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

Utility Indifference Pricing with High Risk Aversion and Small Linear Price Impact

Yan Dolinsky111Department of Statistics, Hebrew University of Jerusalem. . and Shir Moshe222Department of Statistics, Hebrew University of Jerusalem. . [email protected] [email protected]
Abstract

We consider the Bachelier model with linear price impact. Exponential utility indifference prices are studied for vanilla European options and we compute their non-trivial scaling limit for a vanishing price impact which is inversely proportional to the risk aversion. Moreover, we find explicitly a family of portfolios which are asymptotically optimal.

keywords:
Utility Indifference Pricing, Linear Price Impact, Asymptotic Analysis
33footnotetext: Both authors supported by the GIF Grant 1489-304.6/2019 and the ISF grant 230/21.

1 Introduction

In financial markets, trading moves prices against the trader: buying faster increases execution prices, and selling faster decreases them. This aspect of liquidity, known as market depth [4] or price-impact has recently received increasing attention (see, for instance, [2, 5, 7, 9, 10, 12, 13, 14, 15, 16, 17, 18] and the references therein).

In [14] the authors showed that for a reasonable market model, in the presence of price impact, super–replication is prohibitively costly. Namely, in the presence of price impact, even in market models such as the Bachelier model or the Black–Scholes model (which are complete in the frictionless setup) there is no practical way to construct a hedging strategy which eliminates all risk from a financial position. This brings us to utility indifference pricing.

In this paper we study utility indifference pricing for vanilla European options in the multi–dimensional Bachelier model with linear price impact. Our main result is computing the asymptotic behavior of the exponential utility indifference prices where the risk aversion goes to infinity at a rate which is inversely proportional to the linear price impact which goes to zero. In addition we provide a family of asymptotically optimal hedging strategies.

This type of scaling limits goes back to the seminal work of Barles and Soner [8] which determines the scaling limit of utility indifference prices of vanilla options for small proportional transaction costs and high risk aversion. The present note provides an analogous analysis for the case of linear price impact which results in quadratic transaction costs, albeit using probabilistic techniques rather than taking a PDE approach as pursued in [8].

We divide the proof of our main result, namely Theorem 2.1 into two main steps: the proof of the lower bound and the proof of the upper bound.

The proof of the lower bound goes through a dual representation of the certainty equivalent. In the dual problem, there is only one player: a maximizing (adverse) player that controls the probability measure. A key ingredient in the proof of the lower bound is a construction of a family of probability measures which attain (in the asymptotic sense) the desired limit. This is done in Proposition 3.3.

The proof of the upper bound does not use duality and is based on a direct argument (Proposition 4.1). More precisely, we construct a family of trading strategies for which the expected utility converges to the scaling limit. Roughly speaking, these strategies are given by a reversion towards the Δ\Delta–hedging strategy which corresponds to a modified European option and a modified stock price.

The rest of the paper is organized as follows. In the next section we introduce the setup and formulate the main results. In Section 3 we discuss the dual representation and prove the lower bound. In Section 4 we prove the upper bound.

2 Preliminaries and Main Results

Let T<T<\infty be the time horizon and let W=(Wt1,,Wtd)t[0,T]W=\left(W^{1}_{t},...,W^{d}_{t}\right)_{t\in[0,T]} be a standard dd-dimensional Brownian motion defined on the filtered probability space (Ω,,(t)t[0,T],)(\Omega,\mathcal{F},(\mathcal{F}_{t})_{t\in[0,T]},\mathbb{P}) where (t)t[0,T](\mathcal{F}_{t})_{t\in[0,T]} is the (augmented) filtration generated by WW. We consider a simple financial market with a riskless savings account bearing zero interest (for simplicity) and with dd-risky asset S=(St1,,Std)t[0,T]S=\left(S^{1}_{t},...,S^{d}_{t}\right)_{t\in[0,T]} with Bachelier price dynamics

(1) Sti=s0i+μit+j=1dσijWtjS^{i}_{t}=s^{i}_{0}+\mu^{i}t+\sum_{j=1}^{d}\sigma^{ij}W^{j}_{t}

where s0=(s01,,s0d)ds_{0}=(s^{1}_{0},...,s^{d}_{0})\in\mathbb{R}^{d} is the initial position of the risky assets, μ=(μ1,μd)d\mu=(\mu^{1},...\mu^{d})\in\mathbb{R}^{d} is a constant vector (drift) and σ={σij}1i,jdMd()\sigma=\{\sigma^{ij}\}_{1\leq i,j\leq d}\in M_{d}(\mathbb{R}) is a constant nonsingular matrix (volatility). Without loss of generality we assume that the constant nonsingular matrix σ\sigma is a positive definite matrix.

Following [1], we model the investor’s market impact, in a temporary linear form and, thus, when at time tt the investor turns over her position at the ii–asset Φti\Phi^{i}_{t} at the rate ϕti:=Φ˙ti\phi^{i}_{t}:=\dot{\Phi}^{i}_{t} the execution price is St+Λ2ϕtiS_{t}+\frac{\Lambda}{2}\phi^{i}_{t} for some constant Λ>0\Lambda>0. As a result, the profits and losses from a trading strategy ϕ=(ϕt1,,ϕtd)t[0,T]\phi=\left(\phi^{1}_{t},...,\phi^{d}_{t}\right)_{t\in[0,T]} with the initial position Φ0=(Φ01,,Φ0d)\Phi_{0}=\left(\Phi^{1}_{0},...,\Phi^{d}_{0}\right) are given by

(2) VtΦ0,ϕ:=0tΦu,dSuΛ20tϕv2𝑑v,t[0,T]V^{\Phi_{0},\phi}_{t}:=\int_{0}^{t}\langle\Phi_{u},dS_{u}\rangle-\frac{\Lambda}{2}\int_{0}^{t}||\phi_{v}||^{2}dv,\qquad t\in[0,T]

where, for convenience, we assume that the investor marks to market her position Φt=Φ0+0tϕv𝑑v\Phi_{t}=\Phi_{0}+\int_{0}^{t}\phi_{v}dv in the risky asset that she has acquired by time tt. As usual, ,\langle\cdot,\cdot\rangle and ||||||\cdot||, denotes the standard scalar product and the Euclidean norm, respectively. In our setup, the natural class of admissible strategies is

𝒜:={ϕ=(ϕt1,,ϕtd)t[0,T]:ϕ is -adapted with 0Tϕt2𝑑t< a.s.}.\mathcal{A}:=\left\{\phi=\left(\phi^{1}_{t},...,\phi^{d}_{t}\right)_{t\in[0,T]}:\ \phi\text{ is }\ \mathcal{F}\text{-adapted with }\int_{0}^{T}||\phi_{t}||^{2}dt<\infty\ \text{ a.s.}\right\}.
Remark 2.1.

Let us notice that by scaling the risky assets, there is no loss of generality in assuming that the constant Λ>0\Lambda>0 which represents the linear price impact is the same for all risky assets.

Next, consider a vanilla European option with the payoff f(ST)f\left(S_{T}\right) where f:df:\mathbb{R}^{d}\rightarrow\mathbb{R} is a Lipschitz continuous function. The investor will assess the quality of a hedge by the resulting expected utility. Assuming exponential utility with constant absolute risk aversion α>0\alpha>0, the utility indifference price and the certainty equivalent price of one unit of the claim f(ST)f\left(S_{T}\right) (see, e.g., [11] for details on indifference prices) do not depend on the investor’s initial wealth and, respectively, take the well-known forms

(3) π(Λ,α,Φ0,f):=1αlog(infϕ𝒜𝔼[exp(α(f(ST)VTΦ0,ϕ))]infϕ𝒜𝔼[exp(αVTΦ0,ϕ)])\pi(\Lambda,\alpha,\Phi_{0},f):=\frac{1}{\alpha}\log\left(\frac{\inf_{\phi\in\mathcal{A}}\mathbb{E}_{\mathbb{P}}\left[\exp\left(\alpha\left(f(S_{T})-V^{\Phi_{0},\phi}_{T}\right)\right)\right]}{\inf_{\phi\in\mathcal{A}}\mathbb{E}_{\mathbb{P}}\left[\exp\left(-\alpha V^{\Phi_{0},\phi}_{T}\right)\right]}\right)

and

c(Λ,α,Φ0,f):=1αlog(infϕ𝒜𝔼[exp(α(f(ST)VTΦ0,ϕ))]).c(\Lambda,\alpha,\Phi_{0},f):=\frac{1}{\alpha}\log\left(\inf_{\phi\in\mathcal{A}}\mathbb{E}_{\mathbb{P}}\left[\exp\left(\alpha\left(f(S_{T})-V^{\Phi_{0},\phi}_{T}\right)\right)\right]\right).

If the risk aversion α>0\alpha>0 is fixed, then by applying standard density arguments we obtain that for Λ0\Lambda\downarrow 0, the above indifference price converges to the unique price of the continuous time complete (frictionless) market given by (1). A more interesting limit emerges, however, if we re-scale the investor’s risk-aversion in the form α:=A/Λ\alpha:=A/\Lambda. Before we formulate the limit theorem we need some preparations.

For a given A>0A>0 introduce the functions

(4) gA(x):=supyd[f(x+y)yσ1,y2A],x=(x1,,xd)dg^{A}(x):=\sup_{y\in\mathbb{R}^{d}}\left[f(x+y)-\frac{\langle y\sigma^{-1},y\rangle}{2\sqrt{A}}\right],\quad x=(x^{1},...,x^{d})\in\mathbb{R}^{d}

and

(5) uA(t,x):=𝔼[gA(x+WTtσ)],(t,x)[0,T]×du^{A}(t,x):=\mathbb{E}_{\mathbb{P}}\left[g^{A}(x+W_{T-t}\sigma)\right],\quad(t,x)\in[0,T]\times\mathbb{R}^{d}

where the vectors x,y,Wx,y,W are considered as row vectors and for any row vector zdz\in\mathbb{R}^{d}, zσ1,zσdz\sigma^{-1},z\sigma\in\mathbb{R}^{d} are the standard matrix products. The term uA(t,St)u^{A}(t,S_{t}) represents the price at time tt of a European option with the payoff gA(ST)g^{A}(S_{T}) in the complete market given by (1). It is well known that uC1,2([0,T)×d)u\in C^{1,2}([0,T)\times\mathbb{R}^{d}) solves the PDE

(6) uAt+tr(σ2Dx2uA)2=0in[0,T)×\frac{\partial u^{A}}{\partial t}+\frac{tr\left(\sigma^{2}D^{2}_{x}u^{A}\right)}{2}=0\ \ \ \ \mbox{in}\ \ [0,T)\times\mathbb{R}

where tr()tr(\cdot) is the trace of the square matrix \cdot and Dx2uAD^{2}_{x}u^{A} is the Hessian matrix with respect to x=(x1,,xd)x=(x^{1},...,x^{d}) which is given by [Dx2uA]ij:=2uAxixj[D^{2}_{x}u^{A}]_{ij}:=\frac{\partial^{2}u^{A}}{\partial x^{i}\partial x^{j}}, 1i,jd1\leq i,j\leq d.

For a given A,Λ>0A,\Lambda>0 consider the dd–dimensional (random) ODE (Φ\Phi is a row vector)

(7) ϕt:=Φ˙t=AΛ(DxuA(t,StAΦtσ)Φt)σ,t[0,T)\phi_{t}:=\dot{\Phi}_{t}=\frac{\sqrt{A}}{\Lambda}\left(D_{x}u^{A}\left(t,S_{t}-\sqrt{A}\Phi_{t}\sigma\right)-\Phi_{t}\right)\sigma,\ \ \ \ t\in[0,T)

where DxuA:=(uAx1,,uAxd)dD_{x}u^{A}:=\left(\frac{\partial u^{A}}{\partial x^{1}},...,\frac{\partial u^{A}}{\partial x^{d}}\right)\in\mathbb{R}^{d} is the gradient with respect to xx. From the linear growth of ff it follows that for any ϵ>0\epsilon>0 the function DxuA,Dx2uAD_{x}u^{A},D^{2}_{x}u^{A} are uniformly bounded in the domain [0,Tϵ]×d[0,T-\epsilon]\times\mathbb{R}^{d}. In particular DxuAD_{x}u^{A} is Lipschitz continuous with respect to xx in the domain [0,Tϵ]×d[0,T-\epsilon]\times\mathbb{R}^{d}. Hence, from the standard theory of ODE (see Chapter II, Section 6 in [19]) we obtain that for a given initial value Φ0\Phi_{0} there exists a unique solution to (7) which we denote by (ΦtA,Λ)0t<T(\Phi^{A,\Lambda}_{t})_{0\leq t<T}. Next, the Lipschitz continuity of ff implies that gAg^{A} is a Lipschitz continuous function (with the same constant as ff), and so DxuAD_{x}u^{A} is uniformly bounded in [0,T)×d[0,T)\times\mathbb{R}^{d}. This together with the mean reverting structure of the ODE (7) yields that limtTΦtA,Λ\lim_{t\rightarrow T-}\Phi^{A,\Lambda}_{t} exists and finite a.s. Thus, we can extend ΦA,Λ\Phi^{A,\Lambda} to the interval [0,T][0,T] by ΦtA,Λ:=limtTΦtA,Λ\Phi^{A,\Lambda}_{t}:=\lim_{t\rightarrow T-}\Phi^{A,\Lambda}_{t} and we define ϕA,Λ𝒜\phi^{A,\Lambda}\in\mathcal{A} by ϕtA,Λ:=Φ˙tA,Λ\phi^{A,\Lambda}_{t}:=\dot{\Phi}^{A,\Lambda}_{t} for t<Tt<T and ϕTA,Λ=0\phi^{A,\Lambda}_{T}=0. Obviously,

ΦtA,Λ:=Φ0+0tϕvA,Λ𝑑v,t[0,T].\Phi^{A,\Lambda}_{t}:=\Phi_{0}+\int_{0}^{t}\phi^{A,\Lambda}_{v}dv,\ \ t\in[0,T].
Remark 2.2.

In words, the ODE (7) says that the solution ΦA,Λ\Phi^{A,\Lambda} is tracking the Δ\Delta–hedging strategy which corresponds to the modified payoff gAg^{A} and the shifted stock price StAΦtA,ΛσS_{t}-\sqrt{A}\Phi^{A,\Lambda}_{t}\sigma. We notice that the shift depends on the solution ΦA,Λ\Phi^{A,\Lambda}.

We arrive at the main result of the paper which provides an explicit formula for the asymptotic behavior of the certainty equivalent and an optimal family (it should not be unique) of hedging strategies in the asymptotic sense.

Theorem 2.1.

For vanishing linear price impact Λ0\Lambda\downarrow 0 and re-scaled high risk-aversion A/ΛA/\Lambda with A>0A>0 fixed, the certainty equivalent of f(ST)f(S_{T}) has the scaling limit

(8) limΛ0c(Λ,A/Λ,Φ0,f)=uA(0,s0AΦ0σ)+AΦ0σ,Φ02.\lim_{\Lambda\downarrow 0}c(\Lambda,A/\Lambda,\Phi_{0},f)=u^{A}\left(0,s_{0}-\sqrt{A}\Phi_{0}\sigma\right)+\frac{\sqrt{A}\langle\Phi_{0}\sigma,\Phi_{0}\rangle}{2}.

Moreover, we have,

limΛ0ΛAlog(𝔼[exp(AΛ(f(ST)VTΦ0,ϕA,Λ))])\displaystyle\lim_{\Lambda\downarrow 0}\frac{\Lambda}{A}\log\left(\mathbb{E}_{\mathbb{P}}\left[\exp\left(\frac{A}{\Lambda}\left(f(S_{T})-V^{\Phi_{0},\phi^{A,\Lambda}}_{T}\right)\right)\right]\right)
=uA(0,s0AΦ0σ)+AΦ0σ,Φ02.\displaystyle=u^{A}\left(0,s_{0}-\sqrt{A}\Phi_{0}\sigma\right)+\frac{\sqrt{A}\langle\Phi_{0}\sigma,\Phi_{0}\rangle}{2}.

From Theorem 2.1 we obtain immediately the following corollary which says that the asymptotic value of the utility indifference prices is equal to the price of the vanilla European option with the payoff gA(ST)g^{A}(S_{T}) and the shifted initial stock price s0AΦ0σs_{0}-\sqrt{A}\Phi_{0}\sigma.

Corollary 2.2.

For vanishing linear price impact Λ0\Lambda\downarrow 0 and re-scaled high risk-aversion A/ΛA/\Lambda with A>0A>0 fixed, the utility indifference price of f(ST)f(S_{T}) has the scaling limit

limΛ0π(Λ,A/Λ,Φ0,f)=uA(0,s0AΦ0σ).\lim_{\Lambda\downarrow 0}\pi(\Lambda,A/\Lambda,\Phi_{0},f)=u^{A}\left(0,s_{0}-\sqrt{A}\Phi_{0}\sigma\right).

Proof 2.3.

Apply (8) and take f0f\equiv 0 for the denominator of (3).

The following remark provides a possible application of Corollary 2.2.

Remark 3.

Let π^(Λ,α,Φ0,q):=1qπ(Λ,α,Φ0,qf)\hat{\pi}(\Lambda,\alpha,\Phi_{0},q):=\frac{1}{q}\pi(\Lambda,\alpha,\Phi_{0},qf) be the per-unit indifference price for selling qq units of the option. From (2) (apply the bijection ϕϕΛ\phi\rightarrow\frac{\phi}{\Lambda}) it follows that π(Λ,AΛ,Φ0,f)=π^(Λ2,A,Φ0Λ,1Λ)\pi\left(\Lambda,\frac{A}{\Lambda},\Phi_{0},f\right)=\hat{\pi}\left(\Lambda^{2},A,\frac{\Phi_{0}}{\Lambda},\frac{1}{\Lambda}\right). Hence, Corollary 2.2 can be viewed as a limit theorem for per-unit utility indifference prices for the case of vanishing linear price impact and large position sizes. An interesting question is whether the theory which was developed in [3] can be applied for obtaining the asymptotic behaviour of optimal position sizes (for the exact definition see [3]). We leave this question for future research.

We end this section with the following example.

Example 4.

Consider a European option with the payoff

f(x)=(a,x+b)+,xdf(x)=\left(\left\langle a,x\right\rangle+b\right)^{+},\ \ x\in\mathbb{R}^{d}

for some constant ada\in\mathbb{R}^{d} and bb\in\mathbb{R}. Then we have

gA(x):=supyd[(a,x+y+b)+yσ1,y2A],xd.g^{A}(x):=\sup_{y\in\mathbb{R}^{d}}\left[\left(\left\langle a,x+y\right\rangle+b\right)^{+}-\frac{\left\langle y\sigma^{-1},y\right\rangle}{2\sqrt{A}}\right],\quad x\in\mathbb{R}^{d}.

Clearly, the quadratic pattern ya,yyσ1,y2Ay\rightarrow\langle a,y\rangle-\frac{\langle y\sigma^{-1},y\rangle}{2\sqrt{A}} attains its maximum at y:=Aaσy^{*}:=\sqrt{A}a\sigma. This together with the obvious inequality gAf0g^{A}\geq f\geq 0 yields that

gA(x)=(a,x+y+byσ1,y2A)+=(a,x+b+Aaσ,a2)+.g^{A}(x)=\left(\left\langle a,x+y^{*}\right\rangle+b-\frac{\left\langle y^{*}\sigma^{-1},y^{*}\right\rangle}{2\sqrt{A}}\right)^{+}=\left(\left\langle a,x\right\rangle+b+\frac{\sqrt{A}\left\langle a\sigma,a\right\rangle}{2}\right)^{+}.

Refer to caption
Figure 1: We consider a one dimensional Bachelier model with σ=1\sigma=1. We assume that the initial number of stocks is Φ0=0\Phi_{0}=0. The maturity date is T=1T=1 and the initial stock price is s0=8s_{0}=8. Let f(ST)=(ST8)+f(S_{T})=(S_{T}-8)^{+}, i.e. we take at the money call option. We plot the limiting utility indifference price as a function of the parameter AA. Namely we plot the function AuA(0,s0)A\rightarrow u^{A}(0,s_{0}).

3 The Dual Problem and the Lower Bound

In this section we establish the inequality \geq in (8).

We start with the following lemma.

Lemma 3.1.

Denoting by 𝒬\mathcal{Q} the set of all probability measures \mathbb{Q}\sim\mathbb{P} with finite entropy 𝔼[log(dd)]<\mathbb{E}_{\mathbb{Q}}\left[\log\left(\frac{d\mathbb{Q}}{d\mathbb{P}}\right)\right]<\infty relative to \mathbb{P}, we have

c(Λ,α,Φ0,f)\displaystyle c(\Lambda,\alpha,\Phi_{0},f)
(9) sup𝒬𝔼[f(ST)Φ0,STs01αlog(dd)\displaystyle\geq\sup_{\mathbb{Q}\in\mathcal{Q}}\mathbb{E}_{\mathbb{Q}}\bigg{[}f\left(S_{T}\right)-\langle\Phi_{0},S_{T}-s_{0}\rangle-\frac{1}{\alpha}\log\left(\frac{d\mathbb{Q}}{d\mathbb{P}}\right)
12Λ0T||St𝔼(ST|t)||2dt].\displaystyle-\frac{1}{2\Lambda}\int_{0}^{T}\left|\left|S_{t}-\mathbb{E}_{\mathbb{Q}}\left(S_{T}|\mathcal{F}_{t}\right)\right|\right|^{2}dt\bigg{]}.

Proof 3.2.

The proof rests on the classical Legendre-Fenchel duality inequality

(10) pqep+q(logq1),p,q>0.pq\leq e^{p}+q(\log q-1),\ \ \ p\in\mathbb{R},\ q>0.

Let ϕ𝒜\phi\in\mathcal{A} and 𝒬\mathbb{Q}\in\mathcal{Q}. From the Girsanov theorem it follows that there exists a process θ=(θt1,,θtd)t[0,T]\theta=(\theta^{1}_{t},...,\theta^{d}_{t})_{t\in[0,T]} such that Wt:=Wt0tθv𝑑vW^{\mathbb{Q}}_{t}:=W_{t}-\int_{0}^{t}\theta_{v}dv, t[0,T]t\in[0,T] is a \mathbb{Q}–Brownian motion. Moreover, from the equality

Z:=dd=exp(0Tθt,dWt120Tθt2𝑑t)Z:=\frac{d\mathbb{Q}}{d\mathbb{P}}=\exp\left(\int_{0}^{T}\langle\theta_{t},dW_{t}\rangle-\frac{1}{2}\int_{0}^{T}||\theta_{t}||^{2}dt\right)

and the fact that 𝒬\mathbb{Q}\in\mathcal{Q} we obtain that

𝔼[logZ]=12𝔼[0Tθt2𝑑t]<\mathbb{E}_{\mathbb{Q}}\left[\log Z\right]=\frac{1}{2}\mathbb{E}_{\mathbb{Q}}\left[\int_{0}^{T}||\theta_{t}||^{2}dt\right]<\infty

and so from (1)

(11) 𝔼[sup0tTSt2]<.\mathbb{E}_{\mathbb{Q}}\left[\sup_{0\leq t\leq T}||S_{t}||^{2}\right]<\infty.

Without loss of generality we assume that 𝔼[eα(f(ST)VTΦ0,ϕ)]<.\mathbb{E}_{\mathbb{P}}\left[e^{\alpha\left(f(S_{T})-V^{\Phi_{0},\phi}_{T}\right)}\right]<\infty. Thus, from (10) we obtain

(12) α𝔼[(f(ST)VTΦ0,ϕ)+]𝔼[eα(f(ST)VTΦ0,ϕ)]+𝔼[logZ]<.\alpha\mathbb{E}_{\mathbb{Q}}\left[\left(f(S_{T})-V^{\Phi_{0},\phi}_{T}\right)^{+}\right]\leq\mathbb{E}_{\mathbb{P}}\left[e^{\alpha\left(f(S_{T})-V^{\Phi_{0},\phi}_{T}\right)}\right]+\mathbb{E}_{\mathbb{Q}}\left[\log Z\right]<\infty.

Next, from (2) and the integration by parts formula it follows that

(13) VTΦ0,ϕ=Φ0,STs0+0T(ϕt,STStΛ2ϕt2)𝑑t.V^{\Phi_{0},\phi}_{T}=\langle\Phi_{0},S_{T}-s_{0}\rangle+\int_{0}^{T}\left(\langle\phi_{t},S_{T}-S_{t}\rangle-\frac{\Lambda}{2}\ ||\phi_{t}||^{2}\right)dt.

Hence, from the simple inequality

ϕt,STStΛ4ϕt21ΛSTSt2\langle\phi_{t},S_{T}-S_{t}\rangle-\frac{\Lambda}{4}||\phi_{t}||^{2}\leq\frac{1}{\Lambda}||S_{T}-S_{t}||^{2}

we obtain

f(ST)VTΦ0,ϕVTΦ0,ϕ\displaystyle f(S_{T})-V^{\Phi_{0},\phi}_{T}\geq-V^{\Phi_{0},\phi}_{T}
Φ0,STs0+0T(Λ4ϕt21ΛSTSt2)𝑑t.\displaystyle\geq-\langle\Phi_{0},S_{T}-s_{0}\rangle+\int_{0}^{T}\left(\frac{\Lambda}{4}||\phi_{t}||^{2}-\frac{1}{\Lambda}||S_{T}-S_{t}||^{2}\right)dt.

This together with (11)–(12) gives that

(14) 𝔼[0Tϕt2𝑑t]<.\mathbb{E}_{\mathbb{Q}}\left[\int_{0}^{T}||\phi_{t}||^{2}dt\right]<\infty.

From (10)–(11) and (13)–(14) we obtain that for any γ>0\gamma>0

𝔼[eα(f(ST)VTΦ0,ϕ)]\displaystyle\mathbb{E}_{\mathbb{P}}\left[e^{\alpha\left(f(S_{T})-V^{\Phi_{0},\phi}_{T}\right)}\right]
𝔼[αγZ(f(ST)Φ0,STs00T(ϕt,STStΛ2ϕt2)𝑑t)]\displaystyle\geq\mathbb{E}_{\mathbb{P}}\left[\alpha\gamma Z\left(f(S_{T})-\langle\Phi_{0},S_{T}-s_{0}\rangle-\int_{0}^{T}\left(\langle\phi_{t},S_{T}-S_{t}\rangle-\frac{\Lambda}{2}\ ||\phi_{t}||^{2}\right)dt\right)\right]
𝔼[γZ(log(γZ)1)]\displaystyle-\mathbb{E}_{\mathbb{P}}\left[\gamma Z\left(\log(\gamma Z)-1\right)\right]
=αγ𝔼[f(ST)Φ0,STs0]\displaystyle=\alpha\gamma\mathbb{E}_{\mathbb{Q}}\left[f(S_{T})-\langle\Phi_{0},S_{T}-s_{0}\rangle\right]
αγ𝔼[0T(ϕt,𝔼[ST|t]StΛ2ϕt2)𝑑t]\displaystyle-\alpha\gamma\mathbb{E}_{\mathbb{Q}}\left[\int_{0}^{T}\left(\langle\phi_{t},\mathbb{E}_{\mathbb{Q}}[S_{T}|\mathcal{F}_{t}]-S_{t}\rangle-\frac{\Lambda}{2}||\phi_{t}||^{2}\right)dt\right]
γ(logγ1)γ𝔼[logZ]\displaystyle-\gamma(\log\gamma-1)-\gamma\mathbb{E}_{\mathbb{Q}}\left[\log Z\right]
αγ𝔼[f(ST)Φ0,STs012Λ0T||𝔼[ST|t]St||2dt]\displaystyle\geq\alpha\gamma\mathbb{E}_{\mathbb{Q}}\left[f(S_{T})-\langle\Phi_{0},S_{T}-s_{0}\rangle-\frac{1}{2\Lambda}\int_{0}^{T}\left|\left|\mathbb{E}_{\mathbb{Q}}[S_{T}|\mathcal{F}_{t}]-S_{t}\right|\right|^{2}dt\right]
(15) γ(logγ1)γ𝔼[logZ]\displaystyle-\gamma(\log\gamma-1)-\gamma\mathbb{E}_{\mathbb{Q}}\left[\log Z\right]

where the last inequality follows from the maximization of the quadratic pattern

ϕtϕt,𝔼[ST|t]StΛ2ϕt2,t[0,T].\phi_{t}\rightarrow\langle\phi_{t},\mathbb{E}_{\mathbb{Q}}[S_{T}|\mathcal{F}_{t}]-S_{t}\rangle-\frac{\Lambda}{2}||\phi_{t}||^{2},\ \ \ \ t\in[0,T].

Optimizing (3.2) in γ>0\gamma>0 we arrive at

1αlog(𝔼[eα(f(ST)VTΦ0,ϕ)])\displaystyle\frac{1}{\alpha}\log\left(\mathbb{E}_{\mathbb{P}}\left[e^{\alpha\left(f(S_{T})-V^{\Phi_{0},\phi}_{T}\right)}\right]\right)
𝔼[f(ST)Φ0,STs012Λ0T||𝔼[ST|t]St||2dt1αlogZ].\displaystyle\geq\mathbb{E}_{\mathbb{Q}}\left[f(S_{T})-\langle\Phi_{0},S_{T}-s_{0}\rangle-\frac{1}{2\Lambda}\int_{0}^{T}\left|\left|\mathbb{E}_{\mathbb{Q}}[S_{T}|\mathcal{F}_{t}]-S_{t}\right|\right|^{2}dt-\frac{1}{\alpha}\log Z\right].

Since ϕ𝒜\phi\in\mathcal{A} and 𝒬\mathbb{Q}\in\mathcal{Q} were arbitrary we complete the proof.

Remark 1.

By mimicking the arguments of Proposition A.2 in [6] to the multidimensional case one can show that the inequality in (3.1) is in fact an equality. Namely, there is no duality gap. However, since we need only the lower bound of the duality then we just provide a self contained proof for (3.1).

Next, we fix A>0A>0 and prove the following key result.

Proposition 3.3.

Let h:ddh:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d} be a bounded and measurable function and let Y=h(WT)Y=h(W_{T}). Then,

liminfΛ0c(Λ,A/Λ,Φ0,f)𝔼[f(s0+WTσY)+Φ0,Y12AY,Yσ1].\lim\inf_{\Lambda\downarrow 0}c\left(\Lambda,A/\Lambda,\Phi_{0},f\right)\geq\mathbb{E}_{\mathbb{P}}\left[f\left(s_{0}+W_{T}\sigma-Y\right)+\langle\Phi_{0},Y\rangle-\frac{1}{2\sqrt{A}}\left\langle Y,Y\sigma^{-1}\right\rangle\right].

Proof 3.4.

For a matrix valued process Ξ={Ξtij}1i,jd,0tT\Xi=\{\Xi^{ij}_{t}\}_{1\leq i,j\leq d,0\leq t\leq T} denote 0tΞs𝑑Ws\int_{0}^{t}\Xi_{s}\cdot dW_{s} the row vector (Z1,,Zd)(Z^{1},...,Z^{d}) which is given by Zi=0tj=1dΞsijdWsjZ^{i}=\int_{0}^{t}\sum_{j=1}^{d}\Xi^{ij}_{s}dW^{j}_{s}. By applying the martingale representation theorem and standard density arguments it follows that without loss of generality we can assume that Y=y+0TH(s,Ws)𝑑WsY=y+\int_{0}^{T}H(s,W_{s})\cdot dW_{s} where ydy\in\mathbb{R}^{d} is a constant vector and H:[0,T]×dMd()H:[0,T]\times\mathbb{R}^{d}\rightarrow M_{d}(\mathbb{R}) is a continuous and bounded function which satisfies H[Tδ,T]×d0H_{[T-\delta,T]\times\mathbb{R}^{d}}\equiv 0 for some δ>0\delta>0.

The proof of Proposition 3.3 is done in three steps. In the first step we fix Λ>0\Lambda>0 and construct a sequence of probability measures n\mathbb{Q}_{n}\sim\mathbb{P}, nn\in\mathbb{N} with finite entropy. In the second step we estimate (for the constructed probability measures) the asymptotic value of the components in the right hand side of (3.1) for α:=A/Λ\alpha:=A/\Lambda. In the last step we summarize the computations and apply Lemma 3.1.

Step I: Recall the drift vector μ=(μ1,,μd)\mu=(\mu^{1},...,\mu^{d}) which appears in (1). For any Λ>0\Lambda>0 and 0stT0\leq s\leq t\leq T define

Kt,sΛ:=cosh(A(Tt)σ/Λ)(sinh(A(Ts)σ/Λ))1\displaystyle K^{\Lambda}_{t,s}:=\cosh\left(\sqrt{A}(T-t)\sigma/\Lambda\right)\left(\sinh\left(\sqrt{A}(T-s)\sigma/\Lambda\right)\right)^{-1}
andatΛ:=μσ1+AΛyKt,0Λ\displaystyle\mbox{and}\ \ a^{\Lambda}_{t}:=\mu\sigma^{-1}+\frac{\sqrt{A}}{\Lambda}yK^{\Lambda}_{t,0}

where for any square matrix ξ\xi

cosh(ξ):=exp(ξ)+exp(ξ)2,sinh(ξ):=exp(ξ)exp(ξ)2\cosh(\xi):=\frac{\exp(\xi)+\exp(-\xi)}{2},\ \ \sinh(\xi):=\frac{\exp(\xi)-\exp(-\xi)}{2}

and exp(ξ)\exp(\xi) is the matrix exponential of ξ\xi.

Fix Λ>0\Lambda>0. For any nn\in\mathbb{N} define the processes (which are also depend on Λ\Lambda) Wn=(Wtn,1,,Wtn,d)t[0,T]W^{n}=(W^{n,1}_{t},...,W^{n,d}_{t})_{t\in[0,T]} and θn=(θtn,1,,θtn,d)t[0,T]\theta^{n}=(\theta^{n,1}_{t},...,\theta^{n,d}_{t})_{t\in[0,T]} by the following recursive relations. For t[0,T/n]t\in[0,T/n]

θtn:=atΛ,Wtn:=Wt+0tθvn𝑑v\theta^{n}_{t}:=a^{\Lambda}_{t},\ \ W^{n}_{t}:=W_{t}+\int_{0}^{t}\theta^{n}_{v}dv

and for k=1,,n1k=1,...,n-1, t(kT/n,(k+1)T/n]t\in(kT/n,(k+1)T/n]

θtn:=atΛ+𝕀κtn<nκtn,Wtn:=Wt+0tθvn𝑑v\theta^{n}_{t}:=a^{\Lambda}_{t}+\mathbb{I}_{||\kappa^{n}_{t}||<n}\kappa^{n}_{t},\ \ \ W^{n}_{t}:=W_{t}+\int_{0}^{t}\theta^{n}_{v}dv

where 𝕀\mathbb{I} denotes the indicator function and

κtn:=AΛ0kTnKt,sΛH(s,Wsn)𝑑Wsn,t(kT/n,(k+1)T/n].\kappa^{n}_{t}:=\frac{\sqrt{A}}{\Lambda}\int_{0}^{\frac{kT}{n}}K^{\Lambda}_{t,s}H(s,W^{n}_{s})\cdot dW^{n}_{s},\ \ \ \ t\in(kT/n,(k+1)T/n].

Clearly, for any nn the process θn\theta^{n} is a bounded process, and so from the Girsanov theorem there exists a probability measure n\mathbb{Q}^{n} such that WnW^{n} is a n\mathbb{Q}^{n}–Brownian motion. For these probability measures we have the weak convergence (on the space of continuous function Cd[0,T]C_{d}[0,T])

(16) Qnθn(aΛ+κΛ)Q^{n}\circ\theta^{n}\Rightarrow\mathbb{P}\circ\left(a^{\Lambda}+\kappa^{\Lambda}\right)

where κΛ=(κtΛ)t[0,T]\kappa^{\Lambda}=(\kappa^{\Lambda}_{t})_{t\in[0,T]} is given by κtΛ:=AΛ0tKt,sΛH(s,Ws)𝑑Ws.\kappa^{\Lambda}_{t}:=\frac{\sqrt{A}}{\Lambda}\int_{0}^{t}K^{\Lambda}_{t,s}H(s,W_{s})\cdot dW_{s}.

Step II: From the Fubini theorem, (1), (16) and and the simple equality AΛsTKt,sΛdt=σ1\frac{\sqrt{A}}{\Lambda}\int_{s}^{T}K^{\Lambda}_{t,s}dt=\sigma^{-1} s[0,T]\forall s\in[0,T], we get the weak convergence

(17) nST(s0+WTσY).\mathbb{Q}^{n}\circ S_{T}\Rightarrow\mathbb{P}\circ\left(s_{0}+W_{T}\sigma-Y\right).

Since HH is bounded and H[Tδ,T]×d0H_{[T-\delta,T]\times\mathbb{R}^{d}}\equiv 0 then the term Kt,sH(s,Ws)K_{t,s}H(s,W_{s}) is uniformly bounded and so we have the growth bound

(18) supn𝔼n[sup0tTκtnp]<,p>0.\sup_{n\in\mathbb{N}}\mathbb{E}_{\mathbb{Q}^{n}}\left[\sup_{0\leq t\leq T}||\kappa^{n}_{t}||^{p}\right]<\infty,\ \ \forall p>0.

In particular, nST\mathbb{Q}^{n}\circ S_{T}, nn\in\mathbb{N} are uniformly integrable. Thus, (17) implies (ff has a linear growth)

(19) limn𝔼n[f(ST)Φ0,STs0]=𝔼[f(s0+WTσY)+Φ0,Y].\lim_{n\rightarrow\infty}\mathbb{E}_{\mathbb{Q}^{n}}\left[f\left(S_{T}\right)-\left\langle\Phi_{0},S_{T}-s_{0}\right\rangle\right]=\mathbb{E}_{\mathbb{P}}\left[f\left(s_{0}+W_{T}\sigma-Y\right)+\left\langle\Phi_{0},Y\right\rangle\right].

Next, let HH{{}^{\prime}} be the transpose of HH. From the Fubini theorem, the Itô isometry, (18) and the equality 𝔼n[κtn]=0\mathbb{E}_{\mathbb{Q}^{n}}[\kappa^{n}_{t}]=0 it follows that

ΛAlimn𝔼n[log(dnd)]=Λ2Alimn𝔼n[0Tθtn2𝑑t]\displaystyle\frac{\Lambda}{A}\lim_{n\rightarrow\infty}\mathbb{E}_{\mathbb{Q}^{n}}\left[\log\left(\frac{d\mathbb{Q}^{n}}{d\mathbb{P}}\right)\right]=\frac{\Lambda}{2A}\lim_{n\rightarrow\infty}\mathbb{E}_{\mathbb{Q}^{n}}\left[\int_{0}^{T}||\theta^{n}_{t}||^{2}dt\right]
=Λ2A𝔼[0TatΛ+κtΛ2𝑑t]c1Λ+12Λ0TyKt,0Λ2𝑑t\displaystyle=\frac{\Lambda}{2A}\mathbb{E}_{\mathbb{P}}\left[\int_{0}^{T}||a^{\Lambda}_{t}+\kappa^{\Lambda}_{t}||^{2}dt\right]\leq c_{1}\Lambda+\frac{1}{2\Lambda}\int_{0}^{T}||yK^{\Lambda}_{t,0}||^{2}dt
(20) +12Λ𝔼[tr(0TH(s,Ws)(sT(Kt,sΛ)2𝑑t)H(s,Ws)𝑑s)]\displaystyle+\frac{1}{2\Lambda}\mathbb{E}_{\mathbb{P}}\left[tr\left(\int_{0}^{T}H^{\prime}(s,W_{s})\left(\int_{s}^{T}(K^{\Lambda}_{t,s})^{2}dt\right)H(s,W_{s})ds\right)\right]

for some constant c1>0c_{1}>0 which does not depend on Λ\Lambda.

Finally, we estimate the last term in the right hand side of (3.1). From (1)

St𝔼n[ST|t]\displaystyle S_{t}-\mathbb{E}_{\mathbb{Q}^{n}}\left[S_{T}\left|\right.\mathcal{F}_{t}\right]
=yGtΛ+𝔼n[tTκvnσ𝑑v|t]𝔼n[tT𝕀κvnnκvnσ𝑑v|t],t[0,T]\displaystyle=yG^{\Lambda}_{t}+\mathbb{E}_{\mathbb{Q}^{n}}\left[\int_{t}^{T}\kappa^{n}_{v}\sigma dv\left|\right.\mathcal{F}_{t}\right]-\mathbb{E}_{\mathbb{Q}^{n}}\left[\int_{t}^{T}\mathbb{I}_{||\kappa^{n}_{v}||\geq n}\kappa^{n}_{v}\sigma dv\left|\right.\mathcal{F}_{t}\right],\ \ t\in[0,T]

for

GtΛ:=AΛσtTKv,0Λ𝑑v=sinh(A(Tt)σ/Λ)(sinh(ATσ/Λ))1.G^{\Lambda}_{t}:=\frac{\sqrt{A}}{\Lambda}\sigma\int_{t}^{T}K^{\Lambda}_{v,0}dv=\sinh\left(\sqrt{A}(T-t)\sigma/\Lambda\right)\left(\sinh\left(\sqrt{A}T\sigma/\Lambda\right)\right)^{-1}.

By combining the Doob inequality for the martingales

(𝔼n[0T𝕀κvnnκvn𝑑v|t])t[0,T],n\left(\mathbb{E}_{\mathbb{Q}^{n}}\left[\int_{0}^{T}\mathbb{I}_{||\kappa^{n}_{v}||\geq n}||\kappa^{n}_{v}||dv\left|\right.\mathcal{F}_{t}\right]\right)_{t\in[0,T]},\ \ n\in\mathbb{N}

and (18) it follows that

(𝔼n[tT𝕀κvnnκvnσ𝑑v|t])t[0,T]0inL2(dtn).\left(\mathbb{E}_{\mathbb{Q}^{n}}\left[\int_{t}^{T}\mathbb{I}_{||\kappa^{n}_{v}||\geq n}\kappa^{n}_{v}\sigma dv\left|\right.\mathcal{F}_{t}\right]\right)_{t\in[0,T]}\rightarrow 0\ \ \mbox{in}\ \ L^{2}(dt\otimes\mathbb{Q}^{n}).

This together with the equality 𝔼n[κtn]=0\mathbb{E}_{\mathbb{Q}^{n}}[\kappa^{n}_{t}]=0 yields

12Λlimn𝔼n[0T||St𝔼n[ST|t]||2dt]\displaystyle\frac{1}{2\Lambda}\lim_{n\rightarrow\infty}\mathbb{E}_{\mathbb{Q}^{n}}\left[\int_{0}^{T}\left|\left|S_{t}-\mathbb{E}_{\mathbb{Q}^{n}}[S_{T}\left|\right.\mathcal{F}_{t}]\right|\right|^{2}dt\right]
=12Λ0T||yGtΛ||2dt+12Λlimn𝔼n[0T||𝔼n[tTκvnσdv|t]||2dt]\displaystyle=\frac{1}{2\Lambda}\int_{0}^{T}\left|\left|yG^{\Lambda}_{t}\right|\right|^{2}dt+\frac{1}{2\Lambda}\lim_{n\rightarrow\infty}\mathbb{E}_{\mathbb{Q}^{n}}\left[\int_{0}^{T}\left|\left|\mathbb{E}_{\mathbb{Q}^{n}}\left[\int_{t}^{T}\kappa^{n}_{v}\sigma dv\left|\right.\mathcal{F}_{t}\right]\right|\right|^{2}dt\right]
(21) =12Λ0T||yGtΛ||2dt+12Λ𝔼[0T||𝔼[tTκvΛσdv|t]||2dt].\displaystyle=\frac{1}{2\Lambda}\int_{0}^{T}\left|\left|yG^{\Lambda}_{t}\right|\right|^{2}dt+\frac{1}{2\Lambda}\mathbb{E}_{\mathbb{P}}\left[\int_{0}^{T}\left|\left|\mathbb{E}_{\mathbb{P}}\left[\int_{t}^{T}\kappa^{\Lambda}_{v}\sigma dv\left|\right.\mathcal{F}_{t}\right]\right|\right|^{2}dt\right].

From the Fubini theorem

𝔼[tTκvΛ𝑑v|t]\displaystyle\mathbb{E}_{\mathbb{P}}\left[\int_{t}^{T}\kappa^{\Lambda}_{v}dv\left|\mathcal{F}_{t}\right.\right]
=AΛ𝔼[0T(tsTKv,sΛ𝑑v)H(s,Ws)𝑑Ws|t]\displaystyle=\frac{\sqrt{A}}{\Lambda}\mathbb{E}_{\mathbb{P}}\left[\int_{0}^{T}\left(\int_{t\vee s}^{T}K^{\Lambda}_{v,s}dv\right)H(s,W_{s})\cdot dW_{s}\left|\right.\mathcal{F}_{t}\right]
=AΛ0t(tTKv,sΛ𝑑v)H(s,Ws)𝑑Ws.\displaystyle=\frac{\sqrt{A}}{\Lambda}\int_{0}^{t}\left(\int_{t}^{T}K^{\Lambda}_{v,s}dv\right)H(s,W_{s})\cdot dW_{s}.

Hence, the Itô isometry yields

12Λ𝔼[0T||𝔼[tTκvΛσdv|t]||2dt]\displaystyle\frac{1}{2\Lambda}\mathbb{E}_{\mathbb{P}}\left[\int_{0}^{T}\left|\left|\mathbb{E}_{\mathbb{P}}\left[\int_{t}^{T}\kappa^{\Lambda}_{v}\sigma dv\left|\mathcal{F}_{t}\right.\right]\right|\right|^{2}dt\right]
(22) =12Λ𝔼[tr(0TH(s,Ws)(sT(Lt,sΛ)2𝑑t)H(s,Ws)𝑑s)]\displaystyle=\frac{1}{2\Lambda}\mathbb{E}_{\mathbb{P}}\left[tr\left(\int_{0}^{T}H^{\prime}(s,W_{s})\left(\int_{s}^{T}\left(L^{\Lambda}_{t,s}\right)^{2}dt\right)H(s,W_{s})ds\right)\right]

where

Lt,sΛ:=AΛσtTKv,sΛ𝑑v=sinh(A(Tt)σ/Λ)(sinh(A(Ts)σ/Λ))1.L^{\Lambda}_{t,s}:=\frac{\sqrt{A}}{\Lambda}\sigma\int_{t}^{T}K^{\Lambda}_{v,s}dv=\sinh\left(\sqrt{A}(T-t)\sigma/\Lambda\right)\left(\sinh\left(\sqrt{A}(T-s)\sigma/\Lambda\right)\right)^{-1}.

Step III: In this step we take Λ0\Lambda\downarrow 0. First, from the Itô isometry

(23) 𝔼[Y,Yσ1]=y,yσ1+𝔼[tr(0TH(s,Ws)σ1H(s,Ws)𝑑s)].\mathbb{E}_{\mathbb{P}}\left[\left\langle Y,Y\sigma^{-1}\right\rangle\right]=\langle y,y\sigma^{-1}\rangle+\mathbb{E}_{\mathbb{P}}\left[tr\left(\int_{0}^{T}H^{\prime}(s,W_{s})\sigma^{-1}H(s,W_{s})ds\right)\right].

Next, since σ\sigma is positive definite, then for any ϵ>0\epsilon>0 we have the uniform convergence

limΛ0(sup0tTϵ12ΛsT(Kt,sΛ)2𝑑tσ14A)\displaystyle\lim_{\Lambda\downarrow 0}\left(\sup_{0\leq t\leq T-\epsilon}\left|\left|\frac{1}{2\Lambda}\int_{s}^{T}\left(K^{\Lambda}_{t,s}\right)^{2}dt-\frac{\sigma^{-1}}{4\sqrt{A}}\right|\right|\right)
=limΛ0(sup0tTϵ12ΛsT(Lt,sΛ)2𝑑tσ14A)=0.\displaystyle=\lim_{\Lambda\downarrow 0}\left(\sup_{0\leq t\leq T-\epsilon}\left|\left|\frac{1}{2\Lambda}\int_{s}^{T}\left(L^{\Lambda}_{t,s}\right)^{2}dt-\frac{\sigma^{-1}}{4\sqrt{A}}\right|\right|\right)=0.

Hence, by combining (3.1) and (19)–(23) (recall that H[Tδ,T]×d0H_{[T-\delta,T]\times\mathbb{R}^{d}}\equiv 0) we obtain

liminfΛ0c(Λ,A/Λ,Φ0,f)\displaystyle\lim\inf_{\Lambda\downarrow 0}c\left(\Lambda,A/\Lambda,\Phi_{0},f\right)
liminfΛ0limn𝔼n[f(ST)Φ0,STs01αlog(dnd)\displaystyle\geq\lim\inf_{\Lambda\downarrow 0}\lim_{n\rightarrow\infty}\mathbb{E}_{\mathbb{Q}^{n}}\bigg{[}f\left(S_{T}\right)-\langle\Phi_{0},S_{T}-s_{0}\rangle-\frac{1}{\alpha}\log\left(\frac{d\mathbb{Q}^{n}}{d\mathbb{P}}\right)
12Λ0T||St𝔼n(ST|t)||2dt]\displaystyle-\frac{1}{2\Lambda}\int_{0}^{T}\left|\left|S_{t}-\mathbb{E}_{\mathbb{Q}^{n}}\left(S_{T}|\mathcal{F}_{t}\right)\right|\right|^{2}dt\bigg{]}
𝔼[f(s0+WTσY)+Φ0,Y12AY,Yσ1]\displaystyle\geq\mathbb{E}_{\mathbb{P}}\left[f\left(s_{0}+W_{T}\sigma-Y\right)+\langle\Phi_{0},Y\rangle-\frac{1}{2\sqrt{A}}\left\langle Y,Y\sigma^{-1}\right\rangle\right]

as required.

We now have all the pieces in place that we need for the completion of the proof of the inequality """\geq" in (8).

Proof 3.5.

Recall the definition of gAg^{A} given by (4)(\ref{def1}). Choose ϵ>0\epsilon>0. From the Lipschitz continuity of ff it follows that there exists a finitely valued (and hence bounded and measurable) function ζ:dd\zeta:\mathbb{R}^{d}\rightarrow\mathbb{R}^{d} such that

(24) gA(x)<ϵ+f(xζ(x))ζ(x)σ1,ζ(x)2A,xd.g^{A}(x)<\epsilon+f\left(x-\zeta(x)\right)-\frac{\langle\zeta(x)\sigma^{-1},\zeta(x)\rangle}{2\sqrt{A}},\ \ \ \forall x\in\mathbb{R}^{d}.

By applying Proposition 3.3 for the random variable

Y:=ζ(s0AΦ0+WTσ)+AΦ0σY:=\zeta(s_{0}-\sqrt{A}\Phi_{0}+W_{T}\sigma)+\sqrt{A}\Phi_{0}\sigma

and (24) for x:=s0AΦ0σ+WTσx:=s_{0}-\sqrt{A}\Phi_{0}\sigma+W_{T}\sigma we obtain

liminfΛ0c(Λ,A/Λ,Φ0,f)\displaystyle\lim\inf_{\Lambda\downarrow 0}c(\Lambda,A/\Lambda,\Phi_{0},f)
𝔼[gA(s0AΦ0σ+WTσ)+Yσ1AΦ0,YAΦ0σ2A\displaystyle\geq\mathbb{E}_{\mathbb{P}}\bigg{[}g^{A}\left(s_{0}-\sqrt{A}\Phi_{0}\sigma+W_{T}\sigma\right)+\frac{\left\langle Y\sigma^{-1}-\sqrt{A}\Phi_{0},Y-\sqrt{A}\Phi_{0}\sigma\right\rangle}{2\sqrt{A}}
+Φ0,Y12AY,Yσ1]ϵ\displaystyle+\langle\Phi_{0},Y\rangle-\frac{1}{2\sqrt{A}}\left\langle Y,Y\sigma^{-1}\right\rangle\bigg{]}-\epsilon
=uA(0,s0AΦ0σ)+AΦ0σ,Φ02ϵ\displaystyle=u^{A}\left(0,s_{0}-\sqrt{A}\Phi_{0}\sigma\right)+\frac{\sqrt{A}\langle\Phi_{0}\sigma,\Phi_{0}\rangle}{2}-\epsilon

where the equality follows from the definition of uAu^{A} given by (5). By taking ϵ0\epsilon\downarrow 0 we complete the proof.

4 Proof of the Upper Bound

In order to complete the proof of Theorem 2.1 it remains to establish the following result.

Proposition 4.1.

Fix A>0A>0 and recall the trading strategies ΦA,Λ\Phi^{A,\Lambda}, Λ>0\Lambda>0 which are given by (7). Then,

limΛ0ΛAlog(𝔼[exp(AΛ(f(ST)VTΦ0,ϕA,Λ))])\displaystyle\lim_{\Lambda\downarrow 0}\frac{\Lambda}{A}\log\left(\mathbb{E}_{\mathbb{P}}\left[\exp\left(\frac{A}{\Lambda}\left(f(S_{T})-V^{\Phi_{0},\phi^{A,\Lambda}}_{T}\right)\right)\right]\right)
uA(0,s0AΦ0σ)+AΦ0σ,Φ02.\displaystyle\leq u^{A}\left(0,s_{0}-\sqrt{A}\Phi_{0}\sigma\right)+\frac{\sqrt{A}\langle\Phi_{0}\sigma,\Phi_{0}\rangle}{2}.

Proof 4.2.

Introduce the dd–dimensional process

ΘtA:=DxuA(t,StAΦtσ),t[0,T].\Theta^{A}_{t}:=D_{x}u^{A}\left(t,S_{t}-\sqrt{A}\Phi_{t}\sigma\right),\ \ t\in[0,T].

From the ODE (7) it follows that for any Λ>0\Lambda>0

ΦtA,Λ=Φ0exp(Atσ/Λ)+AΛ0tΘvΛσexp(A(vt)σ/Λ)𝑑v,t[0,T].\Phi^{A,\Lambda}_{t}=\Phi_{0}\cdot\exp\left(-\sqrt{A}t\sigma/\Lambda\right)+\frac{\sqrt{A}}{\Lambda}\int_{0}^{t}\Theta^{\Lambda}_{v}\sigma\exp\left(\sqrt{A}(v-t)\sigma/\Lambda\right)dv,\ \ t\in[0,T].

Since σ\sigma is positive definite and DxuAD_{x}u^{A} is uniformly bonded (ff is Lipschitz), then there exists a constant C>0C>0 such that

(25) sup0tTΦtA,ΛC,Λ>0.\sup_{0\leq t\leq T}||\Phi^{A,\Lambda}_{t}||\leq C,\ \ \forall\Lambda>0.

Next, for any Λ>0\Lambda>0 define the process MΛ=(MtΛ)t[0,T]M^{\Lambda}=(M^{\Lambda}_{t})_{t\in[0,T]} by

MtΛ:=exp(AΛ(uA(t,StAΦtA,Λσ)+AΦtA,Λσ,ΦtA,Λ2VtΦ0,ϕA,Λ)).M^{\Lambda}_{t}:=\exp\left(\frac{A}{\Lambda}\left(u^{A}\left(t,S_{t}-\sqrt{A}\Phi^{A,\Lambda}_{t}\sigma\right)+\frac{\sqrt{A}\langle\Phi^{A,\Lambda}_{t}\sigma,\Phi^{A,\Lambda}_{t}\rangle}{2}-V^{\Phi_{0},\phi^{A,\Lambda}}_{t}\right)\right).

From the Itô formula, (2) and (6)

dMtΛMtΛ=AΛDxuA(t,StAΦtA,Λσ)ΦtA,Λ,dSt\displaystyle\frac{dM^{\Lambda}_{t}}{M^{\Lambda}_{t}}=\frac{A}{\Lambda}\left\langle D_{x}u^{A}\left(t,S_{t}-\sqrt{A}\Phi^{A,\Lambda}_{t}\sigma\right)-\Phi^{A,\Lambda}_{t},dS_{t}\right\rangle
+A22Λ2(DxuA(t,StAΦtA,Λσ)ΦtA,Λ)σ2dt\displaystyle+\frac{A^{2}}{2\Lambda^{2}}\left|\left|\left(D_{x}u^{A}\left(t,S_{t}-\sqrt{A}\Phi^{A,\Lambda}_{t}\sigma\right)-\Phi^{A,\Lambda}_{t}\right)\sigma\right|\right|^{2}dt
A3/2ΛϕtA,Λ,(DxuA(t,StAΦtA,Λσ)ΦtA,Λ)σΛ2AϕtA,Λdt\displaystyle-\frac{A^{3/2}}{\Lambda}\left\langle\phi^{A,\Lambda}_{t},\left(D_{x}u^{A}\left(t,S_{t}-\sqrt{A}\Phi^{A,\Lambda}_{t}\sigma\right)-\Phi^{A,\Lambda}_{t}\right)\sigma-\frac{\Lambda}{2\sqrt{A}}\phi^{A,\Lambda}_{t}\right\rangle dt
=AϕtA,Λσ1,dSt\displaystyle=\sqrt{A}\left\langle\phi^{A,\Lambda}_{t}\sigma^{-1},dS_{t}\right\rangle

where the last equality follows from (7). Hence, from (1)

exp(AΦtA,ΛΦ0A,Λ,μσ1)MtΛ,t[0,T]\exp\left(-\sqrt{A}\left\langle\Phi^{A,\Lambda}_{t}-\Phi^{A,\Lambda}_{0},\mu\sigma^{-1}\right\rangle\right)M^{\Lambda}_{t},\ \ t\in[0,T]

is a local–martingale, and so from the obvious inequality MΛ>0M^{\Lambda}>0 we conclude that this process is a super–martingale.

Finally, (4) yields that f(x)gA(xyσ)+yσ,y2Af(x)\leq g^{A}\left(x-y\sigma\right)+\frac{\langle y\sigma,y\rangle}{2\sqrt{A}} for all x,ydx,y\in\mathbb{R}^{d}. This together with the growth bound (25) and the above super–martingale property gives (observe that uA(T,)=gA()u^{A}(T,\cdot)=g^{A}(\cdot)) that for any Λ>0\Lambda>0

ΛAlog(𝔼[exp(AΛ(f(ST)VTΦ0,ϕΛ))])\displaystyle\frac{\Lambda}{A}\log\left(\mathbb{E}_{\mathbb{P}}\left[\exp\left(\frac{A}{\Lambda}\left(f(S_{T})-V^{\Phi_{0},\phi^{\Lambda}}_{T}\right)\right)\right]\right)
ΛAlog(𝔼[MTΛ])ΛAlog(M0Λ)+ΛA2CTμσ1\displaystyle\leq\frac{\Lambda}{A}\log\left(\mathbb{E}_{\mathbb{P}}[M^{\Lambda}_{T}]\right)\leq\frac{\Lambda}{A}\log\left(M^{\Lambda}_{0}\right)+\frac{\Lambda}{\sqrt{A}}2CT||\mu\sigma^{-1}||
=uA(0,s0AΦ0σ)+AΦ0σ,Φ02+ΛA2CTμσ1\displaystyle=u^{A}\left(0,s_{0}-\sqrt{A}\Phi_{0}\sigma\right)+\frac{\sqrt{A}\langle\Phi_{0}\sigma,\Phi_{0}\rangle}{2}+\frac{\Lambda}{\sqrt{A}}2CT||\mu\sigma^{-1}||

and the result follows by taking Λ0\Lambda\downarrow 0.

Acknowledgements

The authors thank the AE and the anonymous reviewers for their valuable reports and comments which helped to improve the quality of this paper.

References

  • [1] R. Almgren and N. Chriss, Optimal execution of portfolio transactions, Journal of Risk, 3, 5–39, (2001).
  • [2] A. Alfonsi, A. Fruth and A. Schied, Optimal execution strategies in limit order books with general shape functions, Quantitative Finance, 2, 143–157, (2010).
  • [3] M. Anthropelos, S. Robertson and K. Spliopoulus, The pricing of contingent claims and optimal positions in asymptotically complete markets, Annals of Applied Probability, 27, 1178–1830, (2017).
  • [4] F. Black, Noise, Journal of Finance, 41, 529–543, (1986).
  • [5] E. Bayraktar, T. Caye, and I. Ekren, Asymptotics for Small Nonlinear Price Impact: a PDE Approach to the Multidimensional Case, Mathematical Finance, 31, 36–108, (2021).
  • [6] P. Bank, Y. Dolinsky and M. Rásonyi, What if we knew what the future brings? https://arxiv.org/abs/2108.04291, (2021).
  • [7] B. Bouchard, G. Loeper and Y.Zou, Almost-sure hedging with permanent price impact, Finance and Stochastics, 20, 741–771, (2016).
  • [8] G. Barles and H.M. Soner, Option pricing with transaction costs and a nonlinear Black–Scholes equation, Finance and Stochastics, 2, 369–397, (1998).
  • [9] B. Bouchard and X. Tan, Understanding the dual formulation for the hedging of path-dependent options with price impact, to appear in Annals of Applied Probability. https://arxiv.org/pdf/1912.03946, (2020).
  • [10] P. Bank and M. Voß, Optimal Investment with Transient Price Impact, SIAM Journal on Financial Mathematics, 10, 723–768, (2019).
  • [11] R. Carmona, Indifference pricing: theory and applications, Princeton University Press, series in Financial Engineering, (2009).
  • [12] T. Caye, M. Herdegen and J. Muhle-Karbe, Trading with Small Nonlinear Price Impact, Annals of Applied Probability, 30, 706–746, (2020).
  • [13] A. Fruth, T. Schöneborn and M. Urusov, Optimal trade execution in order books with stochastic liquidity, Mathematical Finance, 29, 507–541, (2019).
  • [14] P. Guasoni and M. Rásonyi, Hedging, arbitrage and optimality under superlinear friction, Annals of Applied Probability, 25, 2066–2095, (2015).
  • [15] P. Guasoni and M. Weber, Nonlinear price impact and portfolio choice, Mathematical Finance, 30, 341–376, (2020).
  • [16] L. Moreau, J. Muhle-Karbe and H.M. Soner, Trading with Small Price Impact, Mathematical Finance, 27, 350-400, (2017).
  • [17] S. Nadtochiy, A simple microstructural explanation of the concavity of price impact, to appear in Mathematical Finance. https://arxiv.org/pdf/2001.01860, (2020).
  • [18] A. Schied and T. Zhang, A Market Impact Game Under Transient Price Impact, Mathematics of Operations Research, 44, 102–121, (2019).
  • [19] W. Walter, Ordinary Differential Equations, Springer-Verlag New-York, (1998).