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

Power Utility Maximization in Exponential Lévy Models: Convergence of Discrete-Time to Continuous-Time Maximizers

Johannes P. Temme Fakultät für Mathematik, Universität WienNordbergstraße 15
1090 Wien, Austria
johannes.temme@univie.ac.at
(Date: September 3, 2025)
Abstract.

We consider power utility maximization of terminal wealth in a 1-dimensional continuous-time exponential Lévy model with finite time horizon. We discretize the model by restricting portfolio adjustments to an equidistant discrete time grid. Under minimal assumptions we prove convergence of the optimal discrete-time strategies to the continuous-time counterpart. In addition, we provide and compare qualitative properties of the discrete-time and continuous-time optimizers.

Key words and phrases:
utility maximization, power utility, exponential Lévy process, discretization
2000 Mathematics Subject Classification:
Primary 91B28, 91B16, secondary 60G51
The author gratefully acknowledges financial support from the Austrian Science Fund (FWF) under grant P19456. The author thanks two anonymous referees for their valuable input. He also thanks Johannes Muhle-Karbe and Mathias Beiglböck for fruitful discussions and comments on the draft. The final publication is available at springerlink.com

1. Introduction

In exponential Lévy models the stock is given – just as its name suggests – as the (stochastic) exponential of a Lévy process. Continuous-time exponential Lévy models and their discrete-time counterparts share the defining property of independent and identically distributed logarithmic returns. Exponential Lévy models have become widely popular in the last decade since they are analytically tractable and provide a reasonable approximation to financial data.

We consider maximization of expected utility from terminal wealth with finite time horizon for an investor equipped with a power utility function U(x)=(1p)1x1pU(x)=(1-p)^{-1}x^{1-p}, p>0p>0, in the framework of one bond (normalized to 1) and one stock which follows an exponential Lévy process. The problem of maximizing expected power utility from terminal wealth was first studied by Merton ([15]) for Brownian motion with drift and by Samuelson ([20]) for the discrete-time analogon in an NN-period model. The optimal portfolio selection for general time-continuous exponential Lévy processes was investigated in [5], [10] and [2]. Nutz ([16], for p1p\neq 1), as well as Kardaras ([12], for p=1p=1) have recently shown under minimal assumptions that the optimal fraction111Following a usual practice for power utiltiy functions, we note that we interpret in this exposition trading strategies as the fraction of current wealth invested in the stock, rather than the amount of shares the agent holds. Thus, for a given trading strategy πt\pi_{t} the agent invests the fraction πt\pi_{t} of her current wealth in the stock, and the fraction 1πt1-\pi_{t} of current wealth in the bond. π\pi^{*} of wealth invested in the continuous-time exponential Lévy model is constant and given as the maximizer of a deterministic concave function gg, which is defined in terms of the Lévy triplet (also cf., e.g., [6] for sufficient conditions in a general setting, and [4] for sufficient conditions on the optimality of constant trading strategies). Similarly, the optimal trading strategy for an NN-period exponential Lévy model is to invest a constant fraction πN\pi^{*}_{N} of current wealth in each step. This stems from the fact that the logarithmic return of the stock in each period is i.i.d. and consequently the NN-period model becomes an iterated 1-period model. The optimal fraction πN\pi_{N}^{*} is given as the maximizer of a deterministic concave function gNg^{N}, where gNg^{N} only depends on the evolution of the Lévy process in the first step.

In this exposition we assume that a continuous-time exponential Lévy model is given. We define the discrete-time NN-period exponential Lévy model as the restriction of the continuous-time model to discrete time instants of distance TN\frac{T}{N}, where TT is the finite time horizon. In particular, the discrete logarithmic returns are given by i.i.d. random variables. Hence, trading in this discrete-time model amounts to trading in the original continuous-time model with the restriction that the portfolio can only be adjusted at given time instant, i.e., if t=kTNt=\frac{kT}{N}, k=0,,Nk=0,\ldots,N. This restriction also involves a discretization gap in the sense that the set of admissible trading strategies for the NN-period model is in general strictly smaller than the respective set of the continuous-time model. For this reason the optimal fraction π\pi^{*} invested in the stock of the continuous-time model might in general be not admissible for the discrete-time models. This somewhat surprising consequence was already mentioned by Rogers ([19]) in the case of the classical Merton problem, which we consider in more detail below.

If the Lévy process has a non-zero Brownian motion part or allows both positive and negative jumps, it is easily seen that the interval of admissible constant trading strategies of all NN-period models is given by [0,1][0,1]. This means that the agent is allowed to invest in each period the fraction of 0% to 100% of current wealth in the stock. Short-selling and investing more than 100 % of current wealth is prohibited since negative wealth is not allowed by power utility investors. In the other (less important) case of a pure jump process allowing either only positive or only negative jumps, the interval of admissible trading strategies for the NN-period model is more complicated and depends on the number of discretization points NN. Although not technically necessary, in order to simplify the further discussion we will assume throughout this exposition that the Lévy process has a non-zero Brownian motion part or allows both positive and negative jumps. Hence, the interval of admissible constant NN-period trading strategies is in particular given by [0,1][0,1]. We comment separately on the case of a pure jump process with either only positive or only negative jumps.

The objective of this exposition is to answer the following two questions:

  1. 1.)

    Do the optimal discrete-time strategies πN\pi_{N}^{*} converge to the optimal continuous-time strategy π\pi^{*} as NN\to\infty, i.e., as the number of discretization points increases?

  2. 2.)

    How does the sign of the drift of the Lévy process and the risk aversion of the power utility funcion affect the optimal continuous-time strategy π\pi^{*} and the discrete-time strategies πN\pi_{N}^{*}?

Clearly, a necessary assumption for the convergence of πN[0,1]\pi_{N}^{*}\in[0,1] to π\pi^{*} is π[0,1]\pi^{*}\in[0,1]. But π[0,1]\pi^{*}\in[0,1] generally fails to hold already in the Black-Scholes model due to the previously mentioned discretization gap: in the classical Merton problem the Lévy process is given by Lt=μt+σBtL_{t}=\mu t+\sigma B_{t}, where BB denotes Brownian motion. Let U(x)=(1p)1x1pU(x)=(1-p)^{-1}x^{1-p} with p>0p>0, then the optimal continuous-time strategy is given by the constant π=μσ2p\pi^{*}=\frac{\mu}{\sigma^{2}p}, cf. [15]. Thus, π[0,1]\pi^{*}\notin[0,1] if e.g. μ\mu is chosen large enough and πN\pi_{N}^{*} will not converge to π\pi^{*}.

To cope with this problem we introduce in the continuous-time exponential Lévy model the exogenous portfolio constraint 𝒞=[0,1]\mathscr{C}=[0,1]. I.e., in this constrained continuous-time model the agent may only invest between 0% and 100100% of her current wealth in the stock, which is the same restriction as in the NN-period models. The optimal continuous-time strategy π𝒞\pi_{\mathscr{C}}^{*} of the constrained model can be found similarly as in the unconstrained case as the maximium of the same function gg on 𝒞=[0,1]\mathscr{C}=[0,1]. In particular, π[0,1]\pi^{*}\in[0,1] implies π=π𝒞\pi^{*}=\pi^{*}_{\mathscr{C}}.

Coming back to our two questions we see that we have to refine the first one due to the discretization gap and rather ask

  1. 1.)’

    Do the optimal discrete-time strategies πN\pi_{N}^{*} converge to the optimal strategy π𝒞\pi^{*}_{\mathscr{C}} of the constrained continuous-time model as NN\to\infty, i.e., as the number of discretization points increases?

In order to respond to both questions we observe that we can access π𝒞\pi^{*}_{\mathscr{C}} and πN\pi_{N}^{*} as the maximizer of the previously mentioned deterministic and concave functions g|[0,1]g|_{[0,1]} and gNg^{N}, respectively. A major part in answering the questions is done in Theorem 4.1 below, which shows that gNg^{N} converges uniformly on [0,1][0,1] to gg as NN\to\infty. Our main results can then be summarized as follows:

Main Results (Theorem 4.1, Corollaries 4.2 and 4.3, Proposition 4.4).

Let π\pi^{*}, πN\pi_{N}^{*} and π𝒞\pi^{*}_{\mathscr{C}} denote the optimal fraction of current wealth invested in the stock for maximizing expected power utility U(x)=(1p)1x1pU(x)=(1-p)^{-1}x^{1-p}, p>0p>0, from terminal wealth in the unconstrained continuous-time exponential Lévy model, its NN-period discretization and the constrained continuous-time exponential Lévy model with exogenous portfolio constraints 𝒞=[0,1],\mathscr{C}=[0,1], respectively.

Then

  1. (i)

    limNπN=π𝒞\lim_{N\to\infty}\pi_{N}^{*}=\pi^{*}_{\mathscr{C}},

  2. (ii)

    the terminal expected utility of the NN-period model converges to the terminal expected utility of the constrained continuous-time model as NN\to\infty, and

  3. (iii)

    if additionally the Lévy process is square integrable, then the optimal NN-period terminal payoffs converge in L2(Ω)L^{2}(\Omega) to the optimal terminal payoff of the constrained continuous-time model as NN\to\infty.

Hence, if π[0,1]\pi^{*}\in[0,1], i.e., if π\pi^{*} is admissible in the NN-period models, (i)-(iii) hold for the unconstrained continuous-time exponential Lévy model.

We also find the following qualitative properties of the optimal unconstrained continuous-time and NN-period strategies under the assumption that the Lévy process is integrable:

  1. (iv)

    π,πN\pi^{*},\pi^{*}_{N} are non-negative (non-positive) if the drift of the Lévy process – i.e., the expectation of the Lévy process – is non-negative (non-positive). Both |π||\pi^{*}| and πN[0,1]\pi^{*}_{N}\in[0,1] are increasing if the relative risk aversion pp of UU is decreasing.222Note that π\pi^{*} as well as πN\pi^{*}_{N} depend on U(x)=(1p)1x1pU(x)=(1-p)^{-1}x^{1-p} and can be considered as functions of pp

In the light of our results, we see that the exogenous portfolio constraint 𝒞=[0,1]\mathscr{C}=[0,1] is a natural assumption to make the continuous-time model resemble the discrete-time models. As we have already mentioned, π[0,1]\pi^{*}\in[0,1] implies π=π𝒞\pi^{*}=\pi^{*}_{\mathscr{C}}. On the other hand, if π[0,1]\pi^{*}\notin[0,1] then π𝒞\pi_{\mathscr{C}}^{*} and πN\pi^{*}_{N} are easily found: if π<0\pi^{*}<0 then πN=π𝒞=0\pi_{N}^{*}=\pi_{\mathscr{C}}^{*}=0, and if π>1\pi^{*}>1 then πN=π𝒞=1\pi_{N}^{*}=\pi_{\mathscr{C}}^{*}=1.

The approximation of a continuous-time Lévy model by discretizations has been covered under different aspects in the literature. Sufficient conditions for the convergence of optimal trading strategies in certain diffusion models and their discrete approximations were found in [7]. Rogers ([19]) compared optimal investors of a diffusion model and its discretization in terms of efficiency and his work was extended in [1] to the case of partially available information. Another aspect of discretization in terms of time-lagged trading and its impact on utility maximzation was discussed in [18] for exponential Lévy models.

The paper is organized as follows. In Section 1 we specify the continuous-time and discrete-time exponential Lévy models, as well as the optimization problem of maximizing expected utility from terminal wealth. We state the necessary assumptions (Assumption 2.1) for our results and recall certain findings of [16], as well as some properties of stochastic exponentials of Lévy processes that we shall frequently use. Section 3 summarizes important analytic properties of the deterministic functions gg and gNg^{N}. In Section 4 we state and prove our main results. We also comment on the extension of Theorem 4.1 to higher dimensional exponential Lévy processes. Certain technical results needed for the convergence of the wealth processes (Corollary 4.3) are presented in the Appendix.

It is important to mention that our convergence theorem (Theorem 4.1) relies on the work of [8] and [3] on moment asymptotics for Lévy processes. We also use results on the finiteness of gg, that are proved in [16]. The convergence results in the Appendix are extensions of the results in [13] on Euler discretizations.

2. Preliminaries and Assumptions

We fix a finite time horizon T>0T>0 and a filtered probability space (Ω,,(t)t0,)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\geq 0},{\mathbb{P}}) satisfying the usual assumptions of right-continuity and completeness. The exponential Lévy model of one bond normalized to 11 and one stock SS with initial value S0>0S_{0}>0 is given as the solution of the SDE

(2.1) St=S0+0tSu𝑑Lu,S_{t}=S_{0}+\int_{0}^{t}S_{u-}\;dL_{u},

where (Lt)t(L_{t})_{t} is a Lévy process with Lévy-Khintchine triplet (b(h),c,F)(b(h),c,F) relative to a continuously differentiable truncation function hh, with drift b(h)b(h)\in\mathbb{R}, diffusion c>0c>0 and Lévy measure FF on \mathbb{R}. We refer to [9, II.4] for more information on Lévy processes.

By (2.1), SS is given by the stochastic exponential of LL, i.e., St=S0(L)tS_{t}=S_{0}\mathcal{E}\left(L\right)_{t}. The agent is endowed with initial capital x0>0x_{0}>0 and her preferences are given by a power utility function U(x)=x1p1pU(x)=\frac{x^{1-p}}{1-p} with p>0p>0, where we set throughout this exposition x00:=log(x)\frac{x^{0}}{0}:=\log(x), to cover also the case of logarithmic utility. Let 𝒞\mathcal{C}\subseteq\mathbb{R} denote an interval of portfolio constraints. Throughout this exposition we are only interested in the unconstrained case 𝒞=\mathcal{C}=\mathbb{R} and in the exogenous constraint 𝒞=𝒞:=[0,1]\mathcal{C}=\mathscr{C}:=[0,1].

We define the set of all admissible constant trading strategies by the convex set

(2.2) 𝒜0\displaystyle\mathcal{A}_{0} :={π|F(x:1+πx<0)=0}if p(0,1),\displaystyle:=\{\pi\in\mathbb{R}|F\left(x\in\mathbb{R}:1+\pi x<0\right)=0\}\quad\text{if }p\in(0,1),
(2.3) 𝒜0,\displaystyle\mathcal{A}_{0,*} :={π|F(x:1+πx0)=0}if p1,\displaystyle:=\{\pi\in\mathbb{R}|F\left(x\in\mathbb{R}:1+\pi x\leq 0\right)=0\}\quad\text{if }p\geq 1,

where π\pi denotes the fraction of wealth invested in the stock. Since we fix the utility function UU, we denote the corresponding set of admissible constant trading strategies by 𝒜\mathcal{A} to simplify further notation, while we still have in mind that we mean either of (2.2), (2.3). The requirements in (2.2) and (2.3), respectively, ensure that x0(πL)t0(>0)x_{0}\mathcal{E}\left(\pi L\right)_{t}\geq 0\;(>0) holds for all π𝒜0(π𝒜0,)\pi\in\mathcal{A}_{0}\;(\pi\in\mathcal{A}_{0,*}) and U(x0(πL)T)>U(x_{0}\mathcal{E}\left(\pi L\right)_{T})>-\infty, cf. [17, Lemma 2.5]. Here, x0(πL)tx_{0}\mathcal{E}\left(\pi L\right)_{t} is the wealth process corresponding to the constant trading strategy π𝒜\pi\in\mathcal{A}, i.e., it is the solution of

(2.4) Xt=x0+0tπXu𝑑Lu.X_{t}=x_{0}+\int_{0}^{t}\pi X_{u-}\;dL_{u}.

We also consider non-constant admissible trading strategies πt(ω)\pi_{t}(\omega) given by

𝒯\displaystyle\mathcal{T} :={(πt)t|π pred. and L-integrable, πt(ω)𝒞,(πL)t>0} if p(0,1),\displaystyle:=\{(\pi_{t})_{t}\rvert\,\pi\text{ pred.\ and }L\text{-integrable, }\pi_{t}(\omega)\in\mathcal{C},\,\mathcal{E}\left(\pi\;\begin{picture}(1.0,1.0)(0.0,-3.0)\circle*{3.0}\end{picture}\;L\right)_{t}>0\}\text{ if }p\in(0,1),
𝒯\displaystyle\mathcal{T} :={(πt)t|π pred. and L-integrable, πt(ω)𝒞,(πL)t>0} if p1,\displaystyle:=\{(\pi_{t})_{t}\rvert\,\pi\text{ pred.\ and }L\text{-integrable, }\pi_{t}(\omega)\in\mathcal{C},\,\mathcal{E}\left(\pi\;\begin{picture}(1.0,1.0)(0.0,-3.0)\circle*{3.0}\end{picture}\;L\right)_{t}>0\}\text{ if }p\geq 1,

where πL\pi\;\begin{picture}(1.0,1.0)(0.0,-3.0)\circle*{3.0}\end{picture}\;L denotes the stochastic integral of πt\pi_{t} w.r.t. LL.

It is the agent’s aim to find an optimal trading strategy πt\pi^{*}_{t} for the problem of maximizing expected utility problem from terminal wealth on [0,T][0,T]

(2.5) u(x0):=supπ𝒯𝔼[U(x0(πL)T)]=supπ𝒯𝔼[(x0(πL)T)1p1p].u(x_{0}):=\sup_{\pi\in\mathcal{T}}\mathbb{E}\left[U(x_{0}\mathcal{E}\left(\pi\;\begin{picture}(1.0,1.0)(0.0,-3.0)\circle*{3.0}\end{picture}\;L\right)_{T})\right]=\sup_{\pi\in\mathcal{T}}\mathbb{E}\left[\frac{(x_{0}\mathcal{E}\left(\pi\;\begin{picture}(1.0,1.0)(0.0,-3.0)\circle*{3.0}\end{picture}\;L\right)_{T})^{1-p}}{1-p}\right].

In order to discuss the optimal trading strategy πt\pi^{*}_{t} in more detail we now state the assumptions that we make throughout this exposition:

Assumption 2.1.
  1. (i)

    S>0S>0.

  2. (ii)

    c0c\neq 0 or both F()>0F(\mathbb{R}_{-})>0 and F(+)>0F(\mathbb{R}_{+})>0.

  3. (iii)

    u(x0)<u(x_{0})<\infty, i.e., the maximization problem (2.5) is finite.

By [6, Lemma A.8] assumption (i) implies F(,1]=0F(-\infty,-1]=0 and (i) is in fact equivalent to the existence of a Lévy process L~\widetilde{L} such that St=(L)t=exp(L~t)S_{t}=\mathcal{E}\left(L\right)_{t}=\exp(\widetilde{L}_{t}). [6, Lemma A.8] also gives the following formulae to determine the Lévy triplet (b~(h),c~,F~)(\widetilde{b}(h),\widetilde{c},\widetilde{F}) of L~\widetilde{L}

(2.6) b~(h)\displaystyle\widetilde{b}(h) =b(h)c2+(1,)(h(log(1+x))h(x))𝑑F(x),\displaystyle=b(h)-\frac{c}{2}+\int_{(-1,\infty)}\big{(}h(\log(1+x))-h(x)\big{)}\;dF(x),
(2.7) c~\displaystyle\widetilde{c} =c,\displaystyle=c,
(2.8) F~(G)\displaystyle\widetilde{F}(G) =1G(log(1+x))𝑑F(x)for all Borel sets G,\displaystyle=\int 1_{G}(\log(1+x))\;dF(x)\quad\text{for all Borel sets }G\in\mathcal{B},

where hh is a bounded truncation function, i.e., h(x)=0h(x)=0 for |x|>R>0|x|>R>0. From now on we fix a bounded and continuously differentiable trunction function hh and write in the sequel – if not stated differently – b:=b(h)b:=b(h), b~:=b~(h)\widetilde{b}:=\widetilde{b}(h) to alleviate notation.

As already mentioned in the introduction, by (ii) we make the assumption that LL has a non-zero Brownian motion part or that LL has both positive and negative jumps. Although this is not technically necessary it simplifies further discussion on the admissibility of NN-period trading strategies below. In particular, assumption (ii) assures a no arbitrage condition known as no unbounded increasing profit, which has been introduced in [12]. We comment on the less important case of a Lévy process with c=0c=0 and either only positive or only negative jumps at the end of this section.

Assumption (iii) is a common assumption assuring existence and uniqueness of a maximizer of (2.5), i.e., the sup\sup in (2.5) is in fact a max\max.

Under Assumption 2.1, the optimal trading strategy πt\pi^{*}_{t} of (2.5) becomes trackable and is of a simple structure: Nutz (cf. [16, Theorem 3.2 and Remark 3.3], for p1p\neq 1) and Kardaras (cf. [12, Lemma 5.1], for p=1p=1) have shown that under our assumptions the optimal trading strategy π\pi^{*} is unique and constant, π𝒜𝒞\pi^{*}\in\mathcal{A}\cap\mathcal{C}. In the following, we denote the optimal strategy of the unconstrained model (𝒞=\mathcal{C}=\mathbb{R}) by π\pi^{*} and the optimal strategy under the exogenous portfolio constraint 𝒞=𝒞=[0,1]\mathcal{C}=\mathscr{C}=[0,1] by π𝒞\pi^{*}_{\mathscr{C}}, which we refer to as the optimal strategy of the constrained continuous-time model.

The optimal wealth processes are given by x0(πL)x_{0}\mathcal{E}\left(\pi^{*}L\right), x0(π𝒞L)x_{0}\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right) and an investigation of [14] even shows that the Inada conditions upon UU ensure

(2.9) (πL)t>0t[0,T],\mathcal{E}\left(\pi^{*}L\right)_{t}>0\quad t\in[0,T],

i.e., π𝒜0,\pi^{*}\in\mathcal{A}_{0,*} for all p>0p>0. (π𝒞L)t>0\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right)_{t}>0 holds true since 𝒜[0,1]=[0,1]\mathcal{A}\cap[0,1]=[0,1] and (πL)t>0\mathcal{E}\left(\pi L\right)_{t}>0 for all π[0,1]\pi\in[0,1]. [16, Theorem 3.2] (for p1p\neq 1) and [11, Lemma 5.1] (for p=1p=1) show that π\pi^{*}, π𝒞\pi^{*}_{\mathscr{C}} are given as the unique maximizer of the deterministic concave function g:𝒜𝒞g:\mathcal{A}\cap\mathcal{C}\to\mathbb{R} defined by

(2.10) g(π):=πbpπ2c2+(1,)((1+πx)1p11pπh(x))𝑑F(x)for p1,g(π):=πbpπ2c2+(1,)(log(1+πx)πh(x))𝑑F(x)for p=1,\displaystyle\begin{split}g(\pi)&:=\pi b-\frac{p\pi^{2}c}{2}+\int_{(-1,\infty)}\left(\frac{\left(1+\pi x\right)^{1-p}-1}{1-p}-\pi h(x)\right)\;dF(x)\quad\text{for }p\neq 1,\\ g(\pi)&:=\pi b-\frac{p\pi^{2}c}{2}+\int_{(-1,\infty)}\left(\log(1+\pi x)-\pi h(x)\right)\;dF(x)\quad\text{for }p=1,\end{split}

where we have used F(,1]=0F(-\infty,-1]=0. In particular, the function gg corresponding to the constrained case 𝒞=[0,1]\mathscr{C}=[0,1] is the restriction on [0,1][0,1] of the function gg corresponding to unconstrained case 𝒞=\mathcal{C}=\mathbb{R}. This also shows that π[0,1]\pi^{*}\in[0,1] implies π=π𝒞\pi^{*}=\pi^{*}_{\mathscr{C}}. Moreover, the concavity of gg shows that π>1\pi^{*}>1 (π<0\pi^{*}<0) implies π𝒞=1\pi^{*}_{\mathscr{C}}=1 (π𝒞=0\pi^{*}_{\mathscr{C}}=0).

(2.10) implies that π\pi^{*}, π𝒞\pi^{*}_{\mathscr{C}} are independent of the initial capital x0x_{0}, which is a well known property of power utility functions even in more general market models. By [16, Theorem 3.2] the optimal wealth processes x0(πL)tx_{0}\mathcal{E}\left(\pi^{*}L\right)_{t} and x0(π𝒞L)tx_{0}\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right)_{t}, respectively, lead to the terminal utility

(2.11) 𝔼[U(x0(πL)T)]=x01p1pe(1p)g(π)Tfor π=π,π𝒞 and p1,𝔼[U(x0(πL)T)]=log(x0)+g(π)Tfor π=π,π𝒞 and p=1.\displaystyle\begin{split}\mathbb{E}\left[U(x_{0}\mathcal{E}\left(\pi L\right)_{T})\right]&=\frac{x_{0}^{1-p}}{1-p}e^{(1-p)g(\pi)T}\quad\text{for }\pi=\pi^{*},\pi^{*}_{\mathscr{C}}\text{ and }p\neq 1,\\ \mathbb{E}\left[U(x_{0}\mathcal{E}\left(\pi L\right)_{T})\right]&=\log(x_{0})+g(\pi)T\quad\text{for }\pi=\pi^{*},\pi^{*}_{\mathscr{C}}\text{ and }p=1.\end{split}

We now construct the NN-period model associated to (2.1): let σN:={t0=0,t1=TN,t2=2TN,,tN=T}\sigma_{N}:=\{t_{0}=0,t_{1}=\frac{T}{N},t_{2}=\frac{2T}{N},\ldots,t_{N}=T\} for NN\in{\mathbb{N}}. We define the discrete-time process SNS^{N} by

StiN:=Stii=0,,N,S^{N}_{t_{i}}:=S_{t_{i}}\quad i=0,\ldots,N,

i.e., SNS^{N} is the restriction of SS to σN\sigma_{N}. Hence, investing in SNS^{N} amounts to investing in SS with the restriction that the portfolio may only be adapted at time instants tiσNt_{i}\in\sigma_{N}. From now on we refer to SNS^{N} as the NN-period model approximation of (2.1). SNS^{N} evolves in each period according to

Sti+1N=StiN(1+ZiN)=S0j=0i(1+ZjN)=:S0(ZN)ti+1,S^{N}_{t_{i+1}}=S^{N}_{t_{i}}\left(1+Z^{N}_{i}\right)=S_{0}\prod_{j=0}^{i}\left(1+Z^{N}_{j}\right)=:S_{0}\mathcal{E}\left(Z^{N}\right)_{t_{i+1}},

where ZiN=S(ti+1)S(ti)S(ti)=(L)ti+1(L)ti(L)ti=eL~ti+1L~ti1Z^{N}_{i}=\frac{S(t_{i+1})-S(t_{i})}{S(t_{i})}=\frac{\mathcal{E}\left(L\right)_{t_{i+1}}-\mathcal{E}\left(L\right)_{t_{i}}}{\mathcal{E}\left(L\right)_{t_{i}}}=e^{\widetilde{L}_{t_{i+1}}-\widetilde{L}_{t_{i}}}-1 are i.i.d. random variables. Investing in each period the fraction π\pi of current wealth in the stock SNS^{N} amounts to the wealth process

(2.12) x0(πZN)ti+1:=x0j=0i(1+πZjN).x_{0}\mathcal{E}\left(\pi Z^{N}\right)_{t_{i+1}}:=x_{0}\prod_{j=0}^{i}\left(1+\pi Z^{N}_{j}\right).

For the same economic agent and utility function UU as above, the set of admissible constant trading strategies for SNS^{N} is given by

𝒜N:={π|1+πZ0N0}={π|1+πZ0N>0}=[0,1].\displaystyle\mathcal{A}^{N}:=\{\pi\in\mathbb{R}|1+\pi Z_{0}^{N}\geq 0\}=\{\pi\in\mathbb{R}|1+\pi Z_{0}^{N}>0\}=[0,1].

𝒜N\mathcal{A}^{N} is given by [0,1][0,1] and does not depend on pp and NN since we assume that LL satisfies c0c\neq 0 or allows both positive and negative jumps. We also find 𝒜N=[0,1]𝒜\mathcal{A}^{N}=[0,1]\subseteq\mathcal{A}.

The set of non-constant admissible trading strategies (πi(ω))i(\pi_{i}(\omega))_{i} is given by

𝒯N:={(πi)i=0,,N1|πi is ti measurable, j=0i(1+πiZjN)>0}.\mathcal{T}^{N}:=\{(\pi_{i})_{i=0,\ldots,N-1}\rvert\,\pi_{i}\text{ is }\mathcal{F}_{t_{i}-}\text{ measurable, }\prod_{j=0}^{i}\left(1+\pi_{i}Z^{N}_{j}\right)>0\}.

Similarly to (2.5) it is the agent’s aim to find for each NN a maximizer (πN,i)i=0N1(\pi_{N,i}^{*})_{i=0}^{N-1} of

(2.13) uN(x0):=maxπ𝒯N𝔼[U(x0(πZN)T)]=maxπ𝒯N𝔼[(x0(πZN)T)1p1p].u^{N}(x_{0}):=\max_{\pi\in\mathcal{T}^{N}}\mathbb{E}\left[U(x_{0}\mathcal{E}\left(\pi Z^{N}\right)_{T})\right]=\max_{\pi\in\mathcal{T}^{N}}\mathbb{E}\left[\frac{(x_{0}\mathcal{E}\left(\pi Z^{N}\right)_{T})^{1-p}}{1-p}\right].

Maximizing (2.13) leads to a constant maximizer πN𝒜N=[0,1]\pi^{*}_{N}\in\mathcal{A}^{N}=[0,1], i.e., πN,i=πN[0,1]\pi_{N,i}^{*}=\pi_{N}^{*}\in[0,1]. Moreover, πN\pi_{N}^{*} is in fact the maximizer of the following 1-period model

(2.14) maxπ[0,1]𝔼[U(x0(1+π((L)TN1)))],\max_{\pi\in[0,1]}\mathbb{E}\left[U\left(x_{0}\left(1+\pi(\mathcal{E}\left(L\right)_{\frac{T}{N}}-1)\right)\right)\right],

and is independent of the initial wealth x0x_{0}, cf. [20]. We want to give a short sketch by backward induction why πN\pi^{*}_{N} is constant: Given current wealth XtN1=xX_{t_{N-1}}=x, the Bellman principle and ZNZ^{N} being i.i.d. show that the optimal strategy πN,N1(x)\pi^{*}_{N,N-1}(x) for the last step maximizes (2.14) with x0=xx_{0}=x. Moreover, since UU is a power utility function, πN,N1\pi^{*}_{N,N-1} is in fact independent of xx. Hence, πN,N1=πN\pi_{N,N-1}^{*}=\pi^{*}_{N}. Similar arguments as well as (Zi=n+1N)N(Z^{N}_{i=n+1})^{N} being i.i.d. – and in particular independent of the current wealth XtnX_{t_{n}} – prove πN,n=πN\pi_{N,n}^{*}=\pi^{*}_{N} for the induction step.

Analogously to gg above, we define for each NN the deterministic concave function gN:[0,1]g^{N}:[0,1]\to\mathbb{R} by

(2.15) gN(π)=NT𝔼[(1+π((L)TN1))1p11p]for p1,gN(π)=NT𝔼[log(1+π((L)TN1))]for p=1.\displaystyle\begin{split}g^{N}(\pi)&=\frac{N}{T}\mathbb{E}\left[\frac{\left(1+\pi(\mathcal{E}\left(L\right)_{\frac{T}{N}}-1)\right)^{1-p}-1}{1-p}\right]\quad\text{for }p\neq 1,\\ g^{N}(\pi)&=\frac{N}{T}\mathbb{E}\left[\log\left(1+\pi(\mathcal{E}\left(L\right)_{\frac{T}{N}}-1)\right)\right]\quad\text{for }p=1.\end{split}

Clearly, πN\pi^{*}_{N} is a maximizer of gNg^{N}. The reason why we have slightly modified gNg^{N} in comparison to (2.14) will come apparent in the proof of Lemma 4.6 below.

The case c=0c=0 and either F()=0F(\mathbb{R}_{-})=0 or F(+)=0F(\mathbb{R}_{+})=0:

We now shortly discuss the case of LL being a pure jump process with triplet (b,0,F)(b,0,F) and either F()=0F(\mathbb{R}_{-})=0 or F(+)=0F(\mathbb{R}_{+})=0. In order to exclude arbitrage, we have to assume that bb is negative (positive) if only positive (negative) jumps are allowed.

The set of admissible constant trading strategies for the NN-period model is then given by

𝒜0N\displaystyle\mathcal{A}^{N}_{0} :={π|1+πZ0N=1+π((L)1N1)0}if p(0,1),\displaystyle:=\{\pi\in\mathbb{R}|1+\pi Z_{0}^{N}=1+\pi(\mathcal{E}\left(L\right)_{\frac{1}{N}}-1)\geq 0\}\quad\text{if }p\in(0,1),
𝒜0,N\displaystyle\mathcal{A}^{N}_{0,*} :={π|1+πZ0N=1+π((L)1N1)>0}if p1.\displaystyle:=\{\pi\in\mathbb{R}|1+\pi Z_{0}^{N}=1+\pi(\mathcal{E}\left(L\right)_{\frac{1}{N}}-1)>0\}\quad\text{if }p\geq 1.

By S>0S>0 we have [0,1]𝒜N[0,1]\subseteq\mathcal{A}^{N} and 𝒜N\mathcal{A}^{N} is bounded for all NN. We note that 1+π((L)t1)0(>0)1+\pi(\mathcal{E}\left(L\right)_{t}-1)\geq 0\;(>0) implies 1+π((L)s1)0(>0)1+\pi(\mathcal{E}\left(L\right)_{s}-1)\geq 0\;(>0) for all 0st0\leq s\leq t, which further leads to 𝒜N𝒜N+1\mathcal{A}^{N}\subseteq\mathcal{A}^{N+1}. Letting t0t\to 0 in 1+π((L)t1)0(>0)1+\pi(\mathcal{E}\left(L\right)_{t}-1)\geq 0(>0), [9, II.8a] implies 𝒜N𝒜\mathcal{A}^{N}\subseteq\mathcal{A}. We see that 𝒜N\mathcal{A}^{N} depends on NN and [0,1][0,1] is generally strictly contained in 𝒜N\mathcal{A}^{N}.

As already mentioned, we make in the sequel the assumption that c0c\neq 0 or both F()>0F(\mathbb{R}_{-})>0 and F(+)>0F(\mathbb{R}_{+})>0, i.e., 𝒜N=[0,1]\mathcal{A}^{N}=[0,1] which greatly simplifies further notation. The exogenous portfolio constraint 𝒞=[0,1]\mathscr{C}=[0,1], which implies the convergence limNπN=π𝒞\lim_{N\to\infty}\pi^{*}_{N}=\pi^{*}_{\mathscr{C}} in the case of c0c\neq 0 or both F()>0F(\mathbb{R}_{-})>0 and F(+)>0F(\mathbb{R}_{+})>0 (see Theorem 4.1 below), translates to 𝒞=N1𝒜N¯\mathscr{C}=\bigcup_{N\geq 1}\overline{\mathcal{A}^{N}} in the case of c=0c=0 and either F()=0F(\mathbb{R}_{-})=0 or F(+)=0F(\mathbb{R}_{+})=0.

3. Analytic Properties of gg and gNg^{N}

This section summarizes some results on the continuity and differentiability of the functions gg and gNg^{N}, defined in (2.10) and (2.15), respectively.

Proposition 3.1.

Under Assumption 2.1, gg is continuous on 𝒜0,\mathcal{A}_{0,*}. Moreover, gg and gNg^{N} are finite and differentiable on 𝒜0,\mathcal{A}_{0,*}, respectively [0,1][0,1] with derivatives

(3.16) g(π)\displaystyle g^{\prime}(\pi) =bpπc+(1,)(x(1+πx)ph(x))𝑑F(x),\displaystyle=b-p\pi c+\int_{(-1,\infty)}\left(\frac{x}{\left(1+\pi x\right)^{p}}-h(x)\right)\;dF(x),
(3.17) (gN)(π)\displaystyle\left(g^{N}\right)^{\prime}(\pi) =NT𝔼[exp(L~TN)1(1+π(exp(L~TN)1))p],\displaystyle=\frac{N}{T}\mathbb{E}\left[\frac{\exp(\widetilde{L}_{\frac{T}{N}})-1}{\left(1+\pi(\exp(\widetilde{L}_{\frac{T}{N}})-1)\right)^{p}}\right],

that are finite on (𝒜0,)\left(\mathcal{A}_{0,*}\right)^{\circ}, respectively (0,1)(0,1), where (𝒜0,)\left(\mathcal{A}_{0,*}\right)^{\circ} denotes the interior of 𝒜0,\mathcal{A}_{0,*}

For the proof of Proposition 3.1 we need the following elementary result on exchanging differentiation and integration for concave functions, which is a special case of [17, Lemma 5.14].

Lemma 3.2.

Let η\eta be a Borel-measure on \mathbb{R}, y0y_{0}\in\mathbb{R} and let f:×f:\mathbb{R}\times\mathbb{R}\to\mathbb{R} satisfy

  1. (i)

    xf(x,y)x\mapsto f(x,y) is measurable and |f(x,y)|𝑑η(x)<\int_{\mathbb{R}}|f(x,y)|d\eta(x)<\infty for yy in a neighbourhood of y0y_{0}, and

  2. (ii)

    x:\forall x\in\mathbb{R}: yf(x,y)y\mapsto f(x,y) is concave and differentiable.

Then k(y):=f(x,y)𝑑η(x)k(y):=\int_{\mathbb{R}}f(x,y)d\eta(x) is concave and differentiable with derivative

k(y0)=f(x,y0)𝑑η(x),k^{\prime}(y_{0})=\int_{\mathbb{R}}f^{\prime}(x,y_{0})d\eta(x),

where f(x,y0)f^{\prime}(x,y_{0}) denotes the partial derivative of ff w.r.t. the second argument yy at (x,y0)(x,y_{0}).

Proof.

Cf. [17, Lemma 5.14]. ∎

Proof of Proposition 3.1.

Since u(x0)<u(x_{0})<\infty we find in particular that gNg^{N} is finite on [0,1][0,1]. By [16, Corollary 3.7], u(x0)<u(x_{0})<\infty is equivalent to

(3.18) x>1(1+x)1p1p𝑑F(x)<for p>0,\int_{x>1}\frac{(1+x)^{1-p}}{1-p}\;dF(x)<\infty\quad\text{for }p>0,

and [16, Lemma 5.3] gives that gg is finite and continuous on 𝒜0,\mathcal{A}_{0,*}.

Lemma 3.2 implies that gg and gNg^{N} are (left- and right-) differentiable on 𝒜0,\mathcal{A}_{0,*} and [0,1][0,1], respectively, and their derivatives are given by (3.16) and (3.17), respectively. The concavity of gg and gNg^{N} implies |g|,|(gN)|<|g^{\prime}|,|(g^{N})^{\prime}|<\infty on (𝒜0,)\left(\mathcal{A}_{0,*}\right)^{\circ} and (0,1)(0,1), respectively. ∎

4. Convergence Results for the Exponential Lévy model

Our main convergence result is stated in Theorem 4.1 below, which also implies two important corollaries. Proposition 4.4 gives qualitative properties of the optimal strategies πN\pi_{N}^{*} and π\pi^{*} related to their sign and the risk aversion of UU. To improve readability, we move the proofs of Theorem 4.1 and Proposition 4.4 to the end of this section. We comment on the extension of Theorem 4.1 to higher dimensions at the end of its proof.

Theorem 4.1.

Under Assumption 2.1, gNg^{N} converges uniformly on [0,1][0,1] to gg. In particular, limNπN=π𝒞\lim_{N\to\infty}\pi^{*}_{N}=\pi^{*}_{\mathscr{C}} and limNgN(πN)=g(π𝒞)\lim_{N\to\infty}g^{N}(\pi^{*}_{N})=g(\pi^{*}_{\mathscr{C}}). Thus, if π[0,1]\pi^{*}\in[0,1] then limNπN=π\lim_{N\to\infty}\pi^{*}_{N}=\pi^{*}, and limNgN(πN)=g(π)\lim_{N\to\infty}g^{N}(\pi^{*}_{N})=g(\pi^{*}).

Corollary 4.2.

Under Assumption 2.1, the NN-period value functions converge to the value function of the constrained continuous-time model, i.e.,

limN𝔼[U(x0(πNZN)T)]=𝔼[U(x0(π𝒞L)T)].\lim_{N\to\infty}\mathbb{E}\left[U\left(x_{0}\mathcal{E}\left(\pi^{*}_{N}Z^{N}\right)_{T}\right)\right]=\mathbb{E}\left[U(x_{0}\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right)_{T})\right].

Thus, if π[0,1]\pi^{*}\in[0,1] then

limN𝔼[U(x0(πNZN)T)]=𝔼[U(x0(πL)T)].\lim_{N\to\infty}\mathbb{E}\left[U\left(x_{0}\mathcal{E}\left(\pi^{*}_{N}Z^{N}\right)_{T}\right)\right]=\mathbb{E}\left[U(x_{0}\mathcal{E}\left(\pi^{*}L\right)_{T})\right].
Proof.

We give the proof for 0<p0<p, p1p\neq 1, since the case p=1p=1 is shown similarly. W.l.o.g. we set x0=1x_{0}=1. By (2.12) and ZjZ_{j} being i.i.d. we find

𝔼[U((πNZN)T)]=𝔼[(1+πNZ0N)1p]N1p=(1+(1p)TgN(πN)N)N1p.\mathbb{E}\left[U\left(\mathcal{E}\left(\pi^{*}_{N}Z^{N}\right)_{T}\right)\right]=\frac{\mathbb{E}\left[(1+\pi_{N}^{*}Z^{N}_{0})^{1-p}\right]^{N}}{1-p}=\frac{\left(1+\frac{(1-p)Tg^{N}(\pi_{N}^{*})}{N}\right)^{N}}{1-p}.

Since (1+aN)Nexp(a)(1+\frac{a}{N})^{N}\to\exp(a) converges uniformly on compacts and gN(πN)g(π𝒞)g^{N}(\pi_{N}^{*})\to g(\pi^{*}_{\mathscr{C}}) as NN\to\infty, equation (2.11) shows

𝔼[U((πNZN)T)]exp((1p)Tg(π𝒞))1p=𝔼[U((π𝒞L)T)].\mathbb{E}\left[U\left(\mathcal{E}\left(\pi^{*}_{N}Z^{N}\right)_{T}\right)\right]\to\frac{\exp\left((1-p)Tg(\pi^{*}_{\mathscr{C}})\right)}{1-p}=\mathbb{E}\left[U(\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right)_{T})\right].

The proof of the next corollary of Theorem 4.1 requires some results on the convergence of Euler approximations of (2.4), which can be found in the Appendix.

Corollary 4.3.

Let Assumption 2.1 be satisfied and let LL be square integrable. Then the optimal terminal payoffs of the NN-period models converge in L2(Ω)L^{2}(\Omega) to the optimal terminal payoff of the constrained continuous-time model, i.e.,

limN𝔼[(x0(πNZN)Tx0(π𝒞L)T)2]=0.\lim_{N\to\infty}\mathbb{E}\left[\left(x_{0}\mathcal{E}\left(\pi^{*}_{N}Z^{N}\right)_{T}-x_{0}\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right)_{T}\right)^{2}\right]=0.

Thus, if π[0,1]\pi^{*}\in[0,1] then

limN𝔼[(x0(πNZN)Tx0(πL)T)2]=0.\lim_{N\to\infty}\mathbb{E}\left[\left(x_{0}\mathcal{E}\left(\pi^{*}_{N}Z^{N}\right)_{T}-x_{0}\mathcal{E}\left(\pi^{*}L\right)_{T}\right)^{2}\right]=0.
Proof.

W.l.o.g. we set x0=1x_{0}=1. By Proposition A.4 in the Appendix ((πNZN)T(πNL)T)(\mathcal{E}\left(\pi^{*}_{N}Z^{N}\right)_{T}-\mathcal{E}\left(\pi^{*}_{N}L\right)_{T}) converges in L2(Ω)L^{2}(\Omega) to 0, where (πNL)\mathcal{E}\left(\pi^{*}_{N}L\right) denotes the Euler approximation of (πNL)\mathcal{E}\left(\pi^{*}_{N}L\right), cf. (A.23) in the Appendix.

πN[0,1]\pi^{*}_{N}\in[0,1] implies (πNL)>0\mathcal{E}\left(\pi^{*}_{N}L\right)>0. In particular, Theorem A.3 in the Appendix implies that the Euler approximations (πNL)T\mathcal{E}\left(\pi^{*}_{N}L\right)_{T} converge to (π𝒞L)T\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right)_{T} as NN\to\infty. In total, (πNZN)T\mathcal{E}\left(\pi^{*}_{N}Z^{N}\right)_{T} converges in L2(Ω)L^{2}(\Omega) to (π𝒞L)T\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right)_{T}. ∎

Proposition 4.4.

Let Assumption 2.1 be satisfied and let LL be integrable, i.e., we can choose h(x)=xh(x)=x.

  1. (i)

    Letting b=b(x)b=b(x) denote the drift w.r.t. h(x)=xh(x)=x, we find

    (4.19) b0π,πN0andb0π0,πN=0.\displaystyle b\geq 0\Longrightarrow\pi^{*},\pi^{*}_{N}\geq 0\quad\text{and}\quad b\leq 0\Longrightarrow\pi^{*}\leq 0,\pi^{*}_{N}=0.

    If π1\pi^{*}\geq 1, then πN=1\pi^{*}_{N}=1.

  2. (ii)

    Assume that the maximum in (2.5) is finite for all U=x1p1pU=\frac{x^{1-p}}{1-p} with p(pa,pb)p\in(p_{a},p_{b}). We denote by πp\pi_{p}^{*}, πp,N\pi_{p,N}^{*} the corresponding optimal strategy for UU in the continuous-time and discrete-time model, respectively. Then the mappings pπpp\mapsto\pi_{p}^{*}, pπp,Np\mapsto\pi_{p,N}^{*} are montone decreasing (increasing)(\text{increasing}) on (pa,pb)(p_{a},p_{b}), if b>0b>0 (b<0)(b<0), where b=b(x)b=b(x) is the drift w.r.t. h(x)=xh(x)=x.

Remark 4.5.
  1. (i)

    (4.19) can be shown in a more general framework using sub- and supermartingale arguments. However, we want to establish (i) by taking advantage of the special structure of the exponential Lévy model and prove it by montonicity properities of the functions gg and gNg^{N}, see the proof below.

  2. (ii)

    By U′′(x)U(x)=px-\frac{U^{\prime\prime}(x)}{U^{\prime}(x)}=\frac{p}{x} the constant pp determines the index of relative risk aversion of UU. Proposition 4.4.(ii) formalizes the intuitive idea that higher risk aversion leads to a lower fraction of wealth invested in the stock SS.

Proof of Theorem 4.1:

The proof of Theorem 4.1 requires the pointwise convergence of gNg^{N} to gg that is established by Lemma 4.6.

Lemma 4.6.

Let Assumption 2.1 be satisfied and let π[0,1]\pi\in[0,1]. Then

limNgN(π)=g(π).\lim_{N\to\infty}g^{N}(\pi)=g(\pi).
Proof.

We only give the proof for 0<p0<p, p1p\neq 1, since the case p=1p=1 is obtained similarly. A Taylor expansion of x(1+π(ex1))1p1px\mapsto\frac{\left(1+\pi\left(e^{x}-1\right)\right)^{1-p}}{1-p} around 0 shows

2πpπ2((1+π(ex1))1p11pπx)x2as x0,\frac{2}{\pi-p\pi^{2}}\left(\frac{\left(1+\pi(e^{x}-1)\right)^{1-p}-1}{1-p}-\pi x\right)\sim x^{2}\quad\text{as }x\to 0,

where we use the common notation f(x)g(x),x0limx0f(x)g(x)=1f(x)\sim g(x),\;x\to 0\Leftrightarrow\lim_{x\to 0}\frac{f(x)}{g(x)}=1. In particular, we find

k(x):=2πpπ2((1+π(ex1))1p11pπh(x))x2as x0.k(x):=\frac{2}{\pi-p\pi^{2}}\left(\frac{\left(1+\pi(e^{x}-1)\right)^{1-p}-1}{1-p}-\pi h(x)\right)\sim x^{2}\quad\text{as }x\to 0.

Since kk is locally bounded we can apply [3, Theorem 1.1 (ii)], which implies

(4.20) limNNT𝔼[k(L~TN)]=c~+k(x)𝑑F~(x).\lim_{N\to\infty}\frac{N}{T}\mathbb{E}\left[k(\widetilde{L}_{\frac{T}{N}})\right]=\widetilde{c}+\int_{\mathbb{R}}k(x)\;d\widetilde{F}(x).

[8, Lemma 4.2] implies limNNTπh(L~TN)=πb~(h)\lim_{N\to\infty}\frac{N}{T}\pi h(\widetilde{L}_{\frac{T}{N}})=\pi\widetilde{b}(h). Shifting this limit on the right side of (4.20) and using equations (2.6)-(2.8) we finally get

limNgN(π)=πbpπ2c2+(1,)((1+πx)1p11pπh(x))𝑑F(x).\lim_{N\to\infty}g^{N}(\pi)=\pi b-\frac{p\pi^{2}c}{2}+\int_{(-1,\infty)}\left(\frac{\left(1+\pi x\right)^{1-p}-1}{1-p}-\pi h(x)\right)\;dF(x).

Proof of Theorem 4.1.

By Lemma 4.6 we know that gNg^{N} converges pointwise to gg on [0,1][0,1]. Lemma A.2 in the Appendix shows that the concavity of gNg^{N} and gg imply that gNg^{N} converges in fact uniformly to gg on [0,1][0,1] as NN\to\infty. Thus, the NN-period maximizers πN\pi_{N}^{*} satisfy limNπN=argmax[0,1]g=π𝒞\lim_{N\to\infty}\pi_{N}^{*}=\arg\max_{[0,1]}g=\pi^{*}_{\mathscr{C}} and limNgN(πN)=g(π𝒞)\lim_{N\to\infty}g^{N}(\pi^{*}_{N})=g(\pi^{*}_{\mathscr{C}}). ∎

Extensions to Higher Dimensions

If the Lévy process L=(L1,,Ln)L=(L_{1},\ldots,L_{n}) and the stock S=(L)=((L1),,(Ln))S=\mathcal{E}\left(L\right)=(\mathcal{E}\left(L_{1}\right),\ldots,\mathcal{E}\left(L_{n}\right)) are n\mathbb{R}^{n} valued, the only difficulty in extending our results to higher dimensions is the no-arbitrage condition of no unbounded increasing profit, cf. [12]. As we have seen in Assumption 2.1.(ii) this condition is almost trivially satisfied in one dimension, but it gets more involved in higher dimensions. Lemma 4.6 also proves the pointwise convergence of gNg^{N} to gg on [0,1]n[0,1]^{n}, where gg and gNg^{N} are defined verbatim in higher dimensions. This finally proves by concavity the uniform convergence of gNg^{N} to gg on [0,1]n[0,1]^{n}. Thus, under appropriate no-arbitrage assumptions Theorem 4.1 also holds in higher dimensions.

Proof of Proposition 4.4:

Proof.

ad (i): Using (3.16) with h(x)=xh(x)=x we get g(0)=b=b(x)g^{\prime}(0)=b=b(x), which immediately gives (4.19). Moreover, (3.17) implies (gN)(0)=NT1(ebN1T1)(g^{N})^{\prime}(0)=NT^{-1}(e^{bN^{-1}T}-1). Hence, b0b\leq 0 implies πN=0\pi^{*}_{N}=0 and b0b\geq 0 gives πN0\pi^{*}_{N}\geq 0. If π1\pi^{*}\geq 1 we find in particular 0g(1)<0\leq g^{\prime}(1)<\infty. Using (2.6)-(2.8) a straight forward calculation shows

(gN)(1)=exp(pb~+p2c2+(epx1+ph(x))𝑑F~)(eg(1)1)0,(g^{N})^{\prime}(1)=\exp\left(-p\widetilde{b}+\frac{p^{2}c}{2}+\int_{\mathbb{R}}\big{(}e^{-px}-1+ph(x)\big{)}\,d\widetilde{F}\right)\cdot(e^{g^{\prime}(1)}-1)\geq 0,

thus πN=1\pi_{N}^{*}=1.

ad (ii): Let b0b\neq 0 since otherwise π=πN=0\pi^{*}=\pi^{*}_{N}=0. We first consider the continuous-time case: letting p1<p2p_{1}<p_{2}, p1,p2(pa,pb)p_{1},p_{2}\in(p_{a},p_{b}), (i) shows that we may assume |πp1|,|πp2|>0|\pi_{p_{1}}^{*}|,|\pi_{p_{2}}^{*}|>0 and that πp1\pi_{p_{1}}^{*}, πp2\pi_{p_{2}}^{*} have the same sign. Hence, in order to prove (ii) we have to establish |πp1||πp2||\pi_{p_{1}}^{*}|\geq|\pi_{p_{2}}^{*}|.

We first consider the case when πp1\pi_{p_{1}}^{*} and πp2\pi_{p_{2}}^{*} are in the interior (𝒜0,)(\mathcal{A}_{0,*})^{\circ} of 𝒜0,\mathcal{A}_{0,*}. As πpi0\pi_{p_{i}}^{*}\neq 0, πpi\pi_{p_{i}}^{*} satisfy in particular the equation

G(πpi,pi):=πpibpi(πpi)2c+(1,)(πpix(1+πpix)piπpix)𝑑F(x)=0G(\pi_{p_{i}}^{*},p_{i}):=\pi_{p_{i}}^{*}b-p_{i}(\pi_{p_{i}}^{*})^{2}c+\int_{(-1,\infty)}\left(\frac{\pi_{p_{i}}^{*}x}{(1+\pi_{p_{i}}^{*}x)^{p_{i}}}-\pi_{p_{i}}^{*}x\right)\,dF(x)=0

with i=1,2i=1,2, since G(π,p)=πg(π,p)G(\pi,p)=\pi g^{\prime}(\pi,p). Here g(,p)g^{\prime}(\cdot,p) denotes the partial derivative w.r.t. π\pi of the function gg corresponding to pp.

The integral I(π,p):=(1,)x(1+πx)pxdF(x)I(\pi,p):=\int_{(-1,\infty)}x(1+\pi x)^{-p}-x\,dF(x) is partially differentiable w.r.t. π\pi and pp: indeed, the integrand (π,p)x(1+πx)px(\pi,p)\mapsto x(1+\pi x)^{-p}-x is convex (concave) in π\pi (with pp fixed) and in pp (with π\pi fixed) if x0x\geq 0 (x0x\leq 0). Hence, we can apply Lemma 3.2 by splitting the domain of integration into (1,0](-1,0] and [0,)[0,\infty), which yields that II is partially differentiable w.r.t. π(𝒜0,)\{0}\pi\in(\mathcal{A}_{0,*})^{\circ}\backslash\{0\} and p(pa,pb)p\in(p_{a},p_{b}). Moreover, differentiation and integration can be interchanged. In particular, the partial derivatives of GG are finite for π(𝒜0,)\{0}\pi\in(\mathcal{A}_{0,*})^{\circ}\backslash\{0\} and p(pa,pb)p\in(p_{a},p_{b}). Lemma A.6 in the Appendix shows that the partial derivatives of GG are continuous for any p(pa,pb)p\in(p_{a},p_{b}) and π(𝒜0,)\{0}\pi\in(\mathcal{A}_{0,*})^{\circ}\backslash\{0\}, which further implies that GG is continuously differentiable.

Hence, any pair of optimal strategies (πp,p)(\pi_{p}^{*},p) with G(πp,p)=0G(\pi_{p}^{*},p)=0 satisfies

πpGπ(πp,p)\displaystyle\pi_{p}^{*}\frac{\partial G}{\partial\pi}(\pi_{p}^{*},p) =G(πp,p)p(πp)2c(1,)π2x2p(1+πpx)p1𝑑F(x),\displaystyle=G(\pi_{p}^{*},p)-p(\pi_{p}^{*})^{2}c-\int_{(-1,\infty)}\pi^{2}x^{2}p(1+\pi_{p}^{*}x)^{-p-1}\;dF(x),
Gπ(πp,p)\displaystyle\Longleftrightarrow\;\frac{\partial G}{\partial\pi}(\pi_{p}^{*},p) =pπp(c+(1,)x2(1+πpx)p1𝑑F(x)),\displaystyle=-p\pi_{p}^{*}\left(c+\int_{(-1,\infty)}x^{2}(1+\pi_{p}^{*}x)^{-p-1}\;dF(x)\right),

where we have used |πp|>0|\pi_{p}^{*}|>0. In particular, sgn(Gπ(πp,p))=sgn(πp)0\text{sgn}\left(\frac{\partial G}{\partial\pi}(\pi_{p}^{*},p)\right)=-\text{sgn}(\pi_{p}^{*})\neq 0.

By the implicit function theorem, there exists a differentiable function ϕ\phi defined on a neighbourhood NN of p1p_{1} such that pN:G(ϕ(p),p)=0\forall p\in N:G(\phi(p),p)=0, ϕ(p1)=πp1\phi(p_{1})=\pi^{*}_{p_{1}} and

(4.21) ϕ(p)=(Gπ(ϕ(p),p))1Gp(ϕ(p),p).\phi^{\prime}(p)=-\left(\frac{\partial G}{\partial\pi}(\phi(p),p)\right)^{-1}\frac{\partial G}{\partial p}(\phi(p),p).

By continuity we may also assume that ϕ(p)0\phi(p)\neq 0 for all pNp\in N. Differentiating GG with respect to pp implies for all π𝒜0,\{0}\pi\in\mathcal{A}_{0,*}\backslash\{0\}

Gp(π,p)=π2c(1,)πxlog(1+πx)(1+πx)p𝑑F(x)<0,\frac{\partial G}{\partial p}(\pi,p)=-\pi^{2}c-\int_{(-1,\infty)}\frac{\pi x\log(1+\pi x)}{\left(1+\pi x\right)^{p}}\;dF(x)<0,

where the right inequality follows from (1+πx)>0(1+\pi x)>0 FF-a.s. and πxlog(1+πx)0\pi x\log(1+\pi x)\geq 0. Hence, (4.21) yields sgn(ϕ(p))=sgn(ϕ(p))\text{sgn}(\phi^{\prime}(p))=-\text{sgn}(\phi(p)). This also implies that we may assume w.l.o.g. that ϕ\phi is defined on an interval containing [p1,p2][p_{1},p_{2}]: let pp1p\geq p_{1} be at the right boundary of NN and πp:=limNpnpϕ(pn)\pi_{p}:=\lim_{N\ni p_{n}\to p}\phi(p_{n}), then |ϕ(p1)|=|πp1||ϕ(p)|=|πp||\phi(p_{1})|=|\pi_{p_{1}}^{*}|\geq|\phi(p)|=|\pi_{p}| and

G(πp,p):=limNpnpG(ϕ(pn),pn)=limNpnp0=0.G(\pi_{p},p):=\lim_{N\ni p_{n}\to p}G(\phi(p_{n}),p_{n})=\lim_{N\ni p_{n}\to p}0=0.

The continuity of G(π,p)=πg(π,p)G(\pi,p)=\pi g^{\prime}(\pi,p) as well as the uniqueness of a solution πp0\pi^{*}_{p}\neq 0 of the utility maximization problem related to pp implies g(πp,p)=0g^{\prime}(\pi_{p},p)=0, πp=πp0\pi^{*}_{p}=\pi_{p}\neq 0 and G(πp,p)=0G(\pi_{p}^{*},p)=0. By |Gπ(πp,p)|0|\frac{\partial G}{\partial\pi}(\pi_{p}^{*},p)|\neq 0 the implicit function theorem can also be applied at pp, which implies that ϕ\phi can be defined on an interval containing [p1,p2][p_{1},p_{2}]. Finally, sgn(ϕ(p))=sgn(ϕ(p))\text{sgn}(\phi^{\prime}(p))=-\text{sgn}(\phi(p)) shows |ϕ(p1)|=|πp1||πp2|=|ϕ(p2)||\phi(p_{1})|=|\pi_{p_{1}}^{*}|\geq|\pi_{p_{2}}^{*}|=|\phi(p_{2})|, as desired.

If πp2\pi_{p_{2}}^{*} is at the boundary of 𝒜0,\mathcal{A}_{0,*}, we only have to treat the case when πp2\pi_{p_{2}}^{*} is at the right boundary while πp1<πp2\pi_{p_{1}}^{*}<\pi_{p_{2}}^{*}. We then find

G(π,p2)0π𝒜0,andG(πp1,p1)0.G(\pi,p_{2})\geq 0\quad\forall\pi\in\mathcal{A}_{0,*}\quad\text{and}\quad G(\pi_{p_{1}}^{*},p_{1})\leq 0.

Since Gp(π,p)<0\frac{\partial G}{\partial p}(\pi,p)<0 we find G(πp1,p2)<0G(\pi_{p_{1}}^{*},p_{2})<0, which contradicts the above assumption. Thus |πp1||πp2||\pi_{p_{1}}^{*}|\geq|\pi_{p_{2}}^{*}|. The case of πp1\pi_{p_{1}}^{*} being at the boundary of 𝒜0,\mathcal{A}_{0,*} is treated analogously.

In the discrete-time case, similar arguments as above with G(π,p):=π𝔼[(1+π(exp(L~1N)1))p(exp(L~1N)1)]G(\pi,p):=\pi\mathbb{E}[(1+\pi(\exp(\widetilde{L}_{\frac{1}{N}})-1))^{-p}(\exp(\widetilde{L}_{\frac{1}{N}})-1)] show that (ii) also holds for πN\pi^{*}_{N}. ∎

Appendix A

The Appendix contains results on the integrability of the stochastic exponential (Lemma A.1), on the pointwise and uniform convergence of concave functions (Lemma A.2) and on the convergence of the optimal NN-period wealth processes (πNZN)\mathcal{E}\left(\pi_{N}^{*}Z^{N}\right) to the Euler approximation of (π𝒞L)\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right), i.e., the optimal wealth process of the constrained continuous-time model (Proposition A.4). In the proof of Corollary 4.3 we also need the convergence of the Euler approximations to (π𝒞L)\mathcal{E}\left(\pi^{*}_{\mathscr{C}}L\right). More explicitly, for our purpose we need that the convergence holds in fact uniformly in π𝒞\pi^{*}_{\mathscr{C}} (Theorem A.3). Since this result only requires a slight modification of a result of Kohatsu and Protter ([13]), we omit a detailed proof of Theorem A.3 and just remark on the adaption of the proof in [13]. At the end of the Appendix we also present a technical continuity result (Lemma A.6) that we have omitted in the proof of Proposition 4.4.

Lemma A.1.

Let LL be a Lévy process satisfying 𝔼[(Lt)p]<\mathbb{E}\left[(L_{t})^{p}\right]<\infty with p>0p>0 and (L)t>0\mathcal{E}\left(L\right)_{t}>0. Then there exists a constant c(p)c(p)\in\mathbb{R} such that

(A.22) 𝔼[(L)tp]=ec(p)t.\mathbb{E}\left[\mathcal{E}\left(L\right)_{t}^{p}\right]=e^{c(p)t}.
Proof.

Using (L)t=exp(L~t)\mathcal{E}\left(L\right)_{t}=\exp(\widetilde{L}_{t}) this is a simple consequence of [21, Theorem 25.3]. ∎

Lemma A.2.

Let c,cn:[0,1]c,c_{n}:[0,1]\to\mathbb{R} be concave functions and let cc be continuous on [0,1][0,1]. Then the pointwise convergence of cnc_{n} to cc on [0,1][0,1] implies the uniform convergence of cnc_{n} to cc on [0,1][0,1].

Proof.

Let σN:={ti=iN|i=0,,N}\sigma_{N}:=\{t_{i}=\frac{i}{N}|i=0,\ldots,N\} and let ϵ>0\epsilon>0 be given. Define for any function f:[0,1]f:[0,1]\to\mathbb{R}

η(f,N):=maxi=0,,N|f(ti)f(ti+1)|.\eta(f,N):=\max_{i=0,\ldots,N}\left|f\left(t_{i}\right)-f\left(t_{i+1}\right)\right|.

Define [x]N[x]_{N} for all x[0,1]x\in[0,1] by [x]N:=ti[x]_{N}:=t_{i} if tix<ti+1t_{i}\leq x<t_{i+1}. Any concave function c~\tilde{c} satisfies

|c~(x)c~([x]N)|2η(c~,N) for N2.|\tilde{c}(x)-\tilde{c}([x]_{N})|\leq 2\eta(\tilde{c},N)\quad\text{ for }N\geq 2.

By assumption there exists M(N)M(N)\in{\mathbb{N}} such that maxi=0,,N|cn(ti)c(ti)|<ϵ\max_{i=0,\ldots,N}|c_{n}(t_{i})-c(t_{i})|<\epsilon for all nM(N)n\geq M(N). Hence, we find for all nM(N)n\geq M(N) and all x[0,1]x\in[0,1]

|c(x)cn(x)|2η(c,N)+2η(cn,N)+|cn([x]N)c([x]N)|.|c(x)-c_{n}(x)|\leq 2\eta(c,N)+2\eta(c_{n},N)+\left|c_{n}([x]_{N})-c([x]_{N})\right|.

The uniform continuity of cc implies that we can choose NN large enough such that η(c,N)<ϵ\eta(c,N)<\epsilon. Together with the last inequality this further implies maxx[0,1]|c(x)cn(x)|<9ϵ\max_{x\in[0,1]}|c(x)-c_{n}(x)|<9\epsilon for all nM(N)n\geq M(N). ∎

Theorem A.3.

Let LL be square integrable and XNX^{N} be positive and solution of

XtN=X0+0tπNXsN𝑑Ls,X_{t}^{N}=X_{0}+\int_{0}^{t}\pi_{N}X_{s-}^{N}\;dL_{s},

where πN\pi_{N} is a fixed sequence of real numbers converging to π\pi, i.e., XN=(πNL)X^{N}=\mathcal{E}\left(\pi_{N}L\right). Let σn:={t0=0,t1=Tn,,tn=T}\sigma_{n}:=\{t_{0}=0,t_{1}=\frac{T}{n},\ldots,t_{n}=T\} and define by XtkN,n=i=0k1(1+πNΔinL)=:(πNΔnL)tkX^{N,n}_{t_{k}}=\prod_{i=0}^{k-1}(1+\pi_{N}\Delta_{i}^{n}L)=:\mathcal{E}\left(\pi_{N}\Delta^{n}L\right)_{t_{k}} the Euler approximation of XNX^{N}, where ΔinL:=L(i+1)TnLiTn\Delta_{i}^{n}L:=L_{\frac{(i+1)T}{n}}-L_{\frac{iT}{n}}. Then

limn𝔼[suptT|XtN,nXtN|2]=0,\lim_{n\to\infty}\mathbb{E}\left[\sup_{t\leq T}|X^{N,n}_{t}-X^{N}_{t}|^{2}\right]=0,

and the limit holds uniformly in NN. In particular, under the above assumptions limN𝔼[((πNΔNL)T(πL)T)2]\lim_{N\to\infty}\mathbb{E}\left[(\mathcal{E}\left(\pi_{N}\Delta^{N}L\right)_{T}-\mathcal{E}\left(\pi L\right)_{T})^{2}\right].

Proof.

This result follows in fact from a slight modification of [13, Theorem 2.5]. Kohatsu and Protter prove the result for SDEs of the type XtN=X0+0tFN(X)s𝑑YsX_{t}^{N}=X_{0}+\int_{0}^{t}F^{N}(X)_{s-}\,dY_{s}, where FF is bounded and YY is a special semimartingale. Although in our case FN(X)s=πNXsF^{N}(X)_{s}=\pi_{N}X_{s} is unbounded, a moment’s reflection reveals that the independence of the increments of Y=LY=L implies that one only has to assume 𝔼[(XtkN,n)2]<C<\mathbb{E}\left[(X^{N,n}_{t_{k}})^{2}\right]<C<\infty with CC independent of kk, which is satisfied since LL is square integrable. As πNπ\pi_{N}\to\pi it is also not hard to see that the above convergence holds uniformly in NN. ∎

To simplify the reading of Proposition A.4, we recall the following notation

(πNZn)tk\displaystyle\mathcal{E}\left(\pi_{N}Z^{n}\right)_{t_{k}} =j=0k1(1+πNZjn) where Zjn=exp(L~tj+1L~tj)1,\displaystyle=\prod_{j=0}^{k-1}\left(1+\pi_{N}Z^{n}_{j}\right)\quad\text{ where }\quad Z^{n}_{j}=\exp(\widetilde{L}_{t_{j+1}}-\widetilde{L}_{t_{j}})-1,
(A.23) (πNΔnL)tk\displaystyle\mathcal{E}\left(\pi_{N}\Delta^{n}L\right)_{t_{k}} =j=0k1(1+πNΔjnL) where ΔjnL=Ltj+1Ltj,\displaystyle=\prod_{j=0}^{k-1}(1+\pi_{N}\Delta_{j}^{n}L)\quad\text{ where }\quad\Delta_{j}^{n}L=L_{t_{j+1}}-L_{t_{j}},

with ti=iTnt_{i}=\frac{iT}{n}.

Proposition A.4.

Let πN\pi_{N} be a sequence of real numbers converging to π\pi and let LL be square integrable. Then

limn𝔼[((πNZn)T(πNΔnL)T)2]=0,\lim_{n\to\infty}\mathbb{E}\left[\left(\mathcal{E}\left(\pi_{N}Z^{n}\right)_{T}-\mathcal{E}\left(\pi_{N}\Delta^{n}L\right)_{T}\right)^{2}\right]=0,

where the limit holds uniformly in NN. In particular, under the above assumptions limN𝔼[((πNZN)T(πNΔNL)T)2]=0\lim_{N\to\infty}\mathbb{E}\left[(\mathcal{E}\left(\pi_{N}Z^{N}\right)_{T}-\mathcal{E}\left(\pi_{N}\Delta^{N}L\right)_{T})^{2}\right]=0.

Proof.

Letting Xtkn:=(πNZn)tkX_{t_{k}}^{n}:=\mathcal{E}\left(\pi_{N}Z^{n}\right)_{t_{k}} and Ytk:=(πNΔnL)tkY_{t_{k}}:=\mathcal{E}\left(\pi_{N}\Delta^{n}L\right)_{t_{k}}, they satisfy

(A.24) Xtk+1n=Xtkn+πNXtknZknandYtk+1n=Ytkn+πNYtknΔknL.\displaystyle X_{t_{k+1}}^{n}=X_{t_{k}}^{n}+\pi_{N}X_{t_{k}}^{n}Z^{n}_{k}\quad\text{and}\quad Y_{t_{k+1}}^{n}=Y_{t_{k}}^{n}+\pi_{N}Y_{t_{k}}^{n}\Delta^{n}_{k}L.

Defining the i.i.d. random variables ϵkn:=ZknΔknL=(L)k+1n(L)kn(L)kn(Lk+1nLkn)\epsilon^{n}_{k}:=Z^{n}_{k}-\Delta^{n}_{k}L=\frac{\mathcal{E}\left(L\right)_{\frac{k+1}{n}}-\mathcal{E}\left(L\right)_{\frac{k}{n}}}{\mathcal{E}\left(L\right)_{\frac{k}{n}}}-\left(L_{\frac{k+1}{n}}-L_{\frac{k}{n}}\right), we deduce from (A.24)

𝔼[(Xtk+1nYtk+1n)2]3𝔼[(XtknYtkn)2]+3πN2\displaystyle\mathbb{E}\left[\left(X_{t_{k+1}}^{n}-Y_{t_{k+1}}^{n}\right)^{2}\right]\leq 3\mathbb{E}\left[\left(X_{t_{k}}^{n}-Y_{t_{k}}^{n}\right)^{2}\right]+3\pi_{N}^{2} 𝔼[(XtknYtkn)2(ΔknL)2]\displaystyle\mathbb{E}\left[\left(X_{t_{k}}^{n}-Y_{t_{k}}^{n}\right)^{2}\left(\Delta^{n}_{k}L\right)^{2}\right]
+3πN2𝔼[(Xtknϵkn)2].\displaystyle+3\pi_{N}^{2}\mathbb{E}\left[\left(X_{t_{k}}^{n}\epsilon^{n}_{k}\right)^{2}\right].

Using the independence of ϵkn\epsilon^{n}_{k} and tk\mathcal{F}_{t_{k}} as well as of ΔknL\Delta^{n}_{k}L and tk\mathcal{F}_{t_{k}}, and the estimate 𝔼[(ΔknL)2]C1n\mathbb{E}\left[\left(\Delta^{n}_{k}L\right)^{2}\right]\leq\frac{C_{1}}{n} – where C1>0C_{1}>0 only depends on the Lévy triplet of LL – this further implies

𝔼[(Xtk+1nYtk+1n)2]3n+πN2C1n𝔼[(XtknYtkn)2]+3πN2𝔼[(Xtkn)2]𝔼[(ϵkn)2].\mathbb{E}\left[\left(X_{t_{k+1}}^{n}-Y_{t_{k+1}}^{n}\right)^{2}\right]\leq 3\frac{n+\pi_{N}^{2}C_{1}}{n}\mathbb{E}\left[\left(X_{t_{k}}^{n}-Y_{t_{k}}^{n}\right)^{2}\right]+3\pi_{N}^{2}\mathbb{E}\left[\left(X_{t_{k}}^{n}\right)^{2}\right]\mathbb{E}\left[\left(\epsilon_{k}^{n}\right)^{2}\right].

Since LL is square integrable, Lemma A.1 implies that ZknZ^{n}_{k} is also square integrable. This further gives 𝔼[(Xtkn)2]C2\mathbb{E}\left[(X^{n}_{t_{k}})^{2}\right]\leq C_{2}, where C2C_{2} can be chosen to be independent of nn and NN. The discrete Gronwall inequality of Lemma A.5 as well as X0n=Y0nX_{0}^{n}=Y_{0}^{n} imply

𝔼[(XTnYTn)2]C3i=0n1𝔼[(ϵin)2]ei=0n1πN2C1nC4n𝔼[(ϵ0n)2],\mathbb{E}\left[\left(X_{T}^{n}-Y_{T}^{n}\right)^{2}\right]\leq C_{3}\sum_{i=0}^{n-1}\mathbb{E}\left[\left(\epsilon^{n}_{i}\right)^{2}\right]e^{\sum_{i=0}^{n-1}\frac{\pi_{N}^{2}C_{1}}{n}}\leq C_{4}n\mathbb{E}\left[\left(\epsilon^{n}_{0}\right)^{2}\right],

where C4C_{4} is independent of nn and NN, since (πN)N(\pi_{N})_{N} are bounded.

We now establish 𝔼[(ϵ0n)2]=O(n2)\mathbb{E}[\left(\epsilon^{n}_{0}\right)^{2}]=O(n^{-2}) which concludes the proof: since LL is square integrable with triplet (b(x),c,F)(b(x),c,F), we can decompose it canonically into Lt=(Ltbt)+btL_{t}=(L_{t}-bt)+bt, where Mt:=(Ltbt)M_{t}:=(L_{t}-bt) is a square integrable martingale and Vt:=btV_{t}:=bt is the predictable finite variation process. Denoting by M,M\langle M,M\rangle the predictable quadratic variation of MM, we find M,Mt=(c+x2F(dx))t\langle M,M\rangle_{t}=(c+\int x^{2}\;F(dx))t. Lemma A.1 and [9, Theorem 4.40] show that 01/n(L)s1dMs\int_{0}^{1/n}\mathcal{E}\left(L\right)_{s-}-1\,dM_{s} is a square integrable martingale. Using the Itô isometry (for MtM_{t}) as well as Jensen’s inequality (for VtV_{t}), we find

𝔼[(ϵ0n)2]=𝔼[(01n(L)s1dLs)2]C5n+C5n01n𝔼[((L)s1)2]𝑑s,\mathbb{E}\left[\left(\epsilon^{n}_{0}\right)^{2}\right]=\mathbb{E}\left[\left(\int_{0}^{\frac{1}{n}}\mathcal{E}\left(L\right)_{s-}-1dL_{s}\right)^{2}\right]\leq\frac{C_{5}n+C_{5}}{n}\int_{0}^{\frac{1}{n}}\mathbb{E}\left[(\mathcal{E}\left(L\right)_{s}-1)^{2}\right]ds,

where C5C_{5} only depends on the Lévy triplet of LL. Using the same arguments as before we find for 0s1n0\leq s\leq\frac{1}{n}

𝔼[((L)s1)2]\displaystyle\mathbb{E}\left[(\mathcal{E}\left(L\right)_{s}-1)^{2}\right] =𝔼[(0s(L)u𝑑Lu)2],\displaystyle=\mathbb{E}\left[\left(\int_{0}^{s}\mathcal{E}\left(L\right)_{u-}\,dL_{u}\right)^{2}\right],
C5n+C5n0s𝔼[((L)u)2]𝑑uC6n,\displaystyle\leq\frac{C_{5}n+C_{5}}{n}\int_{0}^{s}\mathbb{E}\left[(\mathcal{E}\left(L\right)_{u})^{2}\right]\,du\leq\frac{C_{6}}{n},

where the last inequality follows from LL being square integrable and Lemma A.1. Hence, 𝔼[(ϵ0n)2]=O(n2)\mathbb{E}\left[\left(\epsilon^{n}_{0}\right)^{2}\right]=O(n^{-2}). ∎

Lemma A.5 (Discrete Gronwall Lemma).

Let (δi)i=0N(\delta_{i})_{i=0}^{N}, (ei)i=0N(e_{i})_{i=0}^{N}, (ηi)i=0N(\eta_{i})_{i=0}^{N} satisfy δi,ei,ηi0\delta_{i},e_{i},\eta_{i}\geq 0 and

ei+1(1+δi)ei+ηii=0,,N1.e_{i+1}\leq(1+\delta_{i})e_{i}+\eta_{i}\quad i=0,\ldots,N-1.

Then

ei(e0+j=0i1ηj)exp(j=0i1δj)i=0,,N.e_{i}\leq\left(e_{0}+\sum_{j=0}^{i-1}\eta_{j}\right)\exp\big{(}\sum_{j=0}^{i-1}\delta_{j}\big{)}\quad i=0,\ldots,N.
Proof.

Straight forward using induction. ∎

Lemma A.6 (omitted part of the proof of Proposition 4.4.(ii)).

Under the assumptions of Proposition 4.4.(ii), the functions

K(π,p)\displaystyle K(\pi,p) :=Iπ(π,p)=(1,)(x(1+πx)pxpx2(1+πx)1+p)𝑑F(x),\displaystyle:=\frac{\partial I}{\partial\pi}(\pi,p)=\int_{(-1,\infty)}\left(\frac{x}{(1+\pi x)^{p}}-x-\frac{px^{2}}{(1+\pi x)^{1+p}}\right)\,dF(x),
M(π,p)\displaystyle M(\pi,p) :=Ip(π,p)=(1,)xπ(1+πx)plog(1+xπ)𝑑F(x),\displaystyle:=\frac{\partial I}{\partial p}(\pi,p)=\int_{(-1,\infty)}\frac{-x\pi}{(1+\pi x)^{p}}\log(1+x\pi)\;dF(x),

are continuous for all π(𝒜0,)\{0}\pi\in(\mathcal{A}_{0,*})^{\circ}\backslash\{0\} and p(pa,pb)p\in(p_{a},p_{b}).

Proof.

We denote the integrand of KK by k(x,π,p)k(x,\pi,p), and the integrand of MM by m(x,π,p)m(x,\pi,p).

Let π(𝒜0,)\{0}\pi\in(\mathcal{A}_{0,*})^{\circ}\backslash\{0\}, p(pa,pb)p\in(p_{a},p_{b}) and assume first π>0\pi>0. Let πn(𝒜0,)\pi_{n}\in(\mathcal{A}_{0,*})^{\circ}, pn(pa,pb)p_{n}\in(p_{a},p_{b}) with πn>0\pi_{n}>0 and πnπ\pi_{n}\to\pi, pnpp_{n}\to p. Letting π¯:=infnπn\underline{\pi}:=\inf_{n}\pi_{n}, π¯:=supnπn\overline{\pi}:=\sup_{n}\pi_{n}, p¯:=infnpn\underline{p}:=\inf_{n}p_{n}, p¯:=supnpn\overline{p}:=\sup_{n}p_{n}, we find the following estimates which hold FF-a.s.:

|k(x,πn,pn)|\displaystyle|k(x,\pi_{n},p_{n})| |x|(1+π¯x)p¯+|x|+p¯x2(1+π¯x)1p¯for 1<x<0,\displaystyle\leq|x|(1+\overline{\pi}x)^{-\overline{p}}+|x|+\overline{p}x^{2}(1+\overline{\pi}x)^{-1-\overline{p}}\quad\text{for }-1<x<0,
|k(x,πn,pn)|\displaystyle|k(x,\pi_{n},p_{n})| x(1+π¯x)p¯+x+p¯x2(1+π¯x)1p¯for x0,\displaystyle\leq x(1+\underline{\pi}x)^{-\underline{p}}+x+\overline{p}x^{2}(1+\underline{\pi}x)^{-1-\underline{p}}\quad\text{for }x\geq 0,
|m(x,πn,pn)|\displaystyle|m(x,\pi_{n},p_{n})| |π¯xlog(1+π¯x)|(1+π¯x)p¯for 1<x<0,\displaystyle\leq\left|\overline{\pi}x\log(1+\overline{\pi}x)\right|(1+\overline{\pi}x)^{-\overline{p}}\quad\text{for }-1<x<0,
|m(x,πn,pn)|\displaystyle|m(x,\pi_{n},p_{n})| p¯1π¯xfor x0,\displaystyle\leq\underline{p}^{-1}\overline{\pi}x\quad\text{for }x\geq 0,

where we have used in the last inequality the estimate log(1+xπn)p1(1+xπn)p\log(1+x\pi_{n})\leq p^{-1}(1+x\pi_{n})^{p} for all x0x\geq 0, p>0p>0. We claim that the right hand side estimates are FF-integrable functions, which implies by dominated convergence that KK and MM are continuous for π>0\pi>0: this follows from our assumption that LL is integrable as well as that the integral of the functions on the right hand sides can be considered as derivatives of component-wise convex (concave) functions, as shown in the proof of Proposition 4.4. The integrals are finite since p¯,p¯,π¯,π¯\overline{p},\underline{p},\overline{\pi},\underline{\pi} are in the interior of (pa,pb)(p_{a},p_{b}) and 𝒜0,\mathcal{A}_{0,*}, respectively.

Letting π(𝒜0,)\{0}\pi\in(\mathcal{A}_{0,*})^{\circ}\backslash\{0\} with π<0\pi<0, p(pa,pb)p\in(p_{a},p_{b}) and πn(𝒜0,)\pi_{n}\in(\mathcal{A}_{0,*})^{\circ}, pn(pa,pb)p_{n}\in(p_{a},p_{b}) with πn<0\pi_{n}<0 and πnπ\pi_{n}\to\pi, pnpp_{n}\to p, we find with the same notation as above

|k(x,πn,pn)|\displaystyle|k(x,\pi_{n},p_{n})| |x|(1+π¯x)p¯+|x|+p¯x2(1+π¯x)1p¯for 1<x<0,\displaystyle\leq|x|(1+\overline{\pi}x)^{-\overline{p}}+|x|+\overline{p}x^{2}(1+\overline{\pi}x)^{-1-\underline{p}}\quad\text{for }-1<x<0,
|k(x,πn,pn)|\displaystyle|k(x,\pi_{n},p_{n})| x(1+π¯x)p¯+x+p¯x2(1+π¯x)1p¯for x0,\displaystyle\leq x(1+\underline{\pi}x)^{-\underline{p}}+x+\overline{p}x^{2}(1+\underline{\pi}x)^{-1-\overline{p}}\quad\text{for }x\geq 0,
|m(x,πn,pn)|\displaystyle|m(x,\pi_{n},p_{n})| p¯1|x||π¯|for 1<x<0,\displaystyle\leq\underline{p}^{-1}|x||\underline{\pi}|\quad\text{for }-1<x<0,
|m(x,πn,pn)|\displaystyle|m(x,\pi_{n},p_{n})| |π¯xlog(1+π¯x)|(1+π¯x)p¯for x0.\displaystyle\leq\left|\underline{\pi}x\log(1+\underline{\pi}x)\right|(1+\underline{\pi}x)^{-\overline{p}}\quad\text{for }x\geq 0.

Similar arguments as above show that the right hand side estimates are FF-integrable. Hence, KK and MM are also continuous for π<0\pi<0. ∎

References

  • [1] N. Bäuerle, S.P. Urban, and L.A.M. Veraart. The relaxed investor with partial information. Preprint, 2010.
  • [2] F.E. Benth, K. H. Karlsen, and K. Reikvam. Optimal portfolio selection with consumption and nonlinear integro-differential equations with gradient constraint: a viscosity solution approach. Finance and Stochastics, 5(3):275–303, 2001.
  • [3] J.E. Figueroa-López. Small-time moment asymptotics for Lévy processes. Statistics & Probability Letters, 78(18):3355–3365, 2008.
  • [4] L. Foldes. Certainty equivalence in the continuous-time portfolio-cum-saving model. In Applied stochastic analysis (London, 1989), volume 5 of Stochastics Monogr., pages 343–387. Gordon and Breach, New York, 1991.
  • [5] N.C. Framstad, B.K. Øksendal, and A. Sulem. Optimal consumption and portfolio in a jump diffusion market. Journal of Mathematical Economics, 1999.
  • [6] T. Goll and J. Kallsen. Optimal portfolios for logarithmic utility. Stochastic Processes and their Applications, 89(1):31–48, 2000.
  • [7] H. He. Optimal consumption-portfolio policies: a convergence from discrete to continuous time models. J. Econom. Theory, 55(2):340–363, 1991.
  • [8] J. Jacod. Asymptotic properties of power variations of Lévy processes. ESAIM. Probability and Statistics, 11:173–196, 2007.
  • [9] J. Jacod and A.N. Shiryaev. Limit Theorems for Stochastic Processes. Springer Berlin, second edition, 2003.
  • [10] J. Kallsen. Optimal portfolios for exponential Lévy processes. Mathematical Methods of Operations Research, 51(3):357–374, 2000.
  • [11] I. Karatzas and C. Kardaras. The numéraire portfolio in semimartingale financial models. Finance Stoch., 11(4):447–493, 2007.
  • [12] C. Kardaras. No-Free-Lunch equivalences for exponential Lévy models under convex constraints on investment. Mathematical Finance, 19(2):161–187, 2009.
  • [13] A. Kohatsu and P.E. Protter. The euler scheme for SDE’s driven by semimartingales. Stochastic Analysis on Infinite Dimensional Spaces, pages 141–151, 1994.
  • [14] D. Kramkov and W. Schachermayer. The asymptotic elasticity of utility functions and optimal investment in incomplete markets. Annals of Applied Probability, 9(3):904–950, 1999.
  • [15] R.C. Merton. Lifetime portfolio selection under uncertainty: The continuous-time case. The Review of Economics and Statistics, 51(3):247–257, 1969.
  • [16] M. Nutz. Power utility maximization in constrained exponential Lévy models. Mathematical Finance (forthcoming), 2009.
  • [17] M. Nutz. The bellman equation for power utility maximization with semimartingales. The Annals of Applied Probability, 22(1):363–406, 2012.
  • [18] L. C. G. Rogers and E. J. Stapleton. Utility maximisation with a time lag in trading. In Computational methods in decision-making, economics and finance, volume 74 of Appl. Optim., pages 249–269. Kluwer Acad. Publ., Dordrecht, 2002.
  • [19] L.C.G. Rogers. The relaxed investor and parameter uncertainty. Finance and Stochastics, 5(2):131–154, 2001.
  • [20] P.A. Samuelson. Lifetime portfolio selection by dynamic stochastic programming. The Review of Economics and Statistics, 51(3):239–246, 1969.
  • [21] K.I. Sato. Lévy Processes and Infinitely Divisible Distributions. Cambridge University Press, 1999.