)}\Bigr{]}=\mathbb{E}\Biggl{[}\left.\mathbb{E}\biggl{[}\bar{U}^{*}\biggl{(}y\mathcal{E}\Bigl{(}-\int_{0}^{\bf\cdot}\theta_{t}\,dW^{1}_{t}\Bigr{)}_{T}\mathcal{E}\Bigl{(}\int_{0}^{\bf\cdot}\lambda_{t}\,dW^{2}_{t}\Bigr{)}_{T}\biggr{)}\,\right|\,\mathcal{F}_{\bf\cdot}^{W^{1}}\biggr{]}\Biggr{]}
𝔼[U¯(y(0θtdWt1)T𝔼[(0λtdWt2)T|W1])].\displaystyle\geq\mathbb{E}\Biggl{[}\bar{U}^{*}\Biggl{(}y\mathcal{E}\Bigl{(}-\int_{0}^{\bf\cdot}\theta_{t}\,dW^{1}_{t}\Bigr{)}_{T}\mathbb{E}\biggl{[}\mathcal{E}\Bigr{(}\int_{0}^{\bf\cdot}\lambda_{t}\,dW^{2}_{t}\Bigr{)}_{T}\Big{|}\mathcal{F}_{\bf\cdot}^{W^{1}}\biggr{]}\Biggr{)}\Biggr{]}.

Since W1W^{1} and W2W^{2} are independent, 𝔼[(0λt𝑑Wt2)T|W1]=1\mathbb{E}\bigl{[}\mathcal{E}\bigl{(}\int_{0}^{\bf\cdot}\lambda_{t}\,dW^{2}_{t}\bigr{)}_{T}\big{|}\mathcal{F}_{\bf\cdot}^{W^{1}}\bigr{]}=1 a.s., and we note that

𝔼[U¯(yZTθ(λ))]𝔼[U¯(y(0θt𝑑Wt1)T)]=𝔼[U¯(yZTθ(0))].\displaystyle\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}yZ_{T}^{\theta}(\lambda)\bigr{)}\Bigr{]}\geq\mathbb{E}\biggl{[}\bar{U}^{*}\biggl{(}y\mathcal{E}\Bigl{(}-\int_{0}^{\bf\cdot}\theta_{t}\,dW^{1}_{t}\Bigr{)}_{T}\biggr{)}\biggr{]}=\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}yZ_{T}^{\theta}(0)\bigr{)}\Bigr{]}.

Since, of course, v(y)𝔼[U¯(yZTθ(0))]v(y)\leq\mathbb{E}\bigl{[}\bar{U}^{*}\bigl{(}yZ_{T}^{\theta}(0)\bigr{)}\Bigr{]}, we conclude that

v(y)\displaystyle v(y) =𝔼[U¯(yZTθ(0))].\displaystyle=\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}yZ_{T}^{\theta}(0)\bigr{)}\Bigr{]}.

Thus, we can use Lemma 1 and apply Theorem 2.2 directly:

Theorem 4.1

Assume that for a stochastic volatility model Assumption 4 holds. Additionally, assume that θttW1\theta_{t}\in\mathcal{F}_{t}^{W_{1}}, and 0Tθt2𝑑t>0\int_{0}^{T}\theta_{t}^{2}\,dt>0~\mathbb{P}-a.s. Then the original problem (3) has a maximizer, which is also a maximizer of the concavified problem (6).

Finally, we want to show that the Assumption 4 is satisfied in many standard volatility models. First we remark that if the volatility function σt=σ(Yt)\sigma_{t}=\sigma(Y_{t}) is a smooth function in YtY_{t}, bounded and bounded away from zero, and the volatility process satisfies Y𝕃1,2Y\in\mathbb{L}^{1,2}, then the assumption is satisfied. This is also enough to ensure the existence of an equivalent local martingale measure. Turning to more standard models, we observe that it can be shown that many standard volatility processes, such as e.g., Ornstein-Uhlenbeck, CIR or geometric Brownian motion, satisfy the Malliavin differentiability condition (for the CIR process, at least in the nice regime when Feller’s condition holds – cf. AlosEwald ). Thus, all possible problems arise, not from the Malliavin smoothness condition, but from the requirement that θttW1\theta_{t}\in\mathcal{F}_{t}^{W^{1}}, which is usually not satisfied for constant drift. As mentioned above, the standard way to circumvent this problem is to allow for a volatility-dependent excess appreciation:

Example 4

(Correlated Hull-White model): We consider a bond with constant interest rate rr and the stock price given by

dSt\displaystyle dS_{t} =(r+Ytf(Wt1))Stdt+YtStdWt1,S0=s,\displaystyle=\Bigl{(}r+Y_{t}f(W_{t}^{1})\Bigr{)}S_{t}\,dt+Y_{t}S_{t}\,dW_{t}^{1},\quad S_{0}=s,
dYt\displaystyle dY_{t} =bYtdt+ϱaYtdWt1+1ϱ2aYtdWt2,Y0=y,\displaystyle=bY_{t}\,dt+\varrho aY_{t}\,dW_{t}^{1}+\sqrt{1-\varrho^{2}}aY_{t}\,dW_{t}^{2},\quad Y_{0}=y,

for constants, bb\in\mathbb{R}, aa, ss, y>0y>0 and ϱ(1,1)\varrho\in(-1,1) and independent Brownian motions W1W^{1}, W2W^{2}. Moreover we assume that the excess appreciation rate f(Wt1)f(W_{t}^{1}) is given via a bounded C1(0)C^{1}({\mathbb{R}_{\geq 0}})-function ff with bounded derivative, and that it is not identically zero. This guarantees that the market price of risk θt=f(Wt1)\theta_{t}=f(W_{t}^{1}) remains bounded and ensures that the integrability condition of Assumption 4 is satisfied.

Calculating the Malliavin derivative of the θ2\theta^{2}

Dtθs2=Dt(f(Ws1)2)=2f(Ws1)DtWs11l[t,T](s)=(2f(Ws1)1l[t,T](s)0).D_{t}\theta_{s}^{2}=D_{t}\bigl{(}f(W_{s}^{1})^{2}\bigr{)}=2f^{\prime}(W_{s}^{1})D_{t}W^{1}_{s}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{[t,T]}(s)=\begin{pmatrix}2f^{\prime}(W_{s}^{1}){\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{[t,T]}(s)\\ 0\end{pmatrix}.

We can conclude that θ2𝕃1,2\theta^{2}\in\mathbb{L}^{1,2} since ff and ff^{\prime} are bounded. Moreover, 0Tθt2𝑑t>0\int_{0}^{T}\theta_{t}^{2}\,dt>0~\mathbb{P}-a.s since ff is continuous and not identically zero. Thus, all the conditions of Lemma 1 are satisfied.

The proofs for the additional two examples follow the proof of the previous example, and are thus omitted.

Example 5

(Correlated Scott model): The Scott (or exponential Ornstein-Uhlenbeck) model is given (besides the bond with constant interest rate rr) by the stock price dynamics

dSt\displaystyle dS_{t} =(r+f(Wt1)eYt)Stdt+eYtStdWt1,S0=s,\displaystyle=\Bigl{(}r+f(W_{t}^{1})e^{Y_{t}}\Bigr{)}S_{t}\,dt+e^{Y_{t}}S_{t}\,dW_{t}^{1},\quad S_{0}=s,
dYt\displaystyle dY_{t} =κ(θYt)dt+ϱξdWt1+1ϱ2ξdWt2,Y0=y,\displaystyle=\kappa(\theta-Y_{t})\,dt+\varrho\xi\,dW_{t}^{1}+\sqrt{1-\varrho^{2}}\xi\,dW_{t}^{2},\quad Y_{0}=y,

for constants, κ\kappa, θ\theta, ξ\xi, ss, y>0y>0 and ϱ(1,1)\varrho\in(-1,1) and independent Brownian motions W1W^{1}, W2W^{2}. Again we assume that the excess appreciation rate f(Wt1)f(W_{t}^{1}) is given via a bounded C1()C^{1}(\mathbb{R})-function ff with bounded derivative, and that it is not identically zero. This guarantees that the market price of risk θt=f(Wt1)\theta_{t}=f(W_{t}^{1}) remains bounded and ensures the integrability condition of Assumption 4.

Example 6

(Correlated Heston model under Feller condition): We consider a bond with constant interest rate rr and the stock price given by

dSt\displaystyle dS_{t} =(r+f(Wt1)Yt)Stdt+YtStdWt1,S0=s,\displaystyle=\bigl{(}r+f(W_{t}^{1})\sqrt{Y_{t}}\bigr{)}S_{t}\,dt+\sqrt{Y_{t}}S_{t}\,dW_{t}^{1},\quad S_{0}=s,
dYt\displaystyle dY_{t} =κ(θYt)dt+ϱξYtdWt1+1ϱ2ξYtdWt2,Y0=y,\displaystyle=\kappa\bigl{(}\theta-Y_{t}\bigr{)}\,dt+\varrho\xi\sqrt{Y_{t}}\,dW_{t}^{1}+\sqrt{1-\varrho^{2}}\xi\sqrt{Y_{t}}\,dW_{t}^{2},\quad Y_{0}=y,

for constants κ\kappa, θ\theta, ξ\xi, ss, y>0y>0 and ϱ(1,1)\varrho\in(-1,1) and independent Brownian motions W1W^{1}, W2W^{2}. Moreover, we impose the Feller condition 2κθ>ξ22\kappa\theta>\xi^{2}. Again, we assume that the excess appreciation rate f(Wt1)f(W_{t}^{1}) is given via a bounded C1()C^{1}(\mathbb{R})-function ff with bounded derivative, and that it is not identically zero. This guarantees that the market price of risk θt=f(Wt1)\theta_{t}=f(W_{t}^{1}) remains bounded and ensures the that integrability condition of Assumption 4 is satisfied.

Finally, we would like to remark that the same reasoning applied above to stochastic volatility models also holds true for Markovian regime switching models, as they have the same kind of representation of equivalent local martingale measures (c.f. Siu ). As long as the market price of risk is positive, sufficiently (Skorohod-) integrable and depends only on the stock-driving Brownian motion, the infimum is attained independently of the volatility risk. Therefore, we conclude that the optimizer has a density.

5 Preliminaries - The Classical Utility Optimization Problem

We will now briefly review the classical results of utility optimization. We adapt statements of BTZ and WZ (mainly Theorem 3.2 of BTZ ) on non-smooth utility maximization for the use in our setting. To keep notation concise and well-integrated with the rest of the paper, we hide our incentive scheme by setting g(x)=xg(x)=x throughout this section. To emphasize this we will talk about the classical utility optimization problems

u(x):=supX𝒳(x)𝔼[U(XT)],u(x):=\sup_{X\in\mathcal{X}(x)}\mathbb{E}\bigl{[}U\bigl{(}X_{T}\bigr{)}\bigr{]}, (38)

and its dual

v(y):=infY𝒴(y)𝔼[U(YT)].v(y):=\inf_{Y\in\mathcal{Y}(y)}\mathbb{E}\Bigl{[}U^{*}\bigl{(}Y_{T}\bigr{)}\Bigr{]}. (39)

We will continue with our standing Assumptions 1 and 2 throughout this section. However, we will later relax Assumption 3. The central result of Kramkov and Schachermayer (KramSchach1, , Theorem 2.2.) is the following:

Theorem 5.1 (Kramkov-Schachermayer)

Under the Assumptions 1, 2, and 3, for the utility maximization problem (38), it holds that

  • a)

    The functions uu and vv are finite on >0\mathbb{R}_{>0} and conjugate, i.e., v=uv=u^{*}. Moreover uu and v-v are strictly concave, strictly increasing, continuously differentiable on >0{\mathbb{R}_{>0}}, satisfy the Inada conditions (4) and uu satisfies the asymptotic elasticity condition (5).

  • b)

    The optimal solutions X^(x)𝒳(x)\hat{X}(x)\in\mathcal{X}(x) for (38) and Y^(y)𝒴(y)\hat{Y}(y)\in\mathcal{Y}(y) for (39) exist, are unique and are for y=u(x)y=u^{\prime}(x) related through

    X^T(x)=(U)(Y^T(y)),Y^T(y)=U(X^T(x)).\hat{X}_{T}(x)=-(U^{*})^{\prime}\bigl{(}\hat{Y}_{T}(y)\bigr{)},\qquad\qquad\hat{Y}_{T}(y)=U^{\prime}\bigl{(}\hat{X}_{T}(x)\bigr{)}.

    Moreover, X^(x)Y^(y)\hat{X}(x)\hat{Y}(y) is a uniformly integrable martingale.

  • c)

    Additionally we have

    v(y)=infe𝔼[U(ydd)],v(y)=\inf_{\mathbb{Q}\in\mathcal{M}^{e}}\mathbb{E}\biggl{[}U^{*}\biggl{(}y\frac{d\mathbb{Q}}{d\mathbb{P}}\biggr{)}\biggr{]},

    however the infimum is in general not attained in e\mathcal{M}^{e}.

Asymptotic elasticity is the minimal condition to assure the duality result in general semimartingale models for smooth utility functions (cf. KramSchach1 ). (If one poses a joint condition on model and utility function, then the minimal condition is the finiteness of the dual value function, cf. KramSchach2 .) However, as previously mentioned, the concavified utility function U¯\bar{U}^{**} will be, in general, neither strictly concave nor satisfy the Inada condition at 0. Thus, we will have to rely on results for nonsmooth utility maximization. While we will still impose Assumptions 1 and 2, we will have to relax Assumption 3. In the nonsmooth case, it turns out that the asymptotic elasticity – following Deelstra, Pham, and Touzi DPhT – has to be written on the convex conjugate of the utility function. The following general result is due to Bouchard, Touzi and Zeghal (BTZ, , Theorem 3.2.). A simplification of the proof can be found in Westray and Zheng (WZ, , Theorem 5.1.).

We relax the conditions on the utility function UU, assuming only that U:(α,)U:(\alpha,\infty)\rightarrow\mathbb{R}, α\alpha\in\mathbb{R}, is nonconstant, nondecreasing and concave (we extend UU again continuously to [α,)[\alpha,\infty), allowing the value -\infty at α\alpha while still assuming that U()>0U(\infty)>0). In particular, we no longer assume that UU is continuously differentiable on (α,)(\alpha,\infty) nor do we require that UU to be strictly increasing or strictly concave. Finally, we no longer impose Inada conditions, but merely that the closure of the domain of the dual function is 0\mathbb{R}_{\geq 0}. As mentioned above, the asymptotic elasticity condition will be written on the dual function. Hence, we substitute the following assumption for Assumption 3

Assumption 5

The investor’s preferences are represented by a utility function U:(α,)U:(\alpha,\infty)\rightarrow\mathbb{R}.

  • a)

    We assume that UU is nonconstant, nondecreasing and concave;

  • b)

    The dual function satisfies domU¯=0\overline{\operatorname{dom\,}{U^{*}}}={\mathbb{R}_{\geq 0}};

  • c)

    Moreover, it satisfies the dual asymptotic elasticity condition

    AE(U):=lim supy0supxU(y)yxU(y)<.AE^{*}(U):=\limsup_{y\to 0}\sup_{x\in-\partial U^{*}(y)}\frac{yx}{U^{*}(y)}<\infty. (40)
  • d)

    There exists y>0y>0 such that v(y)v(y) defined by (39) is finite.

Remark 1

We note that for smooth UU the classical and dual asymptotic elasticity condition are equivalent under Inada-type conditions (cf. (DPhT, , Proposition 4.1.) for a precise statement).

Theorem 5.2 (Bouchard-Touzi-Zeghal)

Assume that Assumptions 1, 2, and 5 are satisfied, then for the optimization problems (38) and (39) it holds that

  • a)

    The functions uu and vv are finite on (α,)(\alpha,\infty) and >0\mathbb{R}_{>0} respectively, and conjugate, i.e., v=uv=u^{*}.

  • b)

    Optimal solutions X^(x)𝒳(x)\hat{X}(x)\in\mathcal{X}(x) for (38) and Y^(y)𝒴(y)\hat{Y}(y)\in\mathcal{Y}(y) for (39) exist such that for some yu(x)y\in\partial u(x) we have that X^(x)Y^(y)\hat{X}(x)\hat{Y}(y) is a uniformly integrable martingale and

    X^T(x)U(Y^T(y)).\displaystyle\hat{X}_{T}(x)\in-\partial U^{*}\bigl{(}\hat{Y}_{T}(y)\bigr{)}. (41)
  • c)

    Additionally we have

    v(y)=infe𝔼[U(ydd)],\displaystyle v(y)=\inf_{\mathbb{Q}\in\mathcal{M}^{e}}\mathbb{E}\biggl{[}U^{*}\biggl{(}y\frac{d\mathbb{Q}}{d\mathbb{P}}\biggr{)}\biggr{]}, (42)

    however the infimum is in general not attained in e\mathcal{M}^{e}.

Note that the subdifferential-valued random variables in part b) should be understood as random variables whose range is a subset of the image of a random variable under a set-valued function. This is a much larger set then just the collection of random variables one would obtain by picking only fixed elements in the subdifferential and looking on images under these mappings. In the first case we can have a different mapping for every ωΩ\omega\in\Omega, whereas in the second case one fixes a single function for all ω\omega.

Proof

We adapted here the statement of BTZ and WZ to better fit the framework with KramSchach1 .The fact that the formulations in [3] and [29] differ from the formulations in [16] stems from the goal of the authors of [3] and [29] to accommodate a discussion of portfolio optimization on the whole real line with random initial endowment (as opposed to [16] who consider a simple portfolio optimization problem on the positive half of real line). However, their formulations (in terms of processes or terminal random variables) are equivalent for our case (without random endowment); they are (using the terminology of Kramkov/Schachermayer) the concrete and the abstract side of the same problem. The ultimate reason for the equality of both formulations is that the set of nonnegative T\mathcal{F}_{T}-measurable random variables dominated by some YTY_{T}, Y𝒴(y)Y\in\mathcal{Y}(y), is the bipolar of the set {ydd:e}\bigl{\{}y\frac{d\mathbb{Q}}{d\mathbb{P}}\,:\,\mathbb{Q}\in\mathcal{M}^{e}\bigr{\}}. This is due to the bipolar theorem on the cone of nonnegative random variables proved by Brannath and Schachermayer BS . For details, see (KramSchach1, , Proposition 3.1 and Section 4).

First, without loss of generality, we may assume that α=0\alpha=0, otherwise, shift everything by α\alpha. Next, we can apply Theorem 3.2 of BTZ with B=0,β=0B=0,\beta=0. It is not hard to see that, by Assumption 1 and by the concavity of UU, the function uu is finite on 0{\mathbb{R}_{\geq 0}}. The fact that vv is finite follows directly from Lemma 5.4 of WZ and from Assumption 5. One concludes that v=uv=u^{*} from Theorem 3.2 part (iii) of BTZ .

The existence of optimal solutions X^(x)\hat{X}(x) and Y^(y)\hat{Y}(y) follows from parts (i) and (ii) of Theorem 3.2 of BTZ , respectively. The fact that yu(x)y\in\partial u(x) is a consequence of part (i). Additionally, (41) and the fact that X^(x)Y^(y)\hat{X}(x)\hat{Y}(y) is a (uniformly integrable) martingale follow from part (iii). Finally, (42) is obtained from Remark 3.9 part 1. of BTZ . ∎

Finally, we note that the solutions are, in general, not unique and that the value function may not be smooth. Moreover, there may well exist a random variable ZU(Y^T(y))Z\in-\partial U^{*}\bigl{(}\hat{Y}_{T}(y)\bigr{)} satisfying 𝔼[ZY^T(y)]=xy\mathbb{E}[Z\hat{Y}_{T}(y)]=xy, which is not dominated by the terminal value of any X(x)𝒳(x)X(x)\in\mathcal{X}(x), as shown by Westray and Zheng in WZcounter .

6 The Dual and the Concavified Problem

Keeping in mind our standing Assumptions 1-3, we resume our discussion about the portfolio manager’s maximization problem 3, the concavified problem 6, and their common dual problem 7. Our plan now is to apply Theorem 5.2. Therefore, must first ensure that U¯\bar{U}^{**} satisfies all the conditions of Theorem 5.2. We also collect some properties of this function:

Proposition 2

For the concavified utility function U¯\bar{U}^{**} we have

domU¯¯=[β,),β:=inf{x>0:U¯(x)>}[0,).\overline{\operatorname{dom\,}{\bar{U}^{**}}}=[\beta,\infty),\qquad\beta:=\inf\{x>0:\bar{U}(x)>-\infty\}\in[0,\infty).

Furthermore, U¯\bar{U}^{**}, together with its conjugate U¯\bar{U}^{*}, enjoys the following regularity properties. U¯\bar{U}^{**} is continuously differentiable on (β,)(\beta,\infty); U¯\bar{U}^{*} is strictly convex on the whole domain if U(0)=U(0)=-\infty, otherwise it is strictly convex on (0,(U¯)(0))\bigl{(}0,\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(0)\bigr{)} and constant U¯(0)=U¯(0)\bar{U}(0)=\bar{U}^{**}(0) on [(U¯)(0),)\bigl{[}\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(0),\infty\bigr{)}. Finally, U¯\bar{U}^{**} satisfies Assumption 1 and Assumption 5.

We divide the proof into three lemmas. The proof is elementary, but rather technical, so it can be safely skipped on the first reading.

Lemma 5

For the concavified utility function U¯\bar{U}^{**} it holds that domU¯¯=[β,)\overline{\operatorname{dom\,}{\bar{U}^{**}}}=[\beta,\infty), and it satisfies Assumption 1 and Assumption 5, a), b) and d).

Proof

Consider first the case U(0)>U(0)>-\infty. Note that since U¯\bar{U} is continuous, its epigraph is closed and thus U¯\bar{U}^{**} is its concave hull. Thus, by Caratheodory’s theorem (cf. (HUL, , Theorem A.1.3.6.)), we know that

(0,U¯(0)){i=13λizi:λi0,i=13λi=1,zihypoU¯}.\big{(}0,\bar{U}^{**}(0)\bigr{)}\in\Biggl{\{}\sum_{i=1}^{3}\lambda_{i}z_{i}\,:\,\lambda_{i}\geq 0,\,\sum_{i=1}^{3}\lambda_{i}=1,\,z_{i}\in\operatorname{hypo\,}\bar{U}\Biggr{\}}.

Since hypoU¯0×\operatorname{hypo\,}\bar{U}\subseteq{\mathbb{R}_{\geq 0}}\times\mathbb{R}, it follows that the linear combination has to be the trivial, i.e, (0,U¯(0))hypoU¯\big{(}0,\bar{U}^{**}(0)\bigr{)}\in\operatorname{hypo\,}\bar{U} and U¯(0)=U¯(0)>\bar{U}(0)=\bar{U}^{**}(0)>-\infty. Thus, it follows that domU¯=[0,)=[β,)\operatorname{dom\,}\bar{U}^{**}=[0,\infty)=[\beta,\infty).

Similarly, if U(0)=U(0)=-\infty, we note first that if g(0)>0g(0)>0, we have U¯(0)>\bar{U}(0)>-\infty and β=0\beta=0. Thus, we can conclude, exactly as in the previous case, that domU¯=[β,)\operatorname{dom\,}\bar{U}^{**}=[\beta,\infty). However, if g(0)=0g(0)=0, we know by the definition of β\beta that U¯(x0)\bar{U}(x_{0}) is real valued if and only if x0(β,)x_{0}\in(\beta,\infty). In this case, the assumption that U¯(β)>\bar{U}^{**}(\beta)>-\infty leads to a contradiction by Caratheodory’s theorem. It follows that U¯(β)=U¯(β)\bar{U}^{**}(\beta)=\bar{U}(\beta) and hence domU¯=(β,)\operatorname{dom\,}\bar{U}^{**}=(\beta,\infty). Putting the information from all three cases together we recover the statement domU¯¯=[β,)\overline{\operatorname{dom\,}\bar{U}^{**}}=[\beta,\infty).

Now, set b:=g(β)b:=g(\beta) and note that U¯(x)U(x+b)\bar{U}(x)\leq U(x+b). We have, for y>0y>0

U¯(y)=supx>β(U¯(x)xy)supx>β(U(x+b)(x+b)y)+byU(y)+by<.\bar{U}^{*}(y)=\sup_{x>\beta}\Bigl{(}\bar{U}(x)-xy\Bigr{)}\leq\sup_{x>\beta}\Bigl{(}U(x+b)-(x+b)y\Bigr{)}+by\leq U^{*}(y)+by<\infty. (43)

Hence, we have  dom(U¯)¯=0\overline{\mbox{ dom}(\bar{U}^{*})}={\mathbb{R}_{\geq 0}}, i.e., part b) of Assumption 5 is satisfied. It is straight forward to see that part a) holds for the concavification of a nondecreasing, nonconstant function. Finally, using the above, it follows also for x>βx>\beta

U¯(x)=supy>0(U¯(y)+xy)supy>0(U(y)+by+xy)=U(x+b)=U(x+b).\bar{U}^{**}(x)=\sup_{y>0}\Bigl{(}\bar{U}^{*}(y)+xy\Bigr{)}\leq\sup_{y>0}\Bigl{(}U^{*}(y)+by+xy\Bigr{)}=U^{**}(x+b)=U(x+b).

We conclude by Theorem 5.1 that

w(x)supX𝒳(x)𝔼[U(XT+b)]supX𝒳(x+b)𝔼[U(XT)]w(x)\leq\sup_{X\in\mathcal{X}(x)}\mathbb{E}\bigl{[}{U}\bigl{(}X_{T}+b\bigr{)}\bigr{]}\leq\sup_{X\in\mathcal{X}(x+b)}\mathbb{E}\bigl{[}U\bigl{(}X_{T}\bigr{)}\bigr{]}

is finite on (β,)(\beta,\infty). This proves Assumption 1. Moreover, from (43), we see that

v(y)infY𝒴(y)𝔼[U(YT)]+by.v(y)\leq\inf_{Y\in\mathcal{Y}(y)}\mathbb{E}\bigl{[}{U}^{*}\bigl{(}Y_{T}\bigr{)}\bigr{]}+by. (44)

The right hand side of (44) is finite by Theorem 5.1. Thus, vv is finite on >0{\mathbb{R}_{>0}}, and Assumption 5 part d) is satisfied. Hence, all the requirements of Theorem 5.2 are satisfied except c) of Assumption 5, whose proof we postpone to Lemma 7. ∎

Lemma 6

The concavified utility function U¯\bar{U}^{**} and its conjugate U¯\bar{U}^{*} enjoy the following regularity properties:

  • a)

    The concavified utility function U¯\bar{U}^{**} is continuously differentiable on (β,)(\beta,\infty).

  • b)

    The dual utility function U¯\bar{U}^{*} is strictly convex on the whole domain if U(0)=U(0)=-\infty, otherwise it is strictly convex on (0,(U¯)(0))\bigl{(}0,\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(0)\bigr{)} and constant U¯(0)=U¯(0)\bar{U}(0)=\bar{U}^{**}(0) on [(U¯)(0),)\bigl{[}\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(0),\infty\bigr{)}.

Proof

To prove a), we note first that the set A:={x>β:U¯(x)U¯(x)}A:=\{x>\beta\,:\,\bar{U}(x)\neq\bar{U}^{**}(x)\} where U¯\bar{U} and U¯\bar{U}^{**} do not agree, is a countable union of pairwise disjoint open intervals. We also note that the function U¯\bar{U} is continuous on (β,)(\beta,\infty) since it is the composition of continuous functions (gg is as convex, nondecreasing function, thus continuous). The same is true for U¯\bar{U}^{**}, which is a concave function by definition. Hence, AA is the 0-sublevel set of the continuous function U¯U¯\bar{U}^{**}-\bar{U} and is thus open. But every open set in \mathbb{R} can be written as countable union of pairwise disjoint open intervals, say A=n=1(an,an+)A=\bigcup_{n=1}^{\infty}(a^{-}_{n},a^{+}_{n}), an<an+a_{n}^{-}<a_{n}^{+}. We note explicitly that a1=βa_{1}^{-}=\beta and an+=a_{n}^{+}=\infty for some nn are allowed. On every interval in AA the function U¯\bar{U}^{**} is affine (the straight linear interpolation between U¯(an)\bar{U}(a_{n}^{-}) and U¯(an+)\bar{U}(a_{n}^{+})) and hence we can write it as U¯(x)=gammanx+αn\bar{U}^{**}(x)=\ gamma_{n}x+\alpha_{n} for some γn>0\gamma_{n}\in\mathbb{R}_{>0}, αn\alpha_{n}\in\mathbb{R}, with {γn}\{\gamma_{n}\} a sequence satisfying that if indices nn and mm are such that an+ama_{n}^{+}\leq a_{m}^{-} then γnγm\gamma_{n}\geq\gamma_{m}. Thus, clearly U¯\bar{U}^{**} is differentiable in AA.

Now, denote by BB the open interior of the set where U¯\bar{U} and U¯\bar{U}^{**} agree, i.e., B:={x>β:U¯(x)=U¯(x)}B:=\{x>\beta\,:\,\bar{U}(x)=\bar{U}^{**}(x)\}^{\circ}. We will prove that, on the set BB, the function U¯\bar{U} is continuously differentiable. Pick some point xBx\in B. Since gg is convex, it holds that gr(x)gl(x)g^{\prime}_{r}(x)\geq g^{\prime}_{l}(x), where gr,glg^{\prime}_{r},g^{\prime}_{l} are the left- and right-hand derivatives, respectively. Thus, it follows by the differentiability of UU that U¯r(x)=U(g(x)))gr(x)U(g(x)))gl(x)=U¯l(x)\bar{U}^{\prime}_{r}(x)=U^{\prime}\bigl{(}g(x)\bigr{)})g^{\prime}_{r}(x)\geq U^{\prime}\bigl{(}g(x)\bigr{)})g^{\prime}_{l}(x)=\bar{U}^{\prime}_{l}(x). But, on the other hand, the concavity of U¯\bar{U}^{**} implies U¯r(x)=(U¯)r(x)(U¯)l(x)=U¯l(x)\bar{U}^{\prime}_{r}(x)=\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}_{r}(x)\leq\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}_{l}(x)=\bar{U}^{\prime}_{l}(x). Thus, the left- and right-derivatives have to agree for every xBx\in B. We conclude, then, that the function is continuously differentiable there.

We then note that (A¯B)\β=(β,)\left(\overline{A}\cup B\right)\backslash\beta=(\beta,\infty). Thus to complete our argument, it remains only to prove continuous differentiability on one of the points aA¯\(Aβ)a\in\overline{A}\backslash\left(A\cup\beta\right). Note that for such aa, we can find a sequence ank±a_{n_{k}}^{\pm} (assume without loss of generality it is anka_{n_{k}}^{-}, as we can handle ank+a_{n_{k}}^{+} the same way), such that limkank=a\lim\limits_{k\to\infty}a_{n_{k}}^{-}=a. Additionally, note that by continuity of U¯,U¯\bar{U}^{**},\bar{U}, and the fact that aAa\not\in A it follows that U¯(a)=U¯(a)\bar{U}^{**}(a)=\bar{U}(a). Assume by contradiction that U¯\bar{U}^{**} is not continuously differentiable at aa. It follows that

U¯r(a)U¯l(a)(U¯)l(a)>(U¯)r(a).\bar{U}^{\prime}_{r}(a)\geq\bar{U}^{\prime}_{l}(a)\geq\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}_{l}(a)>\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}_{r}(a). (45)

The first inequality stems from the fact that (since UU is continuously differentiable) every point of non-differentiability of U¯\bar{U} is due to not having an interior derivative of UgU\circ g. However, for the convex function gg we have glgrg_{l}^{\prime}\leq g_{r}^{\prime} (and U0U^{\prime}\geq 0). The strict inequality is the consequence of our assumption that U¯\bar{U}^{**} is not differentiable at aa and that it is a concave function. The second inequality follows from the fact that U¯\bar{U}^{**} is the concave hull of U¯\bar{U} (and both functions agree on aa). Indeed, using the concavity of U¯\bar{U}^{**} and the fact that U¯(a)=U¯(a)\bar{U}(a)=\bar{U}^{**}(a) we write

U¯l(a)limh0+U¯(a)U¯(ah)hlimh0+U¯(a)U¯(ah)h=(U¯)l(a).\bar{U}^{\prime}_{l}(a)\geq\lim\limits_{h\to 0^{+}}\frac{\bar{U}(a)-\bar{U}(a-h)}{h}\geq\lim\limits_{h\to 0^{+}}\frac{\bar{U}^{**}(a)-\bar{U}^{**}(a-h)}{h}=\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}_{l}(a).

However, (45) leads to a contradiction, since, by a similar argument

(U¯)r(a)limh0+U¯(a+h)U¯(a)hlimh0+U¯(a+h)U¯(a)hU¯r(a).\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}_{r}(a)\geq\lim\limits_{h\to 0^{+}}\frac{\bar{U}^{**}(a+h)-\bar{U}^{**}(a)}{h}\geq\lim\limits_{h\to 0^{+}}\frac{\bar{U}(a+h)-\bar{U}(a)}{h}\geq\bar{U}^{\prime}_{r}(a).

Thus, U¯\bar{U}^{**} has to be continuously differentiable in aa and hence, on the whole interval (β,)(\beta,\infty).

In passing we note that the differentiability of U¯\bar{U}^{**} implies that U¯\bar{U}^{*} cannot be differentiable at any γn\gamma_{n}. Assume indirectly that it would be differentiable. Then, there exists some a~\tilde{a}\in\mathbb{R} such that (U¯)(γn)=a~-\bigl{(}\bar{U}^{*}\bigr{)}^{\prime}(\gamma_{n})=\tilde{a}. Furthermore, convex duality implies γnU¯(a~)\gamma_{n}\in\partial\bar{U}^{**}(\tilde{a}). However, the differentiability of U¯\bar{U}^{**} reduces the subdifferential to a singleton. This means that γn\gamma_{n} can only be the slope of U¯\bar{U}^{**} at the single point a~\tilde{a}, which is in contradiction to the fact that it is the slope on the whole interval (an,an+)(a_{n}^{-},a_{n}^{+}).

Finally to show b), we note that the strict convexity in the range of the gradient mapping is a classical consequence in convex Analysis, see e.g., (HUL, , Theorem E.4.1.2.), i.e., U¯\bar{U}^{*} is strictly convex on {(U¯)(x):x(β,)}\bigl{\{}\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(x)\,:\,x\in(\beta,\infty)\bigr{\}}. We claim that {(U¯)(x):x(β,)}=(0,(U¯)(β))\bigl{\{}\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(x)\,:\,x\in(\beta,\infty)\bigr{\}}=\bigl{(}0,(\bar{U}^{**})^{\prime}(\beta)\bigr{)}. Indeed, (U¯)\bigl{(}\bar{U}^{**}\bigr{)}^{\prime} is nonincreasing, and for x>max{a1+,β}x>\max\{a_{1}^{+},\beta\}

(U¯)(x)\displaystyle\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(x) ={U(g(x))g(x)xA¯γnx[an,an+]}\displaystyle=\left\{\begin{array}[]{ll}U^{\prime}\bigl{(}g(x)\bigr{)}g^{\prime}(x)&x\notin\overline{A}\\ \gamma_{n}&x\in[a_{n}^{-},a_{n}^{+}]\end{array}\right\}
={U(g(x))g(x)xA¯U(g(an))g(an)x[an,an+]}U(g(x)),\displaystyle=\left\{\begin{array}[]{ll}U^{\prime}\bigl{(}g(x)\bigr{)}g^{\prime}(x)&x\notin\overline{A}\\ U^{\prime}\bigl{(}g(a_{n}^{-})\bigr{)}g^{\prime}(a_{n}^{-})&x\in[a_{n}^{-},a_{n}^{+}]\end{array}\right\}\leq U^{\prime}\bigl{(}g(x)\bigr{)},

with an±a_{n}^{\pm} the boundary points of intervals in AA as above. Thus, since gg is convex, nonconstant and nondecreasing function, it must satisfy limxg(x)=\lim\limits_{x\to\infty}g(x)=\infty. It follows by the Inada condition at \infty that 0(U¯)()U()=00\leq\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(\infty)\leq U^{\prime}(\infty)=0. For the right hand of the domain of strict convexity of U¯\bar{U}^{*} we have to consider three cases. First, if U(0)=U(0)=-\infty, then we have (U¯)(β)=\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(\beta)=\infty, since U¯(β)=U¯(β)=\bar{U}(\beta)=\bar{U}^{**}(\beta)=-\infty. We therefore obtain {(U¯)(x):x(β,)}=(0,)\bigl{\{}(\bar{U}^{**})^{\prime}(x)\,:\,x\in(\beta,\infty)\bigr{\}}=(0,\infty). Second, if U(0)U(0) is real and (U¯)(β)=\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(\beta)=\infty, then we can conclude in similar manner that {(U¯)(x):x(β,)}=(0,)\bigl{\{}(\bar{U}^{**})^{\prime}(x)\,:\,x\in(\beta,\infty)\bigr{\}}=(0,\infty). Finally, if U(0)U(0) is real and (U¯)(β)<\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(\beta)<\infty, then U¯\bar{U}^{*} is strictly convex on (0,(U¯)(0))\bigl{(}0,\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(0)\bigr{)}. However, for y(U¯)(0)=maxx>β(U¯)(x)y\geq\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(0)=\max_{x>\beta}\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(x), we can conclude that U¯(0)supx>β(U¯(x)xy)supx>β(U¯(x)xy)=U¯(0)\bar{U}(0)\leq\sup_{x>\beta}\bigl{(}\bar{U}(x)-xy\bigr{)}\leq\sup_{x>\beta}\bigl{(}\bar{U}^{**}(x)-xy\bigr{)}=\bar{U}^{**}(0). Since U¯(0)=U¯(0)\bar{U}(0)=\bar{U}^{**}(0), it follows that v(y)U¯(0)v(y)\equiv\bar{U}^{**}(0) on [(U¯)(β),)\bigl{[}(\bar{U}^{**})^{\prime}(\beta),\infty\bigr{)}. ∎

Finally, we have to prove the dual asymptotic ellipticity of U¯\bar{U}^{**}. The following result builds on and generalizes (in the one-dimensional case) the equivalence result of dual and classical asymptotic elasticity given by Deelstra, Pham and Touzi (DPhT, , Proposition 4.1.) (their result can be seen as the linear case g(x)=xg(x)=x).

Lemma 7

The concavified function U¯\bar{U}^{**} satisfies the dual asymptotic elasticity condition (40), i.e.,

AE(U¯)=lim supy0supxU¯(y)yxU¯(y)<.AE^{*}\bigl{(}\bar{U}^{**}\bigr{)}=\limsup_{y\to 0}\sup_{x\in-\partial\bar{U}^{*}(y)}\frac{yx}{\bar{U}^{*}(y)}<\infty.
Proof

First, we note that, by the slope bound and the non-constancy of gg

c:=supx0g(x)c:=\sup\bigcup_{x\geq 0}\partial g(x)

is finite and strictly positive. Thus, we obtain on one hand that there exists for every ε>0\varepsilon>0 some x0x_{0} (which we will assume to be bigger then β\beta) such that for all x>x0x>x_{0}

g(x0)+(cε)(xx0)g(x)g(0)+cx,g(x_{0})+(c-\varepsilon)(x-x_{0})\leq g(x)\leq g(0)+cx, (46)

and

(cε)inf[x0,)gsup[x0,)gc.(c-\varepsilon)\leq\inf_{[x_{0},\infty)}\partial g\leq\sup_{[x_{0},\infty)}\partial g\leq c.

Moreover, we note that, in the case of affine g~(x)=ax+b\tilde{g}(x)=ax+b with a>0a\in\mathbb{R}_{>0} and bb\in\mathbb{R}, we have that

supxdomUg~(U(g~(x))xy)=supx>ba(U(ax+b)xy)=supz>0(U(z)zya)+bya=U(ya)+bya.\sup_{x\in\operatorname{dom\,}U\circ\tilde{g}}\Bigl{(}U\bigl{(}\tilde{g}(x)\bigr{)}-xy\Bigr{)}=\sup_{x>-\frac{b}{a}}\Bigl{(}U(ax+b)-xy\Bigr{)}=\sup_{z>0}\biggl{(}U(z)-z\frac{y}{a}\biggr{)}+\frac{by}{a}=U^{*}\Bigl{(}\frac{y}{a}\Bigr{)}+\frac{by}{a}.

Setting a:=cεa:=c-\varepsilon and b:=g(x0)(cε)x0b:=g(x_{0})-(c-\varepsilon)x_{0}, we note that

supxdomUg~(U(g~(x))xy)=supx>x0(U(g~(x))xy)\sup_{x\in\operatorname{dom\,}U\circ\tilde{g}}\Bigl{(}U\bigl{(}\tilde{g}(x)\bigr{)}-xy\Bigr{)}=\sup_{x>x_{0}}\Bigl{(}U\bigl{(}\tilde{g}(x)\bigr{)}-xy\Bigr{)}

as long as y<U(g(x0))(cε)=:y0y<U^{\prime}\bigl{(}g(x_{0})\bigr{)}(c-\varepsilon)=:y_{0}. Thus, we can conclude by (46) that for y(0,y0)y\in(0,y_{0}) it holds that

U¯(y)\displaystyle\bar{U}^{*}(y) =supxdomU¯(U¯(x)xy)=supx>β(U¯(x)xy)supx>x0((Ug)(x)xy)\displaystyle=\sup_{x\in\operatorname{dom\,}\bar{U}}\Bigl{(}\bar{U}(x)-xy\Bigr{)}=\sup_{x>\beta}\Bigl{(}\bar{U}(x)-xy\Bigr{)}\geq\sup_{x>x_{0}}\Bigl{(}\bigl{(}U\circ g\bigr{)}(x)-xy\Bigr{)}
supx>x0((Ug~)(x)xy)=supxdomUg~((Ug~)(x)xy)\displaystyle\geq\sup_{x>x_{0}}\Bigl{(}\bigl{(}U\circ\tilde{g}\bigr{)}(x)-xy\Bigr{)}=\sup_{x\in\operatorname{dom\,}U\circ\tilde{g}}\Bigl{(}\bigl{(}U\circ\tilde{g}\bigr{)}(x)-xy\Bigr{)}
=U(ycε)+g(x0)(cε)x0cεy.\displaystyle=U^{*}\Bigl{(}\frac{y}{c-\varepsilon}\Bigr{)}+\frac{g(x_{0})-(c-\varepsilon)x_{0}}{c-\varepsilon}y.

We note that from Lemma 6, it follows that for x>a1+x>a_{1}^{+} (and all x>βx>\beta in the case of concave U¯\bar{U})

(U¯)(x)\displaystyle\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(x) ={U(g(x))g(x)xA¯γnx[an,an+]}\displaystyle=\left\{\begin{array}[]{ll}U^{\prime}\bigl{(}g(x)\bigr{)}g^{\prime}(x)&\quad x\notin\overline{A}\\ \gamma_{n}&\quad x\in[a_{n}^{-},a_{n}^{+}]\end{array}\right\} (49)
={U(g(x))g(x)xA¯U(g(an))g(an)x[an,an+]}U(g(x))c.\displaystyle=\left\{\begin{array}[]{ll}U^{\prime}\bigl{(}g(x)\bigr{)}g^{\prime}(x)&\quad x\notin\overline{A}\\ U^{\prime}\bigl{(}g(a_{n}^{-})\bigr{)}g^{\prime}(a_{n}^{-})&\quad x\in[a_{n}^{-},a_{n}^{+}]\end{array}\right\}\leq U^{\prime}\bigl{(}g(x)\bigr{)}c. (52)

By convex conjugacy we have

xU¯(y)y=(U¯)(x).x\in-\partial\bar{U}^{*}(y)\qquad\Longleftrightarrow\qquad y=\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(x).

Hence, since by concavity the gradient is nonincreasing, we see that

xinfU¯(y)y(U¯)(x).x\leq-\inf\partial\bar{U}^{*}(y)\qquad\Longleftrightarrow\qquad y\geq(\partial\bar{U}^{**})^{\prime}(x).

Thus, we can conclude that (U)1=(U)(U^{\prime})^{-1}=-(U^{*})^{\prime}. Let f1(y):=inf{z:f(z)>y}f^{-1}(y):=\inf\{z\,:\,f(z)>y\} be the generalized inverse of a function ff. Then

AE(U¯)\displaystyle AE^{*}\bigl{(}\bar{U}^{**}\bigr{)} =lim supy0supxU¯(y)yxU¯(y)=lim supy0sup{x:y=(U¯)(x)}yxU¯(y)\displaystyle=\limsup_{y\to 0}\sup_{x\in-\partial\bar{U}^{*}(y)}\frac{yx}{\bar{U}^{*}(y)}=\limsup_{y\to 0}\sup_{\{x\,:\,y=(\bar{U}^{*})^{\prime}(x)\}}\frac{yx}{\bar{U}^{*}(y)}
lim supy0sup{x:y(U¯)(x)}yxU¯(y)lim supy0sup{x:yU(g(x))c}yxU¯(y)\displaystyle\leq\limsup_{y\to 0}\sup_{\{x\,:\,y\leq(\bar{U}^{*})^{\prime}(x)\}}\frac{yx}{\bar{U}^{*}(y)}\leq\limsup_{y\to 0}\sup_{\{x\,:\,y\leq{U}^{\prime}(g(x))c\}}\frac{yx}{\bar{U}^{*}(y)}
=lim supy0sup{x:(U)(y/c)g(x)}yxU¯(y)lim supy0yg1((U)(yc))U¯(y).\displaystyle=\limsup_{y\to 0}\sup_{\{x\,:\,-(U^{*})^{\prime}(y/c)\geq g(x)\}}\frac{yx}{\bar{U}^{*}(y)}\leq\limsup_{y\to 0}\frac{yg^{-1}\bigl{(}-(U^{*})^{\prime}\bigl{(}\frac{y}{c}\bigr{)}\bigr{)}}{\bar{U}^{*}(y)}.

We discern now two cases. First, if (U)-(U^{*})^{\prime} is bounded, then we can directly conclude that AE(U¯)<AE^{*}\bigl{(}\bar{U}^{**}\bigr{)}<\infty, since U¯(0)=U()>0\bar{U}^{*}(0)=U(\infty)>0. Second, if (U)-(U^{*})^{\prime} is unbounded, then by the Inada condition for UU we have that

lim supy0(U(y))=lim supy0(U)1(y)=.\limsup_{y\to 0}\,\bigl{(}-U^{*}(y)\bigr{)}=\limsup_{y\to 0}\,(U^{\prime})^{-1}(y)=\infty.

From (46) we see that yg(y)g(x0)cε+x0y\leq\frac{g(y)-g(x_{0})}{c-\varepsilon}+x_{0} holds for all yx0y\geq x_{0}. Applying this to y=g1(z)y=g^{-1}(z) (note that gg is here a true inverse since x0x_{0} was assumed to be bigger then β\beta) we conclude that g1(z)zcεg(x0)cε+x0g^{-1}(z)\leq\frac{z}{c-\varepsilon}-\frac{g(x_{0})}{c-\varepsilon}+x_{0}, for all z>g(x0)z>g(x_{0}). It follows that with z=(U)(yc)z=-(U^{*})^{\prime}\bigl{(}\frac{y}{c}\bigr{)} we have

g1((U)(yc))1cε(U)(yc)(g(x0)(cε)x0)cε,g^{-1}\Bigl{(}-(U^{*})^{\prime}\bigl{(}\frac{y}{c}\bigr{)}\Bigr{)}\leq-\frac{1}{c-\varepsilon}(U^{*})^{\prime}\Bigl{(}\frac{y}{c}\Bigr{)}-\frac{(g(x_{0})-(c-\varepsilon)x_{0})}{c-\varepsilon},

for yy satisfying (U)(yc)>g(x0)-(U^{*})^{\prime}\bigl{(}\frac{y}{c}\bigr{)}>g(x_{0}). By the unboundedness of (U)-(U^{*})^{\prime} this is satisfied for all yy small enough. Since UU satisfies the dual asymptotic elasticity condition by (DPhT, , Proposition 4.1.) (cf. Remark 1) we have for some M(0,)M\in(0,\infty) that

AE(U)=lim supy0y(U)(y)U(y)<M<.AE^{*}\bigl{(}U\bigr{)}=\limsup_{y\to 0}\frac{-y(U^{*})^{\prime}(y)}{U^{*}(y)}<M<\infty.

From

U(ycε)U(yc)+εyc(cε)(U)(yc),U^{*}\Bigl{(}\frac{y}{c-\varepsilon}\Bigr{)}\geq U^{*}\Bigl{(}\frac{y}{c}\Bigr{)}+\varepsilon\frac{y}{c(c-\varepsilon)}\bigl{(}U^{*}\bigr{)}^{\prime}\Bigl{(}\frac{y}{c}\Bigr{)},

we conclude that

AE(U¯)\displaystyle AE^{*}\bigl{(}\bar{U}^{**}\bigr{)}\leq lim supy0yg1((U)(yc))U¯(y)lim supy0ycε(U)(yc)(g(x0)(cε)x0)cεyU(ycε)+g(x0)(cε)x0cεy\displaystyle\limsup_{y\to 0}\frac{yg^{-1}\bigl{(}-(U^{*})^{\prime}\bigl{(}\frac{y}{c}\bigr{)}\bigr{)}}{\bar{U}^{*}(y)}\leq\limsup_{y\to 0}\frac{-\frac{y}{c-\varepsilon}(U^{*})^{\prime}\bigl{(}\frac{y}{c}\bigr{)}-\frac{(g(x_{0})-(c-\varepsilon)x_{0})}{c-\varepsilon}y}{U^{*}\bigl{(}\frac{y}{c-\varepsilon}\bigr{)}+\frac{g(x_{0})-(c-\varepsilon)x_{0}}{c-\varepsilon}y}
\displaystyle\leq lim supy01cεy(U)(yc)U(ycε)+11cε11Mεc(cε)+1<,\displaystyle\limsup_{y\to 0}\frac{1}{c-\varepsilon}\frac{-y(U^{*})^{\prime}\bigl{(}\frac{y}{c}\bigr{)}}{U^{*}\bigl{(}\frac{y}{c-\varepsilon}\bigr{)}}+1\leq\frac{1}{c-\varepsilon}\frac{1}{\frac{1}{M}-\frac{\varepsilon}{c(c-\varepsilon)}}+1<\infty,

for ε>0\varepsilon>0 chosen small enough (note that MM only depends on the original utility function, hence it is independent of ε\varepsilon). ∎

We can now look more closely at how the concavified problem relates to the classical Kramkov/Schachermayer setting. The concavified utility function U¯\bar{U}^{**} is indeed continuously differentiable. It will follow from (49) that it satisfies also the Inada condition (U¯)()=0\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(\infty)=0. Hence, by (DPhT, , Proposition 4.1.) the primal asymptotic elasticity condition AE(U¯)<1AE\bigl{(}\bar{U}^{**}\bigr{)}<1 is also satisfied. However, it fails, in general, the Inada condition (U¯)(0)=\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(0)=\infty. Furthermore, it will not necessarily be strictly concave.

Relying heavily on Proposition 2, we can now prove Theorem 2.1, which is the result concerning existence and uniqueness of an optimal solution of the dual problem (7) as well as existence for the concavified problem (6). In the next sections we will use this central result to discuss the uniqueness of the concavified problem as well as discuss how one can use the concavified problem to solve the original problem (3).

Proof (Theorem 2.1)

It follows from Proposition 2 that the conditions of Theorem 5.2 are satisfied for the concavified utility function U¯\bar{U}^{**} with α=β\alpha=\beta. This implies the finiteness and the duality statements of a).

The existence part of b) also follows directly from Theorem 5.2. For the uniqueness part, we note Proposition 2 implies that U¯\bar{U}^{*} is strictly convex on (0,(U¯)(0))(0,(\bar{U}^{**})^{\prime}(0)). Assume, by contradiction, that Y^T1(y)\hat{Y}_{T}^{1}(y) and Y^T2(y)\hat{Y}_{T}^{2}(y) are the terminal values of two different optimizers of the dual problem such that

[{Y^T1(y)Y^T2(y)}{Y^T1(y)(0,(U¯)(0))}{Y^T2(y)(0,(U¯)(0))}]>0.\displaystyle\mathbb{P}\bigl{[}\{\hat{Y}_{T}^{1}(y)\neq\hat{Y}_{T}^{2}(y)\}\cap\{\hat{Y}_{T}^{1}(y)\in(0,(\bar{U}^{**})^{\prime}(0))\}\cap\{\hat{Y}_{T}^{2}(y)\in(0,(\bar{U}^{**})^{\prime}(0))\}\bigr{]}>0.

That is, the random variables Y^T1(y),Y^T2(y)\hat{Y}_{T}^{1}(y),\hat{Y}_{T}^{2}(y) differ on the set (0,(U¯)(0))(0,(\bar{U}^{**})^{\prime}(0)) with positive probability. It follows that for every λ(0,1)\lambda\in(0,1) and YTλ(y):=λY^T1(y)+(1λ)Y^T2(y)Y^{\lambda}_{T}(y):=\lambda\hat{Y}^{1}_{T}(y)+(1-\lambda)\hat{Y}_{T}^{2}(y) we have, by the strict convexity of U¯\bar{U}^{*}, that

𝔼[U¯(YTλ(y))]\displaystyle\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}Y_{T}^{\lambda}(y)\bigr{)}\Bigr{]} =𝔼[U¯(λY^T1(y)+(1λ)Y^T2(y))]\displaystyle=\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}\lambda\hat{Y}_{T}^{1}(y)+(1-\lambda)\hat{Y}_{T}^{2}(y)\bigr{)}\Bigr{]}
<λ𝔼[U¯(Y^T1(y))]+(1λ)𝔼[U¯(Y^T2(y))],\displaystyle<\lambda\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}\hat{Y}_{T}^{1}(y)\bigr{)}\Bigr{]}+(1-\lambda)\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}\hat{Y}_{T}^{2}(y)\bigr{)}\Bigr{]},

which contradicts the optimality of Y^T1(y)\hat{Y}_{T}^{1}(y), or Y^T2(y)\hat{Y}_{T}^{2}(y).

To prove the remaining statements of a) we note that we have for every λ(0,1)\lambda\in(0,1) and y1y_{1}, y2>0y_{2}>0

λY^(y1)+(1λ)Y^(y2)λ𝒴(y1)+(1λ)𝒴(y2)=𝒴(λy1+(1λ)y2).\lambda\hat{Y}(y_{1})+(1-\lambda)\hat{Y}(y_{2})\in\lambda\mathcal{Y}(y_{1})+(1-\lambda)\mathcal{Y}(y_{2})=\mathcal{Y}\bigl{(}\lambda y_{1}+(1-\lambda)y_{2}\bigr{)}.

Thus, we can conclude by the strict convexity of U¯\bar{U}^{*} that for λ(0,1)\lambda\in(0,1), and 0<y1<y2δ0<y_{1}<y_{2}\leq\delta with δ:=sup{y>0:supp(Y^T(y))(0,(U¯)(β)]}\delta:=\sup\{y>0\,:\,\operatorname{supp\,}{(\hat{Y}_{T}(y))}\cap(0,(\bar{U}^{**})^{\prime}(\beta)]\neq\emptyset\} we have that

v(λy1+(1λ)y2)\displaystyle v\bigl{(}\lambda y_{1}+(1-\lambda)y_{2}\bigr{)} =𝔼[U¯(Y^T(λy1+(1λ)y2))]𝔼[U¯(λY^T(y1)+(1λ)Y^T(y2))]\displaystyle=\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}\hat{Y}_{T}(\lambda y_{1}+(1-\lambda)y_{2})\bigr{)}\Bigr{]}\leq\mathbb{E}\Bigl{[}\bar{U}^{*}\bigl{(}\lambda\hat{Y}_{T}(y_{1})+(1-\lambda)\hat{Y}_{T}(y_{2})\bigr{)}\Bigr{]}
<λE[U¯(Y^T(y1))]+(1λ)E[U¯(Y^T(y2))]=λv(y1)+(1λ)v(y2).\displaystyle<\lambda E\Bigl{[}\bar{U}^{*}\bigl{(}\hat{Y}_{T}(y_{1})\bigr{)}\Bigr{]}+(1-\lambda)E\Bigl{[}\bar{U}^{*}\bigl{(}\hat{Y}_{T}(y_{2})\bigr{)}\Bigr{]}=\lambda v(y_{1})+(1-\lambda)v(y_{2}).

Hence, vv is strictly convex on (0,δ)(0,\delta) and constant U¯(0)\bar{U}^{**}(0) on [δ,)[\delta,\infty). By (HUL, , Theorem E.4.1.1.) this implies the continuous differentiability of ww.

Part c) follows directly from Theorem 5.2 and the differentiability of ww in the interior of its domain, (β,)(\beta,\infty).

Finally, d) is a direct consequence of Theorem 5.2. ∎

7 The Original Problem: Wealth-independent Solution

We are now finally ready to give the proof for Theorem 2.2 and Proposition 1. We will rely heavily on the following results from the proof of Proposition 2, specifically Lemma 6. The set AA where the two utility functions disagree is an open subset of >0{\mathbb{R}_{>0}}. As such, AA is a countable union of pairwise disjoint open intervals,

A:=n=1(an,an+)={x>0:U¯(x)U¯(x)},an<an+.A:=\bigcup_{n=1}^{\infty}(a^{-}_{n},a^{+}_{n})=\bigl{\{}x>0\,:\bar{U}(x)\neq\bar{U}^{**}(x)\bigr{\}},\qquad a_{n}^{-}<a_{n}^{+}.

On every one of these intervals the function U¯\bar{U}^{**} is affine, U¯(x)=γnx+αn\bar{U}^{**}(x)=\gamma_{n}x+\alpha_{n} for some γn>0\gamma_{n}\in\mathbb{R}_{>0}, αn\alpha_{n}\in\mathbb{R}, where {γn}\{\gamma_{n}\} is a sequence satisfying that if indices nn and mm are such that an+ama_{n}^{+}\leq a_{m}^{-} then γnγm\gamma_{n}\geq\gamma_{m}. We set

Γ:=n=1{γn},\Gamma:=\bigcup_{n=1}^{\infty}\bigl{\{}\gamma_{n}\bigr{\}},

and note that on every γn\gamma_{n} the dual utility function U¯\bar{U}^{*} has a kink, i.e., the function U¯\bar{U}^{*} is not continuously differentiable. We insist that not every kink of U¯\bar{U}^{*} has to lie in Γ\Gamma, nor is every region of linearity of U¯\bar{U}^{**} necessarily contained in AA (e.g., when U¯=Ug\bar{U}=U\circ g is itself concave and has regions of linearity). However, by the duality relationship of U¯\bar{U}^{**} and U¯\bar{U}^{*}, we know that, for the subdifferentials

(U¯)(A)=ΓandU¯(Γ)A\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(A)=\Gamma\qquad\mbox{and}\qquad-\partial\bar{U}^{*}(\Gamma)\supseteq A (53)

holds true.

Proof (Theorem 2.2)

Given that Y^T(y)\hat{Y}_{T}(y) has a continuous law and is unique where U¯\partial\bar{U}^{*} is not vanishing, it follows that for any f1f_{1}, f2U¯f_{2}\in-\partial\bar{U}^{*} we have f1(Y^T(y))=f2(Y^T(y))f_{1}\bigl{(}\hat{Y}_{T}(y)\bigr{)}=f_{2}\bigl{(}\hat{Y}_{T}(y)\bigr{)} \mathbb{P}-a.s. Hence

W^T(x)=f(Y^T(w(x))),fU¯,\hat{W}_{T}(x)=f\bigl{(}\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)}\bigr{)},\qquad-f\in\partial\bar{U}^{*},

is \mathbb{P}-a.s. uniquely defined by a strictly increasing function ff. Since Y^T(w(x))\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)} has a continuous law, so does W^T(x)\hat{W}_{T}(x), proving a).

By the duality relationship (53) we can conclude that

[W^T(x)A]=\displaystyle\mathbb{P}\bigl{[}\hat{W}_{T}(x)\in A\bigr{]}= [f(Y^T(w(x)))A][Y^T(w(x))(U¯)(A)]\displaystyle\mathbb{P}\bigl{[}f\bigl{(}\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)}\bigr{)}\in A\bigr{]}\leq\mathbb{P}\bigl{[}\hat{Y}_{T}\bigr{(}w^{\prime}(x)\bigr{)}\in\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(A)\bigr{]} (54)
=\displaystyle= [Y^T(w(x))Γ]n=1[Y^T(w(x))=γn]=0,\displaystyle\mathbb{P}\bigl{[}\hat{Y}_{T}\bigr{(}w^{\prime}(x)\bigr{)}\in\Gamma\bigr{]}\leq\sum_{n=1}^{\infty}\mathbb{P}\bigl{[}\hat{Y}_{T}\bigr{(}w^{\prime}(x)\bigr{)}=\gamma_{n}\bigr{]}=0,

since the distribution of Y^T(y)\hat{Y}_{T}(y) has no atoms for any y>0y>0. Thus, W^T(x)\hat{W}_{T}(x) is \mathbb{P}-a.s. equal to 0 on AA. Thus, we have on one hand

w(x)=𝔼[U¯(W^T(x))]=E[U¯(W^T(x))]supX𝒳(x)𝔼[U¯(XT)]=u(x),w(x)=\mathbb{E}\bigl{[}\bar{U}^{**}\bigl{(}\hat{W}_{T}(x)\bigr{)}\bigr{]}=E\bigl{[}\bar{U}\bigl{(}\hat{W}_{T}(x)\bigr{)}\bigr{]}\leq\sup_{X\in\mathcal{X}(x)}\mathbb{E}\bigl{[}\bar{U}\bigl{(}X_{T}\bigr{)}\bigr{]}=u(x),

and on the other hand

u(x)=supX𝒳(x)𝔼[U¯(XT)]supX𝒳(x)𝔼[U¯(XT)]=𝔼[U¯(W^T(x))]=w(x).u(x)=\sup_{X\in\mathcal{X}(x)}\mathbb{E}\bigl{[}\bar{U}\bigl{(}X_{T}\bigr{)}\bigr{]}\leq\sup_{X\in\mathcal{X}(x)}\mathbb{E}\bigl{[}\bar{U}^{**}\bigl{(}X_{T}\bigr{)}\bigr{]}=\mathbb{E}\bigl{[}\bar{U}^{**}\bigl{(}\hat{W}_{T}(x)\bigr{)}\bigr{]}=w(x).

Thus, it is clear that W^(x)\hat{W}(x) is also an optimizer for the original problem, X^(x)=W^(x)\hat{X}(x)=\hat{W}(x), proving b). ∎

Note that we have said nothing about the optimal portfolio of the original problem per se, but only about the coincidence of its maximizer with that of the concavified problem. That is, the statement is as follows: when the law of the dual optimizer has no atoms, then there is no ’biduality gap’, and the original problem can be solved by considering the problem with the concavified utility function.

The following remark discusses the economic consequences of Theorem 2.2:

Remark 2

  • a)

    The optimizer X^(x)\hat{X}(x) of Theorem 2.1 satisfies X^T(x)A\hat{X}_{T}(x)\notin A, \mathbb{P}-a.s. That is, the portfolio manager flees successfully all possible outcomes that underperform the concavification.

  • b)

    Similar to the calculation in (54) we can show that the law of X^T(x)\hat{X}_{T}(x) is atomless, except possibly an atom at β\beta. Indeed, by Theorem 2.2 it is enough to show that the distribution of W^T(x)\hat{W}_{T}(x) is atomless, as it coincides with X^T(x)\hat{X}_{T}(x) a.s. Take z>βz>\beta and fU¯f\in-\partial\bar{U}^{*}, then W^T(x)=f(Y^T(w(x)))\hat{W}_{T}(x)=f\bigl{(}\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)}\bigr{)} and

    [W^T(x)=z]=[f(Y^T(w(x)))=z]=[Y^T(w(x))=(U¯)(z)]=0.\mathbb{P}\bigl{[}\hat{W}_{T}(x)=z\bigr{]}=\mathbb{P}\bigl{[}f\bigl{(}\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)}\bigr{)}=z\bigr{]}=\mathbb{P}\bigl{[}\hat{Y}_{T}\bigr{(}w^{\prime}(x)\bigr{)}=\bigl{(}\bar{U}^{**}\bigr{)}^{\prime}(z)\bigr{]}=0. (55)

    However, there is a possibility that an atom occurs at z=βz=\beta. The same calculation shows that if (U¯)(β)=(\bar{U}^{**})^{\prime}(\beta)=\infty, then the distribution of X^T(x)\hat{X}_{T}(x) cannot have an atom at β\beta. Specifically, the law X^T(x)\hat{X}_{T}(x) has an atom at β\beta if and only if (U¯)(β)<(\bar{U}^{**})^{\prime}(\beta)<\infty and [Y^T(w(x))(U¯)(β)]>0\mathbb{P}\bigl{[}\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)}\geq(\bar{U}^{**})^{\prime}(\beta)\bigr{]}>0. Moreover, in this case,

    [X^T(x)=β]=[Y^T(w(x))(U¯)(β)].\mathbb{P}[\hat{X}_{T}(x)=\beta]=\mathbb{P}\bigl{[}\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)}\geq(\bar{U}^{**})^{\prime}(\beta)\bigr{]}.

    This outcome, which occurs, for example, by pure call option payoffs in Black-Scholes markets with nonzero drift, is not very satisfactory for the investor, since the incentive scheme for the portfolio manager is such that the optimal strategy jeopardizes the whole capital with positive probability. What is worse, a call option incentive scheme leads to a higher probability of the ruin as the benchmark increases.

  • c)

    Carpenter carp also considers the case of a call option with random benchmark, g(x)=(xBT)+g(x)=(x-B_{T})^{+}. It is not to hard to integrate such options in our more general framework as long as BTL(Ω,,)B_{T}\in L^{\infty}(\Omega,\mathcal{F},\mathbb{P}), using the random endowment result of (BTZ, , Theorem 3.2.).

Proof (Proposion 1)

We know by Theorem 2.1 that the value function of the dual problem can be represented as an infimum over equivalent local martingale measures,

v(y)=infe𝔼[U¯(ydd)].v(y)=\inf_{\mathbb{Q}\in\mathcal{M}^{e}}\mathbb{E}\biggl{[}\bar{U}^{*}\biggl{(}y\frac{d\mathbb{Q}}{d\mathbb{P}}\biggr{)}\biggr{]}. (56)

Hence, we can, in particular, extract a sequence Zn{dd:e}Z^{n}\in\bigl{\{}\frac{d\mathbb{Q}}{d\mathbb{P}}\,:\,\mathbb{Q}\in\mathcal{M}^{e}\bigr{\}} so that 𝔼[U¯(yZn)]\mathbb{E}\bigl{[}\bar{U}^{*}(yZ^{n})\bigr{]} converges to v(y)v(y). Note that the sequence ZnZ^{n} is bounded in L1(Ω,,)L^{1}(\Omega,\mathcal{F},\mathbb{P}) , since the expectations of densities are bounded by one. Hence, we can apply Komlós’ Lemma ((Bogachev, , Theorem 4.27)) to find a subsequence ZnkZ^{n_{k}} and a random variable ZZ such that every subsequence ZnklZ^{n_{k_{l}}} of ZnkZ^{n_{k}} converges to ZZ, \mathbb{P}-a.s. in the sense of Cesàro. We note that ZZ is a minimizer of (56) since

𝔼[U¯(ymj=1mZnkj)]1mj=1m𝔼[U¯(yZnkj)].\mathbb{E}\biggl{[}\bar{U}^{*}\biggl{(}\frac{y}{m}\sum_{j=1}^{m}Z^{n_{k_{j}}}\biggr{)}\biggr{]}\leq\frac{1}{m}\sum_{j=1}^{m}\mathbb{E}\biggl{[}\bar{U}^{*}\Bigl{(}yZ^{n_{k_{j}}}\Bigr{)}\biggr{]}.

By the convexity of U¯\bar{U}^{*}, the right hand converges as Cesàro-subsequence of a convergent sequence to v(y)v(y). Whereas the convex combination of the random variables on the left hand is the density corresponding to some equivalent local martingale measure by the convexity of e\mathcal{M}^{e}.

Next, we assert that ZZ has a distribution, which has a continuous law. Indeed, since the laws of all the approximating ZnZ^{n} are uniformly absolutely continuous with respect to Lebesgue measure, so are the approximating Cesàro sums. Denote these sums by Z~n\tilde{Z}^{n}. Uniform absolute continuity with respect to the Lebesgue measure of the laws of Z~n\tilde{Z}^{n} implies that, for the respective cumulative distribution functions, it holds that for every ε>0\varepsilon>0 and all tt\in\mathbb{R} there exists a δ=δ(ε)>0\delta=\delta(\varepsilon)>0 such that supnsupt|FZ~n(t+δ)FZ~n(t)|<ε\sup_{n}\sup_{t\in\mathbb{R}}|F_{\tilde{Z}^{n}}(t+\delta)-F_{\tilde{Z}^{n}}(t)|<\varepsilon. We have that Z~nZ\tilde{Z}^{n}\to Z in distribution, so FZ~nFZF_{\tilde{Z}^{n}}\to F_{Z} at all points of continuity of the cumulative distribution function FZF_{Z}. To prove our assertion, it is enough to show that FZ(x)F_{Z}(x) is continuous for every xx\in\mathbb{R}. Indeed, since FZF_{Z} is increasing and bounded, it has at most countable number of discontinuity points. Take for given varepsilon>0\ varepsilon>0 some x1x_{1}, x2x_{2}\in\mathbb{R}, x1<x<x2x_{1}<x<x_{2}, such that x2x1<δ(ε3)x_{2}-x_{1}<\delta\bigl{(}\frac{\varepsilon}{3}\bigr{)}, and such that FZF_{Z} is continuous at both, x1x_{1} and x2x_{2}. Then FZ~n(x2)FZ~n(x1)<ε3F_{\tilde{Z}^{n}}(x_{2})-F_{\tilde{Z}^{n}}(x_{1})<\frac{\varepsilon}{3} for all nn\in\mathbb{N}. We can also chose nn big enough such that |FZ~n(xi)FZ(xi)|<ε3|\,F_{\tilde{Z}^{n}}(x_{i})-F_{Z}(x_{i})\,|<\frac{\varepsilon}{3}, i=1,2i=1,2. Finally, we can conclude that, for all y[x1,x2]y\in[x_{1},x_{2}]

|FZ(x)FZ(y)|FZ(x2)FZ(x1)FZ~n(x2)FZ~n(x1)+2ε3<ε.\bigl{|}\,F_{Z}(x)-F_{Z}(y)\,\bigr{|}\leq F_{Z}(x_{2})-F_{Z}(x_{1})\leq F_{\tilde{Z}^{n}}(x_{2})-F_{\tilde{Z}^{n}}(x_{1})+\frac{2\varepsilon}{3}<\varepsilon.

Thus, FZF_{Z} is continuous at xx. ∎

Remark 3

The proof becomes even simpler if one switches to the more abstract level of the bipolar theorem on L+0(Ω,,)L^{0}_{+}(\Omega,\mathcal{F},\mathbb{P}) of BS . The set of nonnegative random variables dominated by the terminal values of the processes in 𝒴(y)\mathcal{Y}(y) is the bipolar of {ydd:e}\bigl{\{}y\frac{d\mathbb{Q}}{d\mathbb{P}}\,:\,\mathbb{Q}\in\mathcal{M}^{e}\bigr{\}}, i.e., the smallest solid, convex set closed in the sense of convergence in probability that contains {ydd:e}\bigl{\{}y\frac{d\mathbb{Q}}{d\mathbb{P}}\,:\,\mathbb{Q}\in\mathcal{M}^{e}\bigr{\}}. Thus, every element in this set is given as a limit of yy times a Radon-Nikodým derivative. Thus, by Riesz’s theorem, we can extract a subsequence, which converges almost surely. Moreover, in this abstract perspective we are able to give the following interpretation. The optimizer of utility maximization under a convex incentive scheme is well-behaved (i.e., atomless) if the whole set of possible optimizers is well-behaved. Furthermore, this set is (up to a multiplicative factor) simply the bipolar of the set of Radon-Nikodým derivatives of equivalent local martingale measures. Thus, if this set is nice enough (i.e., the distribution of its elements are uniformly absolute continuous with respect to the Lebesgue measure), we always obtain a unique optimizer for utility maximization under convex incentive schemes, independent of the initial capital and the concrete choice of the incentive scheme.

8 The Original Problem: Wealth-dependent Solution

Inspired by Example 1 and, specifically by the case with zero drift in section 3.2, we try now to deduce how one can extend Theorem 2.1 to get existence and/or uniqueness results for particular initial conditions. For y>0y>0 we denote by

Δ(y)={δ>0:[Y^T(y)=δ]>0}\Delta(y)=\bigl{\{}\delta>0\,:\,\mathbb{P}\bigl{[}\hat{Y}_{T}(y)=\delta\bigr{]}>0\bigr{\}}

the at most countable set of atoms of the law of the dual optimizer Y^T(y)\hat{Y}_{T}(y). Moreover, we recall the notations

A=n=1(an,an+)={x>0:U¯(x)U¯(x)},Γ=n=1{γn},A=\bigcup_{n=1}^{\infty}(a^{-}_{n},a^{+}_{n})=\bigl{\{}x>0\,:\bar{U}(x)\neq\bar{U}^{**}(x)\bigr{\}},\qquad\Gamma=\bigcup_{n=1}^{\infty}\bigl{\{}\gamma_{n}\bigr{\}},

where γn\gamma_{n} is the slope of U¯\bar{U}^{**} on (an,an+)(a_{n}^{-},a_{n}^{+}). We are now able to make the following statement.

Theorem 8.1

The optimizer W^(x)\hat{W}(x) for the concavified problem (6) is unique for x>βx>\beta if

Δ(w(x))Γ=.\Delta\bigl{(}w^{\prime}(x)\bigr{)}\cap\Gamma=\emptyset. (57)

Moreover, in this case, X^(x)=W^(x)\hat{X}(x)=\hat{W}(x) is the unique solution to the original problem (3).

Proof

First, note that condition (57) implies that no atom of the distribution of Y^T(w(x))\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)} lies on a point in the domain of U¯\bar{U}^{*} where this function is not differentiable. Thus, we can conclude, as in the proof of Theorem 2.2, that, for f1f_{1}, f2U¯f_{2}\in-\partial\bar{U}^{*}, we have f1(Y^T(w(x)))=f2(Y^T(w(x)))f_{1}\bigl{(}\hat{Y}_{T}(w^{\prime}(x))\bigr{)}=f_{2}\bigl{(}\hat{Y}_{T}(w^{\prime}(x))\bigr{)}, \mathbb{P}-a.s. Hence

W^T(x)=f(Y^T(w(x))),fU¯,\hat{W}_{T}(x)=f\bigl{(}\hat{Y}_{T}\bigl{(}w^{\prime}(x)\bigr{)}\bigr{)},\qquad f\in-\partial\bar{U}^{*},

is \mathbb{P}-a.s. uniquely defined by a strictly increasing function ff, which proves uniqueness of W^(x)\hat{W}(x). To prove the existence of an optimizer of the original problem, we note that, from (57), we know that [Y^T(w(x))=γn]=0\mathbb{P}\bigl{[}\hat{Y}_{T}\bigr{(}w^{\prime}(x)\bigr{)}=\gamma_{n}\bigr{]}=0. Thus, similar to the proof of Theorem 2.2, we can conclude that X^T(x)\hat{X}_{T}(x) is (the unique) solution to the original problem. ∎

For the case that x>βx>\beta such that Δ(w(x))Γ\Delta\bigl{(}w^{\prime}(x)\bigr{)}\cap\Gamma\neq\emptyset, we cannot generally recover any of our results. In particular:

  • a)

    The optimizer of the concavified problem may not be unique, as discussed in the remark at the end of Example 1, Case 2.

  • b)

    It can happen that the optimum of the concavified problem is not reached by the value function of the original problem, i.e., u(x)<w(x)u(x)<w(x). An example therefore will be given below in Example 7.

  • c)

    Even if the maximum of the concavified problem can be reached by the original value function, i.e., u(x)=w(x)u(x)=w(x), it may happen that the optimizer of the original problem is not unique. To see this, we use the setting of Example 1 (with initial capital 11), changing only the incentive scheme

    gˇ(x)={x2240x6,12(x3)x>6,\check{g}(x)=\left\{\begin{array}[]{ll}\frac{x^{2}}{24}&\quad 0\leq x\leq 6,\\ \frac{1}{2}(x-3)&\quad x>6,\end{array}\right.

    which is a convex function with slope bounded by one. However, Ugˇ=U¯U\circ\check{g}=\bar{U}^{**}. Thus, all of the solutions of the concavified problem in Example 1 are also solutions to the original problem with incentive scheme gˇ\check{g}.

Example 7

To see that the optimizer of the concavified problem can be strictly bigger then any admissible terminal value for the original problem, we once again use the utility function and incentive scheme of (23) from Example 1, namely U(x)=2xU(x)=2\sqrt{x} and g(x)=14(x3)+g(x)=\frac{1}{4}(x-3)^{+}. We also take x=1x=1 as initial capital. To describe the discounted stock price process, we fix an (Ω,,)(\Omega,\mathcal{F},\mathbb{P})-measurable random variable RR that satisfies [R=2]=[R=1/2]=1/2\mathbb{P}[R=2]=\mathbb{P}[R=1/2]=1/2 and consider the process

St={10t<T/2,RT/2tT,S_{t}=\left\{\begin{array}[]{ll}1&\quad 0\leq t<T/2,\\ R&\quad T/2\leq t\leq T,\end{array}\right.

in its natural filtration. Thus, in essence, our model a disguised form of a binomial model. We note that

U¯(y)={14y2+30<y<36,[0,]y=36,0y>36,-\partial\bar{U}^{*}(y)=\left\{\begin{array}[]{ll}\frac{1}{4y^{2}}+3&\quad 0<y<\frac{\sqrt{3}}{6},\\ \left[0,\right]&\quad y=\frac{\sqrt{3}}{6},\\ 0&\quad y>\frac{\sqrt{3}}{6},\end{array}\right.

and e={}\mathcal{M}^{e}=\{\mathbb{Q}\}, where the measure \mathbb{Q} is given via the Radon-Nikodým derivative

ZT:=dd|T=231l{ST=2}+431l{ST=12},Z_{T}:=\left.\frac{d\mathbb{Q}}{d\mathbb{P}}\right|_{\mathcal{F}_{T}}=\frac{2}{3}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=2\}}+\frac{4}{3}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=\frac{1}{2}\}},

implying [R=2]=1/3\mathbb{Q}[R=2]=1/3 and [R=1/2]=2/3\mathbb{Q}[R=1/2]=2/3.

Our goal is to show that u(1)<w(1)u(1)<w(1). To compute u(1)=supX𝒳(1)𝔼[U¯(XT)]u(1)=\sup_{X\in\mathcal{X}(1)}\mathbb{E}\bigl{[}\bar{U}\bigl{(}X_{T}\bigr{)}\bigr{]}, we note that, for any predictable SS-integrable investment strategy HH

XT1,H=x+0THt𝑑St=1+HT/2(ST/2ST/2)={1+HT2R=2,1HT/22R=1/2.X_{T}^{1,H}=x+\int_{0}^{T}H_{t}\,dS_{t}=1+H_{T/2}\Bigl{(}S_{T/2}-S_{T/2-}\Bigr{)}=\left\{\begin{array}[]{ll}1+H_{\frac{T}{2}}&R=2,\\ 1-\frac{H_{T/2}}{2}&R=1/2.\end{array}\right.

Since X1,H𝒳(1)X^{1,H}\in\mathcal{X}(1) has to be nonnegative, it follows that HT/2[1,2]H_{T/2}\in[-1,2]. Hence, 0XT1,H30\leq X_{T}^{1,H}\leq 3, and we can conclude that

u(1)=supH𝔼[U¯(XT1,H)]=0.u(1)=\sup_{H}\mathbb{E}\bigl{[}\bar{U}\bigl{(}X_{T}^{1,H}\bigr{)}\bigr{]}=0.

For the calculation of w(1)w(1) we use the fact that, in a complete market, e={}\mathcal{M}^{e}=\{\mathbb{Q}\}. Thus, the dual value function can be directly computed via the unique dual optimizer Y^T(y)=yZT\hat{Y}_{T}(y)=yZ_{T},

v(y)=infQe𝔼[U¯(ydQdP)]=𝔼[U¯(yZT)]={932y3y0<y38,316yy38<y<34,0y34.v(y)=\inf_{Q\in\mathcal{M}^{e}}\mathbb{E}\biggl{[}\bar{U}^{*}\biggl{(}y\frac{dQ}{dP}\biggr{)}\biggr{]}=\mathbb{E}\Bigl{[}\bar{U}^{*}(yZ_{T})\Bigr{]}=\left\{\begin{array}[]{ll}\frac{9}{32y}-3y&\quad 0<y\leq\frac{\sqrt{3}}{8},\\ \frac{3}{16y}-y&\quad\frac{\sqrt{3}}{8}<y<\frac{\sqrt{3}}{4},\\ 0&\quad y\geq\frac{\sqrt{3}}{4}.\end{array}\right.

Now, calculating the subdifferential,

v(y)={932y2+30<y<38,[5,9]y=38,316y2+138<y<34,[0,2]y=34,0y34,-\partial v(y)=\left\{\begin{array}[]{ll}\frac{9}{32y^{2}}+3&\quad 0<y<\frac{\sqrt{3}}{8},\\ \left[5,9\right]&\quad y=\frac{\sqrt{3}}{8},\\ \frac{3}{16y^{2}}+1&\quad\frac{\sqrt{3}}{8}<y<\frac{\sqrt{3}}{4},\\ \left[0,2\right]&\quad y=\frac{\sqrt{3}}{4},\\ 0&\quad y\geq\frac{\sqrt{3}}{4},\end{array}\right.

and using by convex duality that y=w(x)y=w^{\prime}(x) if and only if xv(y)x\in-\partial v(y), we conclude that for x=1x=1 it follows that w(1)=3/4w^{\prime}(1)=\sqrt{3}/4. Thus, Theorem 5.2 implies that

W^T(1)(U¯)(Y^1(w(1)))\displaystyle\hat{W}_{T}(1)\in-\bigl{(}\partial\bar{U}^{*}\bigr{)}\Bigl{(}\hat{Y}_{1}\bigl{(}w^{\prime}(1)\bigr{)}\Bigr{)} =(U¯)(361l{ST=2}+331l{ST=12})\displaystyle=-\bigl{(}\partial\bar{U}^{*}\bigr{)}\biggr{(}\frac{\sqrt{3}}{6}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=2\}}+\frac{\sqrt{3}}{3}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=\frac{1}{2}\}}\biggr{)}
=[0,6]1l{ST=2}+{0}1l{ST=12}\displaystyle=[0,6]{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=2\}}+\{0\}{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=\frac{1}{2}\}}

and we can conclude by the admissibility constraint 𝔼[W^T(1)Y^1(w(1))]=w(1)\mathbb{E}\bigl{[}\hat{W}_{T}(1)\hat{Y}_{1}\bigl{(}w^{\prime}(1)\big{)}\bigr{]}=w^{\prime}(1) that

W^T(1)=31l{ST=2}+01l{ST=12}.\hat{W}_{T}(1)=3{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=2\}}+0{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=\frac{1}{2}\}}.

This can be seen also in a simpler way. Since XT1,H=x+0THt𝑑StX_{T}^{1,H}=x+\int_{0}^{T}H_{t}\,dS_{t} depends only on HT/2H_{T/2} which, by predictability, has to be T/2=0\mathcal{F}_{T/2-}=\mathcal{F}_{0}-measurable and hence constant. We have by admissibility 1H2-1\leq H\leq 2. Hence

w(1)=supH𝔼[U¯(XT1,H(1))]=supH36([R=2](1+H)+[R=1/2](1H/2))=34.w(1)=\sup_{H}\mathbb{E}\Bigl{[}\bar{U}^{**}\bigl{(}X_{T}^{1,H}(1)\bigr{)}\Bigr{]}=\sup_{H}\frac{\sqrt{3}}{6}\Bigl{(}\mathbb{P}[R=2](1+H)+\mathbb{P}[R=1/2](1-H/2)\Bigr{)}=\frac{\sqrt{3}}{4}.

The maximum is achieved with H=2H=2, i.e., the optimal portfolio is W^T(1)=31l{ST=2}+01l{ST=12}\hat{W}_{T}(1)=3{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=2\}}+0{\mathchoice{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.0mu\mathrm{l}}{1\mskip-4.5mu\mathrm{l}}{1\mskip-5.0mu\mathrm{l}}}_{\{S_{T}=\frac{1}{2}\}}. It follows in either case that w(1)=34w(1)=\frac{\sqrt{3}}{4}. Thus, we conclude that 0=u(1)<w(1)=340=u(1)<w(1)=\frac{\sqrt{3}}{4}.

Note, finally, that such behavior can be excluded in the case of complete markets; Reichlin (Rei, , Section 5) shows that the optimizer of the concavified problem is an optimizer of the original problem if the underlying probability space is atomless.

9 Conclusion

We have considered the non-concave utility maximization problem as seen from the point of view of a fund manager, who manages the capital for an investor and who is compensated by a convex incentive scheme. We have proved the existence and uniqueness of the dual optimizer and also proved the existence and uniqueness of the original problem for arbitrary initial capital in case in which the dual optimizer has a continuous distribution. We have shown that this is true in a large class of (possibly incomplete) market models, independent of the specific incentive scheme. When this condition fails, we have proved the existence of a unique solution for the concavified problem and shown that this solution is also a solution of the original problem under additional assumptions on the initial capital. However, there are models, where, for some initial capital, the optimal value of the concavified problem cannot be reached, as we have demonstrated through a counterexample. Moreover, we have illustrated our findings by specific examples, which in essence contain the explicit solution strategies for complete markets. Finally, we have discussed the economic implications of our findings.

Acknowledgements.
Both authors acknowledge partial financial supported by NSF grant DMS-0739195 and want to thank René Carmona for suggesting the problem under consideration and steady encouragement.They are thankful to two anonymous referees and an associate editor for thoughtful remarks and comments that have improved the quality of the article. Also thanks to Gerard Brunick, Christian Reichlin, Ronnie Sircar and Ramon van Handel for helpful discussions and comments. Thanks to Matt Lorig, who put a lot of effort to improve the readability and grammatical correctness of the text.

References

  • (1) Alòs, E., Ewald, C.O.: Malliavin differentiability of the Heston volatility and applications to option pricing. Adv. in Appl. Probab. 40(1), 144–162 (2008)
  • (2) Amendinger, J.: Martingale representation theorems for initially enlarged filtrations. Stochastic Process. Appl. 89(1), 101–116 (2000)
  • (3) Bogachev, V.I.: Measure theory. Vol. I and II. Berlin: Springer (2007)
  • (4) Bouchard, B., Touzi, N., Zeghal, A.: Dual formulation of the utility maximization problem: the case of nonsmooth utility. Ann. Appl. Probab. 14(2), 678–717 (2004)
  • (5) Brannath, W., Schachermayer, W.: A bipolar theorem for L+0(Ω,,P)L^{0}_{+}(\Omega,\mathcal{F},P). In: Séminaire de Probabilités, XXXIII, Lecture Notes in Math., vol. 1709, pp. 349–354. Springer, Berlin (1999)
  • (6) Brigo, D., Mercurio, F.: Interest rate models—theory and practice, second edn. Springer Finance. Springer-Verlag, Berlin (2006)
  • (7) Carassus, L., Pham, H.: Portfolio optimization for piecewise concave criteria functions (the 8th workshop on stochastic numerics). RIMS Kokyuroku 1620, 81–108 (2009)
  • (8) Carpenter, J.N.: Does option compensation increase managerial risk appetite? The Journal of Finance 55(5), 2311–2331 (2000)
  • (9) Deelstra, G., Pham, H., Touzi, N.: Dual formulation of the utility maximization problem under transaction costs. Ann. Appl. Probab. 11(4), 1353–1383 (2001)
  • (10) Delbaen, F., Schachermayer, W.: The mathematics of arbitrage. Springer Finance. Springer-Verlag, Berlin (2006)
  • (11) Fouque, J., Papanicolaou, G., Sircar, R., Sølna, K.: Multiscale stochastic volatility for equity, interest rate, and credit derivatives. Cambridge University Press, Cambridge (2011)
  • (12) Frey, R.: Derivative asset analysis in models with level-dependent and stochastic volatility. CWI Quarterly 10(1), 1–34 (1997). Mathematics of finance, Part II
  • (13) Guasoni, P., Obłój, J.K.: The incentives of hedge fund fees and high-water marks. to appear in Math. Finance
  • (14) Hiriart-Urruty, J.B., Lemaréchal, C.: Fundamentals of convex analysis. Grundlehren Text Editions. Springer-Verlag, Berlin (2001). Abridged version of Convex analysis and minimization algorithms I and II
  • (15) Janeček, K., Sîrbu, M.: Optimal investment with high-watermark performance fee. SIAM J. Control Optim. 50(2), 790–819 (2012)
  • (16) Karatzas, I., Lehoczky, J.P., Shreve, S.E.: Optimal portfolio and consumption decisions for a ’small investor’ on a finite horizon. SIAM J. Control Optim. 25, 1557–1586 (1987)
  • (17) Kramkov, D., Schachermayer, W.: The asymptotic elasticity of utility functions and optimal investment in incomplete markets. Ann. Appl. Probab. 9(3), 904–950 (1999)
  • (18) Kramkov, D., Schachermayer., W.: Necessary and sufficient conditions in the problem of optimal investment in incomplete markets. Ann. Appl. Probab. 13(4), 1504–1516 (2003)
  • (19) Larsen, K.: Optimal portfolio delegation when parties have different coefficients of risk aversion. Quant. Finance 5(5), 503–512 (2005)
  • (20) Merton, R.C.: Optimum consumption and portfolio rules in a continuous-time model. Journal of Economic Theory 3(4), 373–413 (1971)
  • (21) Nualart, D.: Malliavin calculus and its applications, CBMS Regional Conference Series in Mathematics, vol. 110. Published for the Conference Board of the Mathematical Sciences, Washington, DC (2009)
  • (22) Nualart, E.: Lectures on Malliavin calculus and its applications. Lecture Notes (2011)
  • (23) Ocone, D.L., Karatzas, I.: A generalized Clark representation formula, with application to optimal portfolios. Stochastics Stochastics Rep. 34(3-4), 187–220 (1991)
  • (24) Panageas, S., Westerfield, M.M.: High-water marks: High risk appetites? Convex compensation, long horizons, and portfolio choice. The Journal of Finance 64(1), 1–36 (2009)
  • (25) Pliska, S.R.: A stochastic calculus model of continuous trading: optimal portfolios. Math. Oper. Res. 11(2), 370–382 (1986)
  • (26) Reichlin, C.: Utility maximization with a given pricing measure when the utility is not necessarily concave. Math. Financ. Econ. 7, 531–556 (2013)
  • (27) Ross, S.A.: Compensation, incentives, and the duality of risk aversion and riskiness. The Journal of Finance 59(1), 207–225 (2004)
  • (28) Seifried, F.T.: Optimal investment with deferred capital gains taxes: a simple martingale method approach. Math. Methods Oper. Res. 71(1), 181–199 (2010)
  • (29) Siu, T.K.: Regime-switching risk: to price or not to price? Int. J. Stoch. Anal. pp. 1–14 (2011)
  • (30) Westray, N., Zheng, H.: Constrained nonsmooth utility maximization without quadratic inf convolution. Stochastic Process. Appl. 119(5), 1561–1579 (2009)
  • (31) Westray, N., Zheng, H.: Minimal sufficient conditions for a primal optimizer in nonsmooth utility maximization. Finance Stoch. 15(3), 501–512 (2011)
  • (32) Wong, B., Heyde, C.C.: On changes of measure in stochastic volatility models. J. Appl. Math. Stoch. Anal. (2006)