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

Reducing Obizhaeva-Wang type trade execution problems to LQ stochastic control problems

Julia Ackermann Department of Mathematics & Informatics, University of Wuppertal, Gaußstr. 20, 42119 Wuppertal, Germany. Email: jackermann@uni-wuppertal.de, Phone: +49 (0)202 4395238.    Thomas Kruse Department of Mathematics & Informatics, University of Wuppertal, Gaußstr. 20, 42119 Wuppertal, Germany. Email: tkruse@uni-wuppertal.de, Phone: +49 (0)202 4395239.    Mikhail Urusov Faculty of Mathematics, University of Duisburg-Essen, Thea-Leymann-Str. 9, 45127 Essen, Germany. Email: mikhail.urusov@uni-due.de, Phone: +49 (0)201 1837428.
Abstract

We start with a stochastic control problem where the control process is of finite variation (possibly with jumps) and acts as integrator both in the state dynamics and in the target functional. Problems of such type arise in the stream of literature on optimal trade execution pioneered by Obizhaeva and Wang (models with finite resilience). We consider a general framework where the price impact and the resilience are stochastic processes. Both are allowed to have diffusive components. First we continuously extend the problem from processes of finite variation to progressively measurable processes. Then we reduce the extended problem to a linear quadratic (LQ) stochastic control problem. Using the well developed theory on LQ problems we describe the solution to the obtained LQ one and trace it back up to the solution to the (extended) initial trade execution problem. Finally, we illustrate our results by several examples. Among other things the examples show the Obizhaeva-Wang model with random (terminal and moving) targets, the necessity to extend the initial trade execution problem to a reasonably large class of progressively measurable processes (even going beyond semimartingales) and the effects of diffusive components in the price impact process and/or in the resilience process.

Keywords: optimal trade execution; stochastic price impact; stochastic resilience; finite variation stochastic control; continuous extension of cost functional; progressively measurable execution strategy; linear quadratic stochastic control; backward stochastic differential equation.

2020 MSC: Primary: 91G10; 93E20; 60H10. Secondary: 60G99.

Introduction

In the literature on optimal trade execution in illiquid financial markets there arise stochastic control problems where the control is a process of finite variation (possibly with jumps) that acts as integrator both in the state dynamics and in the target functional. For brevity, we use the term finite variation stochastic control for such problems.111Notice that the class of finite variation stochastic control problems contains the class of singular stochastic control problems. In contrast, for control problems where the state is driven by a controlled stochastic differential equation (SDE) and the control acts as one of the arguments in that SDE and as one of the arguments in the integrand of the target functional, we use the term standard stochastic control problems.

In this article we present a general solution approach to finite variation stochastic control problems that arise in the literature on optimal trade execution. We set up a finite variation stochastic control problem of the type of the one in Obizhaeva and Wang [37] and its extensions like, e.g., Alfonsi and Acevedo [4], Bank and Fruth [13], Fruth et al. [24] and [25]. We then show how it can be transformed into a standard linear quadratic (LQ) stochastic control problem which can be solved with the help of state-of-the-art techniques from stochastic optimal control theory. In the introduction we first describe the finite variation stochastic control problem and showcase its usage in finance, before presenting our solution approach, summarizing our main contributions and embedding our paper into the literature.

Finite variation stochastic control problem: As a starting point we consider in this paper the following stochastic control problem. Let T>0T>0 and let (Ω,T,(t)t[0,T],P)(\Omega,\mathcal{F}_{T},(\mathcal{F}_{t})_{t\in[0,T]},P) be a filtered probability space satisfying the usual conditions. Let ξ\xi be an T\mathcal{F}_{T}-measurable random variable and let ζ=(ζs)s[0,T]\zeta=(\zeta_{s})_{s\in[0,T]} be a progressively measurable process both satisfying suitable integrability assumptions (see (5) below). Further, let λ=(λs)s[0,T]\lambda=(\lambda_{s})_{s\in[0,T]} be a bounded progressively measurable process. Let γ=(γs)s[0,T]\gamma=(\gamma_{s})_{s\in[0,T]} be a positive Itô process driven by some Brownian motion and R=(Rs)s[0,T]R=(R_{s})_{s\in[0,T]} an Itô process driven by a (stochastically) correlated Brownian motion (see (3) and (4) below). Throughout the introduction we fix t[0,T]t\in[0,T], x,dx,d\in\mathbb{R} and denote by 𝒜tfv(x,d)\mathcal{A}^{fv}_{t}(x,d) the set of all adapted, càdlàg, finite variation processes X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} satisfying Xt=xX_{t-}=x, XT=ξX_{T}=\xi, and appropriate integrability assumptions (see (A1)–(A3) below). To each X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d) we associate a process DX=(DsX)s[t,T]D^{X}=(D^{X}_{s})_{s\in[t-,T]} satisfying

dDsX=DsXdRs+γsdXs,s[t,T],DtX=d.dD^{X}_{s}=-D^{X}_{s}dR_{s}+\gamma_{s}dX_{s},\quad s\in[t,T],\quad D^{X}_{t-}=d. (1)

We consider the finite variation stochastic control problem of minimizing the cost functional

Jtfv(x,d,X)=Et[[t,T](DsX+12ΔXsγs)𝑑Xs+tTλsγs(Xsζs)2𝑑s]J^{fv}_{t}(x,d,X)=E_{t}\left[\int_{[t,T]}\left(D^{X}_{s-}+\frac{1}{2}\Delta X_{s}\gamma_{s}\right)dX_{s}+\int_{t}^{T}\lambda_{s}\gamma_{s}\left(X_{s}-\zeta_{s}\right)^{2}ds\right] (2)

over X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d), where Et[]E_{t}[\cdot] is a shorthand notation for E[|t]E[\cdot|\mathcal{F}_{t}].

Financial interpretation: Stochastic control problems with cost functional of the form (2) or a special case thereof play a central role in the scientific literature on optimal trade execution problems (see the literature discussion below). Consider an institutional investor who holds immediately prior to time t[0,T]t\in[0,T] a position xx\in\mathbb{R} (x>0x>0 meaning a long position of xx shares of a stock and x<0x<0 a short position of x-x shares) of a certain financial asset. The investor trades the asset during the period [t,T][t,T] in such a way that at each time s[t,T]s\in[t-,T] the position is given by the value XsX_{s} of the adapted, càdlàg, finite variation process X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} (satisfying Xt=xX_{t-}=x). More precisely, XsX_{s-} represents the position immediately prior to the trade at time ss, while XsX_{s} is the position immediately after that trade. The investor’s goal is to reach the target position

XT=ξX_{T}=\xi

during the course of the trading period [t,T][t,T]. Note that we allow ξ\xi to be random to incorporate the possibility that the target position is not known at the beginning of trading but only revealed at terminal time TT. Such situations may for example be faced by airline companies buying on forward markets the kerosene they need in TT months. Their precise demand for kerosene at that future time depends on several factors, such as ticket sales and flight schedules, that are not known today but only gradually learned.

We assume that the market the investor trades in is illiquid, implying that the investor’s trades impact the asset price. To model this effect, we assume (as is typically done in the literature on optimal trade execution) an additive impact on the price. This means that the realized price at which the investor trades at time r[t,T]r\in[t,T] consists of an unaffected price Sr0S^{0}_{r} plus a deviation DrXD^{X}_{r} that is caused by the investor’s trades during [t,r][t,r]. We assume that the unaffected price process S0=(Sr0)r[0,T]S^{0}=(S^{0}_{r})_{r\in[0,T]} is a càdlàg martingale satisfying appropriate integrability conditions. Then integration by parts and the martingale property of S0S^{0} ensure that expected trading costs due to S0S^{0} are given by

Et[[t,T]Sr0𝑑Xr]=Et[ξST0]xSt0.E_{t}\left[\int_{[t,T]}S^{0}_{r}dX_{r}\right]=E_{t}\left[\xi S_{T}^{0}\right]-xS_{t}^{0}.

Thus, these costs do not depend on the investor’s trading strategy XX and are therefore neglected in the sequel (we refer to Remark 2.2 in [1] for a more detailed discussion in the case ξ=0\xi=0). The deviation process DXD^{X} associated to XX is given by (1). Informally speaking, we see from (1) that a trade of size dXsdX_{s} at time s[t,T]s\in[t,T] impacts DXD^{X} by γsdXs\gamma_{s}dX_{s}. So, the factor γs\gamma_{s} determines how strongly the price reacts to trades, and the process γ\gamma is therefore called the price impact process. In particular, the fact that γ\gamma is nonnegative entails that a buy trade dXs>0dX_{s}>0 leads to higher prices whereas a sell trade dXs<0dX_{s}<0 leads to smaller prices. The second component DsXdRs-D^{X}_{s}dR_{s} in the dynamics (1) describes the behavior of DXD^{X} when the investor is not trading. Typically, it is assumed that RR is an increasing process such that in the absence of trades DXD^{X} is reverting to 0 with relative rate dRsdR_{s}. Therefore, RR is called the resilience process. We refer to [3] for a discussion of the effects of “negative” resilience, where RR might also be decreasing. We highlight that in the present paper we allow RR to have a diffusive part. In summary, we note that the deviation prior to a trade of the investor at time s[t,T]s\in[t,T] is given by DsXD^{X}_{s-} whereas it is equal to DsX=DsX+γsΔXsD^{X}_{s}=D^{X}_{s-}+\gamma_{s}\Delta X_{s} afterwards. We take the mean DsX+12γsΔXsD^{X}_{s-}+\frac{1}{2}\gamma_{s}\Delta X_{s} of these two values as the realized price per unit so that the investor’s overall trading costs due to DXD^{X} amount to [t,T](DsX+12γsΔXs)𝑑Xs\int_{[t,T]}\left(D^{X}_{s-}+\frac{1}{2}\gamma_{s}\Delta X_{s}\right)dX_{s}. This describes the first integral on the right-hand side of (2). Under the assumption that λ\lambda is nonnegative, the second integral tTλsγs(Xsζs)2𝑑s\int_{t}^{T}\lambda_{s}\gamma_{s}\left(X_{s}-\zeta_{s}\right)^{2}ds can be understood as a risk term that penalizes any deviation of the position XX from the moving target ζ\zeta in a quadratic way222The parametrization λsγs\lambda_{s}\gamma_{s}, s[0,T]s\in[0,T], for the weight is chosen out of mathematical convenience since it makes some of the following assumptions and results shorter to state. Likewise, one can use λ~s\tilde{\lambda}_{s}, s[0,T]s\in[0,T], as a weight and replace λ\lambda by λ~/γ\tilde{\lambda}/\gamma in the subsequent assumptions and results.. A possible and natural choice would be ζs=Es[ξ]\zeta_{s}=E_{s}[\xi], s[0,T]s\in[0,T], so that the risk term ensures that any optimal strategy XX does not deviate too much from the (expected) target position ξ\xi in the course of the trading period.

Solution approach: The overarching goal of this paper is to show that the finite variation stochastic control problem (2) is equivalent to a standard LQ stochastic control problem (see Corollary˜2.3 and Corollary˜2.4 below). The derivation of this result is based on the following insights. The first observation is that, in general, the functional (2) does not admit a minimizer in 𝒜tfv(x,d)\mathcal{A}^{fv}_{t}(x,d) (see Section˜4.3 below for a specific example). In [1] the functional (2) was extended to a set of càdlàg semimartingales XX and it was shown that its minimum is attained in this set of semimartingales if and only if a certain process that is derived from the solution of an associated backward stochastic differential equation (BSDE) can be represented by a càdlàg semimartingale (see Theorem 2.4 in [1]). In this work we go even a step further and extend the functional (2) to the set 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) of progressively measurable processes X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} satisfying appropriate integrability conditions (see (A1) below) and the boundary conditions Xt=xX_{t-}=x and XT=ξX_{T}=\xi. To do so, we first derive alternative representations of the first integral inside the expectation in (2) and the deviation in (1) that do not involve X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d) as an integrator (see ˜1.3). It follows that the resulting alternative representation of JfvJ^{fv} (see ˜1.4) is not only well-defined on 𝒜tfv(x,d)\mathcal{A}^{fv}_{t}(x,d) but even on 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d), and we denote this extended functional by JpmJ^{pm} (see Section˜1.3). We next introduce a metric on 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) and prove that JpmJ^{pm} is the unique continuous extension of JfvJ^{fv} from 𝒜tfv(x,d)\mathcal{A}^{fv}_{t}(x,d) to 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) (see ˜1.7). In particular, it follows that the infimum of JfvJ^{fv} over 𝒜tfv(x,d)\mathcal{A}^{fv}_{t}(x,d) and the infimum of JpmJ^{pm} over 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) coincide.

Next, for a given X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) we identify the process H¯sX=γs12DsXγs12Xs\overline{H}^{X}_{s}=\gamma_{s}^{-\frac{1}{2}}D^{X}_{s}-\gamma_{s}^{\frac{1}{2}}X_{s}, s[t,T]s\in[t,T], as a useful tool in our analysis. Despite XX and DXD^{X} having discontinuous paths in general, the process H¯X\overline{H}^{X}, which we call the scaled hidden deviation process, is always continuous. Moreover, we show that H¯X\overline{H}^{X} can be expressed in feedback form as an Itô process with coefficients that are linear in γ12DX\gamma^{-\frac{1}{2}}D^{X} and H¯X\overline{H}^{X} (see Lemma˜1.6). Subsequently, we reinterpret the process γ12DX\gamma^{-\frac{1}{2}}D^{X} as a control process uu and H¯X\overline{H}^{X} as the associated state process. Since the cost functional JpmJ^{pm} is quadratic in H¯X\overline{H}^{X} and u=γ12DXu=\gamma^{-\frac{1}{2}}D^{X}, we arrive at a standard LQ stochastic control problem (see (22) and (23)) whose minimal costs coincide with the infimum of JpmJ^{pm} over 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) (see Corollary˜2.3). Importantly, there is a one-to-one correspondence between square integrable controls uu for this standard problem and strategies X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d), which allows to recover the minimizer X𝒜tpm(x,d)X^{*}\in\mathcal{A}^{pm}_{t}(x,d) of JpmJ^{pm} from a minimizer uu^{*} of the standard problem and vice versa (see Corollary˜2.4).

We then solve the LQ stochastic control problem in (22) and (23) using techniques provided in the literature on stochastic optimal control theory. More precisely, we apply results from Kohlmann and Tang [34]333We moreover indicate in Remark 3.1 how we could alternatively use results from Sun et al. [40]. to provide conditions that guarantee that an optimal control uu^{*} exists (and is unique). This optimal control uu^{*} in the LQ problem is characterized by two BSDEs: one is a quadratic BSDE of Riccati type, the other one is linear, however, with unbounded coefficients (see ˜3.3). In Corollary˜3.4 we trace everything back and obtain a unique optimal execution strategy in the class of progressively measurable processes in a closed form (in terms of the solutions to the mentioned BSDEs).

Summary of our contributions: (a) The Obizhaeva-Wang type finite variation stochastic control problem (1)–(2) is continuously extended to the set 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) of appropriate progressively measurable processes XX.

(b) Problem (1)–(2) is rather general. In particular, it includes the following features:

  • Presence of random terminal and moving targets ξ\xi and (ζs)(\zeta_{s});

  • Price impact is a positive Itô process (γs)(\gamma_{s});

  • Resilience444To expand on this point, it is worth noting that in our current parametrization, only processes (Rs)(R_{s}) with dynamics dRs=ρsdsdR_{s}=\rho_{s}\,ds without a diffusive component were considered by now in the literature on optimal trade execution in Obizhaeva-Wang type models. Moreover, in most papers ρ\rho is assumed to be positive, that is, only the case of an increasing (Rs)(R_{s}) was extensively studied previously. is an Itô process (Rs)(R_{s}) acting as an integrator in (1).

(c) Via introducing the mentioned scaled hidden deviation process (H¯sX)(\overline{H}^{X}_{s}) and reinterpreting the process (γs12DsX)(\gamma_{s}^{-\frac{1}{2}}D^{X}_{s}) as a control in an (a priori, different) stochastic control problem, the extended to 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) problem is reduced to an explicitly solvable LQ stochastic control problem. Thus, a unique optimal execution strategy in 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) is obtained in a closed form (in terms of solutions to two BSDEs).

Literature discussion: Finite variation stochastic control problems arise in the group of literature on optimal trade execution in limit order books with finite resilience. The pioneering work555Posted 2005 on SSRN. Obizhaeva and Wang [37] models the price impact via a block-shaped limit order book, where the impact decays exponentially at a constant rate. This embeds into our model via the price impact process γ\gamma that is a positive constant and the resilience process (Rs)(R_{s}) given by Rs=ρsR_{s}=\rho s with some positive constant ρ>0\rho>0. Alfonsi et al. [5] study constrained portfolio liquidation in the Obizhaeva-Wang model. Subsequent works within this group of literature either extend this framework in different directions or suggest alternative frameworks with similar features. There is a subgroup of models which include more general limit order book shapes, see Alfonsi et al. [6], Alfonsi and Schied [7], Predoiu et al. [38]. Models in another subgroup extend the exponential decay of the price impact to general decay kernels, see Alfonsi et al. [8], Gatheral et al. [27]. Models with multiplicative price impact are analyzed in Becherer et al. [17, 18]. We mention that in [18], the (multiplicative) deviation is of Ornstein-Uhlenbeck type and incorporates a diffusion term (but this is different from our diffusion term that results from a diffusive part in the resilience RR). Superreplication and optimal investment in a block-shaped limit order book model with exponential resilience is discussed in Bank and Dolinsky [11, 12] and in Bank and Voß [16].

The present paper falls into the subgroup of the literature that studies time-dependent (possibly stochastic) price impact (γs)(\gamma_{s}) and resilience (Rs)(R_{s}) in generalized Obizhaeva-Wang models. In this connection we mention the works Alfonsi and Acevedo [4], Bank and Fruth [13], Fruth et al. [24], where deterministically varying price impact and resilience are considered. Fruth et al. [25] allow for stochastically varying price impact (resilience is still deterministic) and study the arising optimization problem over monotone strategies. Optimal strategies in a discrete-time model with stochastically varying resilience and constant price impact are derived in Siu et al. [39]. In Ackermann et al. [1, 3, 2] both price impact and resilience are stochastic. We now describe the differences from our present paper in more detail. In [2] optimal execution is studied in discrete time via dynamic programming. In [1] the framework is the closest to the one in this paper. Essentially, our current framework is the framework from [1] extended by a risk term with some moving target (ζs)(\zeta_{s}), a possibly non-zero (random) terminal target ξ\xi, and a larger class of resilience processes (in [1], as in many previous papers, (Rs)(R_{s}) is assumed to have the dynamics dRs=ρsdsdR_{s}=\rho_{s}\,ds, and (ρs)(\rho_{s}) is called resilience). In [3] the framework is similar to the one in [1], while the aim is to study qualitative effects of “negative” resilience (in the sense that ρs0\rho_{s}\leq 0 with (ρs)(\rho_{s}) as in the previous sentence). Now, to compare the approach in the present paper with the one in [1], we first recall that in [1] the finite variation stochastic control problem of the type (1)–(2) is extended to allow for càdlàg semimartingale trading strategies XX and the resulting optimal execution problem over semimartingales is studied. The approach in [1] is based on (1)–(2) (extended with some additional terms), but this does not work beyond semimartingales, as XX acts as integrator there. In contrast, our continuous extension needs to employ essentially different ideas since we want to consider the set 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) of progressively measurable strategies (in particular, beyond semimartingales). This extension is indeed necessary to get an optimizer (see the discussion in the end of Section˜4.3).

Especially with regard to our extension result we now mention several papers where, in different models with finite resilience, trading strategies are not restricted to be of finite variation. The first instance known to us is Lorenz and Schied [35], who discuss dependence of optimal trade execution strategies on a drift in the unaffected price. In order to react to non-martingale trends they allow for càdlàg semimartingale trading strategies. Gârleanu and Pedersen [26, Section 1.3] allow for strategies of infinite variation in an infinite horizon portfolio optimization problem under market frictions. Becherer et al. [19] prove a continuous extension result for gains of a large investor in the Skorokhod J1J_{1} and M1M_{1} topologies in the class of predictable strategies with càdlàg paths. As discussed in the previous paragraph in more detail, in [1] the strategies are càdlàg semimartingales. In Horst and Kivman [29] càdlàg semimartingale strategies emerge in the limiting case of vanishing instantaneous impact parameter, where the initial modeling framework is inspired by Graewe and Horst [28] and Horst and Xia [31].

To complement the preceding discussion from another perspective, we mention Carmona and Webster [22], who examine high-frequency trading in limit order books in general (not necessarily related with optimal trade execution). It is very interesting that one of their conclusions is a strong empirical evidence for the infinite variation nature of trading strategies of high-frequency traders.

Finally, let us mention that, in the context of trade execution problems, risk terms with zero moving target have been included, e.g., in Ankirchner et al. [9], Ankirchner and Kruse [10], Graewe and Horst [28]. Inequality terminal constraints have been considered in Dolinsky et al. [23], and risk terms with general terminal and moving targets appear in the models of, e.g., Bank et al. [14], Bank and Voß [15], Horst and Naujokat [30], Naujokat and Westray [36]. In particular, [10], [15], and [23] consider random terminal targets ξ\xi within trade execution models where position paths are required to be absolutely continuous functions of time. This restriction of the set of position paths entails technical difficulties that make these problems challenging to analyze. In particular, existence of admissible paths that satisfy the terminal constraint is far from obvious and can in general only be assured under further conditions on ξ\xi. Since in our model position paths are allowed to jump at terminal time we do not face these challenges in our framework.

The paper is structured as follows. Section˜1 is devoted to the continuous extension of our initial trade execution problem to the class of progressively measurable strategies. Section˜2 reduces the problem for the progressively measurable strategies to a standard LQ stochastic control problem. In Section˜3 we present the solution to the obtained LQ problem and trace it back up to the solution to the (extended to progressively measurable strategies) trade execution problem. In Section˜4 we illustrate our results with several examples. Finally, Section˜5 contains the proofs together with some auxiliary results necessary for them.

1 From finite variation to progressively measurable execution strategies

In this section we first set up the finite variation stochastic control problem (see Section˜1.1). In Section˜1.2 we then derive alternative representations of the cost functional and the deviation process which do not require the strategies to be of finite variation. We use these results in Section˜1.3 to extend the cost functional to progressively measurable strategies. In Section˜1.5 we show that this is the unique continuous extension. Section˜1.4 introduces the hidden deviation process as a key tool for the proofs of Section˜1.5. All proofs of this section are deferred to Section˜5.

1.1 The finite variation stochastic control problem

Let T>0T>0 and mm\in\mathbb{N}, m2m\geq 2. We fix a filtered probability space (Ω,T,(s)s[0,T],P)(\Omega,\mathcal{F}_{T},(\mathcal{F}_{s})_{s\in[0,T]},P) satisfying the usual conditions and supporting an mm-dimensional Brownian motion (W1,,Wm)(W^{1},\ldots,W^{m})^{\top} with respect to the filtration (s)(\mathcal{F}_{s}).

We first fix some notation. For t[0,T]t\in[0,T] conditional expectations with respect to t\mathcal{F}_{t} are denoted by Et[]E_{t}[\cdot]. For t[0,T]t\in[0,T] and a càdlàg process X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} a jump at time s[t,T]s\in[t,T] is denoted by ΔXs=XsXs\Delta X_{s}=X_{s}-X_{s-}. We follow the convention that, for t[0,T]t\in[0,T], r[t,T]r\in[t,T] and a càdlàg semimartingale L=(Ls)s[t,T]L=(L_{s})_{s\in[t-,T]}, jumps of the càdlàg integrator LL at time tt contribute to integrals of the form [t,r]𝑑Ls\int_{[t,r]}\ldots dL_{s}. In contrast, we write (t,r]𝑑Ls\int_{(t,r]}\ldots dL_{s} when we do not include jumps of LL at time tt into the integral. The notation tr𝑑Ls\int_{t}^{r}\ldots dL_{s} is sometimes used for continuous integrators LL. For nn\in\mathbb{N} and yny\in\mathbb{R}^{n} let y2=(j=1nyj2)12\lVert y\rVert_{2}=(\sum_{j=1}^{n}y_{j}^{2})^{\frac{1}{2}}. For every t[0,T]t\in[0,T] we mean by L1(Ω,t,P)L^{1}(\Omega,\mathcal{F}_{t},P) the space of all real-valued t\mathcal{F}_{t}-measurable random variables YY such that YL1=E[|Y|]<\lVert Y\rVert_{L^{1}}=E[\lvert Y\rvert]<\infty. For t[0,T]t\in[0,T], let t2=2(Ω×[t,T],Prog(Ω×[t,T]),dP×ds|[t,T])\mathcal{L}_{t}^{2}=\mathcal{L}^{2}(\Omega\times[t,T],\mathrm{Prog}(\Omega\times[t,T]),dP\times ds|_{[t,T]}) denote the space of all (equivalence classes of) real-valued progressively measurable processes u=(us)s[t,T]u=(u_{s})_{s\in[t,T]} such that ut2=(E[tTus2𝑑s])12<\lVert u\rVert_{\mathcal{L}_{t}^{2}}=(E[\int_{t}^{T}u_{s}^{2}ds])^{\frac{1}{2}}<\infty.

The control problem we are about to set up requires as input the real-valued, T\mathcal{F}_{T}-measurable random variable ξ\xi and the real-valued, progressively measurable processes μ=(μs)s[0,T]\mu=(\mu_{s})_{s\in[0,T]}, σ=(σs)s[0,T]\sigma=(\sigma_{s})_{s\in[0,T]}, ρ=(ρs)s[0,T]\rho=(\rho_{s})_{s\in[0,T]}, η=(ηs)s[0,T]\eta=(\eta_{s})_{s\in[0,T]}, r¯=(r¯s)s[0,T]\overline{r}=(\overline{r}_{s})_{s\in[0,T]}, ζ=(ζs)s[0,T]\zeta=(\zeta_{s})_{s\in[0,T]} and λ=(λs)s[0,T]\lambda=(\lambda_{s})_{s\in[0,T]}. We suppose that μ\mu, σ\sigma, ρ\rho, η\eta and λ\lambda are dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. bounded. Moreover, we assume that r¯\overline{r} is [1,1][-1,1]-valued. We define WR=(WsR)s[0,T]W^{R}=(W^{R}_{s})_{s\in[0,T]} by dWsR=r¯sdWs1+1r¯s2dWs2dW^{R}_{s}=\overline{r}_{s}dW^{1}_{s}+\sqrt{1-\overline{r}_{s}^{2}}dW^{2}_{s}, s[0,T]s\in[0,T], W0R=0W^{R}_{0}=0 and refer to r¯\overline{r} as the correlation process. The processes ρ\rho and η\eta give rise to the continuous semimartingale R=(Rs)s[0,T]R=(R_{s})_{s\in[0,T]} with

dRs=ρsds+ηsdWsR,s[0,T],R0=0,dR_{s}=\rho_{s}ds+\eta_{s}dW^{R}_{s},\quad s\in[0,T],\quad R_{0}=0, (3)

which is called the resilience process. We use the processes μ\mu and σ\sigma to define the positive continuous semimartingale γ=(γs)s[0,T]\gamma=(\gamma_{s})_{s\in[0,T]} by

dγs=γs(μsds+σsdWs1),s[0,T],d\gamma_{s}=\gamma_{s}(\mu_{s}ds+\sigma_{s}dW^{1}_{s}),\quad s\in[0,T], (4)

with deterministic initial value γ0>0\gamma_{0}>0. We refer to γ\gamma as the price impact process. Finally, we assume that ξ\xi and ζ\zeta satisfy the integrability conditions

E[γTξ2]<andE[0Tγsζs2𝑑s]<.E[\gamma_{T}\xi^{2}]<\infty\quad\text{and}\quad E\left[\int_{0}^{T}\gamma_{s}\zeta_{s}^{2}ds\right]<\infty. (5)
Remark 1.1.

Note that the components W3,,WmW^{3},\ldots,W^{m} of the Brownian motion are not needed in the dynamics (3) and (4). We introduce these components already here, as in Section˜3, in order to apply the results from the literature on LQ stochastic control, we restrict the present setting a little by assuming that the filtration (s)s[0,T](\mathcal{F}_{s})_{s\in[0,T]} is generated by (W1,,Wm)(W^{1},\ldots,W^{m})^{\top}. The components W3,,WmW^{3},\ldots,W^{m} will therefore serve as further sources of randomness, on which the model inputs may depend.

We next introduce the finite variation strategies that we consider in the sequel. Given t[0,T]t\in[0,T] and dd\in\mathbb{R} we associate to an adapted, càdlàg, finite variation process X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} a process DX=(DsX)s[t,T]D^{X}=(D^{X}_{s})_{s\in[t-,T]} defined by

dDsX=DsXdRs+γsdXs,s[t,T],DtX=d.dD^{X}_{s}=-D^{X}_{s}dR_{s}+\gamma_{s}dX_{s},\quad s\in[t,T],\quad D^{X}_{t-}=d. (6)

If there is no risk of confusion we sometimes simply write DD instead of DXD^{X} in the sequel. For t[0,T]t\in[0,T], x,dx,d\in\mathbb{R} we denote by 𝒜tfv(x,d)\mathcal{A}^{fv}_{t}(x,d) the set of all adapted, càdlàg, finite variation processes X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} satisfying Xt=xX_{t-}=x, XT=ξX_{T}=\xi, and

  1. (A1)

    E[tTγs1(DsX)2𝑑s]<E\left[\int_{t}^{T}\gamma_{s}^{-1}(D_{s}^{X})^{2}ds\right]<\infty,

  2. (A2)

    E[(tT(DsX)4γs2ηs2𝑑s)12]<E\left[\left(\int_{t}^{T}(D_{s}^{X})^{4}\gamma_{s}^{-2}\eta_{s}^{2}ds\right)^{\frac{1}{2}}\right]<\infty,

  3. (A3)

    E[(tT(DsX)4γs2σs2𝑑s)12]<E\left[\left(\int_{t}^{T}(D_{s}^{X})^{4}\gamma_{s}^{-2}\sigma_{s}^{2}ds\right)^{\frac{1}{2}}\right]<\infty.

Any element X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d) is called a finite variation execution strategy. The process D=DXD=D^{X} defined via (6) is called the associated deviation process.

For t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d) and associated DXD^{X}, the cost functional JfvJ^{fv} is given by

Jtfv(x,d,X)=Et[[t,T]DsX𝑑Xs+12[t,T]ΔXsγs𝑑Xs+tTλsγs(Xsζs)2𝑑s].J^{fv}_{t}(x,d,X)=E_{t}\left[\int_{[t,T]}D^{X}_{s-}dX_{s}+\frac{1}{2}\int_{[t,T]}\Delta X_{s}\gamma_{s}dX_{s}+\int_{t}^{T}\lambda_{s}\gamma_{s}\left(X_{s}-\zeta_{s}\right)^{2}ds\right]. (7)

(see the proofs of ˜1.4 and Lemma˜1.6 for well-definedness). The finite variation stochastic control problem consists of minimizing the cost functional JfvJ^{fv} over X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d).

1.2 Alternative representations for the cost functional and the deviation process

For t[0,T]t\in[0,T] we introduce an auxiliary process ν=(νs)s[t,T]\nu=(\nu_{s})_{s\in[t,T]}. It is defined to be the solution of

dνs=νsd(Rs+[R]s),s[t,T],νt=1.d\nu_{s}=\nu_{s}d\left(R_{s}+[R]_{s}\right),\quad s\in[t,T],\quad\nu_{t}=1. (8)

Observe that the inverse is given by

dνs1=νs1dRs,s[t,T],νt1=1.d\nu_{s}^{-1}=-\nu_{s}^{-1}dR_{s},\quad s\in[t,T],\quad\nu_{t}^{-1}=1. (9)
Remark 1.2.

Let t[0,T]t\in[0,T], dd\in\mathbb{R}. With the definition of ν\nu in (8), it holds for all adapted, càdlàg, finite variation processes X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} that the solution DX=(DsX)s[t,T]D^{X}=(D^{X}_{s})_{s\in[t-,T]} of the linear SDE (6) reads DsX=νs1(d+[t,s]νrγr𝑑Xr)D_{s}^{X}=\nu_{s}^{-1}(d+\int_{[t,s]}\nu_{r}\gamma_{r}dX_{r}), s[t,T].s\in[t,T].

Proposition 1.3.

Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}. Suppose that X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} is an adapted, càdlàg, finite variation process with Xt=xX_{t-}=x and with associated process DXD^{X} defined by (6). It then holds that

[t,T]DsX𝑑Xs+12[t,T]ΔXsγs𝑑Xs=12(γT1(DTX)2γt1d2tT(DsX)2νs2d(νs2γs1))\begin{split}&\int_{[t,T]}D^{X}_{s-}dX_{s}+\frac{1}{2}\int_{[t,T]}\Delta X_{s}\gamma_{s}dX_{s}=\frac{1}{2}\left(\gamma_{T}^{-1}(D^{X}_{T})^{2}-\gamma_{t}^{-1}d^{2}-\int_{t}^{T}(D^{X}_{s})^{2}\nu_{s}^{2}d\left(\nu^{-2}_{s}\gamma_{s}^{-1}\right)\right)\end{split} (10)

and

DrX=γrXr+νr1(dγtxtrXsd(νsγs)),r[t,T].\begin{split}D^{X}_{r}&=\gamma_{r}X_{r}+\nu_{r}^{-1}\left(d-\gamma_{t}x-\int_{t}^{r}X_{s}d(\nu_{s}\gamma_{s})\right),\quad r\in[t,T].\end{split} (11)

As a consequence of ˜1.3, and relying on (A1)–(A3), we can rewrite the cost functional JfvJ^{fv} as follows.666Analogues of 1.4 are present in the literature in other related settings; see, e.g., Lemmas 7.4 and 8.6 in [24] and the proof of Lemma 5.3 in Appendix B of [25]. A small technical point, which might be worth noting, is that we present a somewhat different proof below. The idea in [24, 25] is to derive an analogue of (10) by applying the substitution dXs=γs1(dDsX+DsXdRs)dX_{s}=\gamma_{s}^{-1}(dD^{X}_{s}+D^{X}_{s}dR_{s}) and then to compute the expectation. Exactly the same idea would also work in our present setting but it would result in more sustained calculations and, moreover, the right-hand side of (10) would then look rather different (but this would be an equivalent representation, of course). The reason for this is that the process RR, hence DXD^{X}, can have nonvanishing quadratic variation. Here we, essentially, express everything not through DXD^{X} but rather through νDX\nu D^{X}, which has finite variation by Remark 1.2 (as XX has finite variation here). This allows to reduce calculations and provides a somewhat more compact form of (10). To shorten notation, we introduce the process κ=(κs)s[0,T]\kappa=(\kappa_{s})_{s\in[0,T]} defined by

κs=12(2ρs+μsσs2ηs22σsηsr¯s),s[0,T].\kappa_{s}=\frac{1}{2}\big{(}2\rho_{s}+\mu_{s}-\sigma_{s}^{2}-\eta_{s}^{2}-2\sigma_{s}\eta_{s}\overline{r}_{s}\big{)},\quad s\in[0,T]. (12)
Proposition 1.4.

Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}. Suppose that X𝒜tfv(x,d)X\in\mathcal{A}_{t}^{fv}(x,d) with associated deviation process DXD^{X} defined by (6). It then holds that Jtfv(x,d,X)J^{fv}_{t}(x,d,X) in (7) admits the representation

Jtfv(x,d,X)=12Et[γT1(DTX)2+tT(DsX)2γs12κs𝑑s+tT2λsγs(Xsζs)2𝑑s]d22γt a.s.\begin{split}J^{fv}_{t}(x,d,X)&=\frac{1}{2}E_{t}\!\left[\gamma_{T}^{-1}(D_{T}^{X})^{2}+\!\int_{t}^{T}\!(D^{X}_{s})^{2}\gamma_{s}^{-1}2\kappa_{s}ds+\!\int_{t}^{T}\!2\lambda_{s}\gamma_{s}\left(X_{s}-\zeta_{s}\right)^{2}ds\right]\!-\frac{d^{2}}{2\gamma_{t}}\text{ a.s.}\end{split} (13)

1.3 Progressively measurable execution strategies

We point out that the right-hand side of (13) is also well-defined for progressively measurable processes XX satisfying an appropriate integrability condition and with associated deviation DD defined by (11) for which one assumes (A1). This motivates the following extension of the setting from Section˜1.1.

For t[0,T]t\in[0,T], x,dx,d\in\mathbb{R} and a progressively measurable process X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} such that tTXs2𝑑s<\int_{t}^{T}X_{s}^{2}ds<\infty a.s. and Xt=xX_{t-}=x, we define the process DX=(DsX)s[t,T]D^{X}=(D^{X}_{s})_{s\in[t-,T]} by

DsX=γsXs+νs1(dγtxtsXrd(νrγr)),s[t,T],DtX=dD^{X}_{s}=\gamma_{s}X_{s}+\nu_{s}^{-1}\left(d-\gamma_{t}x-\int_{t}^{s}X_{r}d(\nu_{r}\gamma_{r})\right),\quad s\in[t,T],\quad D^{X}_{t-}=d (14)

(recall ν\nu from (8)). Notice that the condition tTXs2𝑑s<\int_{t}^{T}X_{s}^{2}ds<\infty a.s. ensures that the stochastic integral in (14) is well-defined. Again, we sometimes write DD instead of DXD^{X}. Further, for t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, let 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) be the set of (equivalence classes of) progressively measurable processes X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} with Xt=xX_{t-}=x and XT=ξX_{T}=\xi that satisfy tTXs2𝑑s<\int_{t}^{T}X_{s}^{2}ds<\infty a.s. and such that condition (A1) holds true for DXD^{X} defined by (14). To be precise, we stress that the equivalence classes for 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d) are understood with respect to the equivalence relation

X(1)X(2) means\displaystyle X^{(1)}\sim X^{(2)}\text{ means } X.(1)=X.(2)dP×ds-a.e. on Ω×[t,T],\displaystyle X^{(1)}_{.}=X^{(2)}_{.}\;\;dP\times ds\text{-a.e.\ on }\Omega\times[t,T],
Xt(1)=Xt(2)(=x) and XT(1)=XT(2)(=ξ).\displaystyle X^{(1)}_{t-}=X^{(2)}_{t-}\,(=x)\text{ and }X^{(1)}_{T}=X^{(2)}_{T}\,(=\xi). (15)

Any element X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) is called a progressively measurable execution strategy. Again the process D=DXD=D^{X} now defined via (14) is called the associated deviation process. Clearly, we have that 𝒜tfv(x,d)𝒜tpm(x,d)\mathcal{A}^{fv}_{t}(x,d)\subseteq\mathcal{A}^{pm}_{t}(x,d).

Given t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) with associated DXD^{X} (see (14)), we define the cost functional JpmJ^{pm} by

Jtpm(x,d,X)=12Et[γT1(DTX)2+tT(DsX)2γs12κs𝑑s+tT2λsγs(Xsζs)2𝑑s]d22γt.\begin{split}J^{pm}_{t}(x,d,X)&=\frac{1}{2}E_{t}\!\left[\gamma_{T}^{-1}(D_{T}^{X})^{2}+\!\int_{t}^{T}\!(D^{X}_{s})^{2}\gamma_{s}^{-1}2\kappa_{s}ds+\!\int_{t}^{T}\!2\lambda_{s}\gamma_{s}\left(X_{s}-\zeta_{s}\right)^{2}ds\right]\!-\frac{d^{2}}{2\gamma_{t}}.\end{split} (16)

Observe that we have the following corollary of ˜1.3 and ˜1.4.

Corollary 1.5.

Let t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d) with associated deviation process DXD^{X} given by (6). It then holds that X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d), that DXD^{X} satisfies (14), and that Jtfv(x,d,X)=Jtpm(x,d,X)J^{fv}_{t}(x,d,X)=J^{pm}_{t}(x,d,X).

1.4 The hidden deviation process

For t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) with associated deviation process DXD^{X}, we define HX=(HsX)s[t,T]H^{X}=(H^{X}_{s})_{s\in[t,T]} by HsX=DsXγsXsH^{X}_{s}=D^{X}_{s}-\gamma_{s}X_{s}, s[t,T]s\in[t,T]. Observe that if the investor followed a finite variation execution strategy X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d) until time s[t,T]s\in[t,T] and then decided to sell XsX_{s} units of the asset (Xs<0X_{s}<0 means buying) at time ss, then by (6) the resulting deviation at time ss would equal DsXγsXsD^{X}_{s}-\gamma_{s}X_{s}. The value of HsXH^{X}_{s} hence represents the hypothetical deviation if the investor decides to close the position at time s[t,T]s\in[t,T]. We therefore call HXH^{X} the hidden deviation process. Despite X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) and DXD^{X} in general being discontinuous, the hidden deviation process HXH^{X} is always continuous. This can be seen from (14) and the fact that RR (hence also ν\nu) and γ\gamma are continuous. In the case of a finite variation execution strategy X𝒜tfv(x,d)X\in\mathcal{A}^{fv}_{t}(x,d), it holds that dHsX=DsdRsXsdγsdH^{X}_{s}=-D_{s}dR_{s}-X_{s}d\gamma_{s}, s[t,T]s\in[t,T]. In particular, the infinitesimal change of the hidden deviation is driven by the changes of the resilience process and the price impact process.

For t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d), we furthermore introduce the scaled hidden deviation777From the mathematical viewpoint, the scaled hidden deviation plays an extremely important role in what follows. It is, therefore, instructive to see in what kind of units it is measured. The meaning of XX is quantity (of shares), while both DXD^{X} and γ\gamma are measured in $. Thus, the scaled hidden deviation H¯X\overline{H}^{X} is measured in $\sqrt{\$}. H¯X=(H¯sX)s[t,T]\overline{H}^{X}=(\overline{H}^{X}_{s})_{s\in[t,T]} defined by

H¯sX=γs12HsX=γs12DsXγs12Xs,s[t,T].\overline{H}_{s}^{X}=\gamma_{s}^{-\frac{1}{2}}H^{X}_{s}=\gamma_{s}^{-\frac{1}{2}}D^{X}_{s}-\gamma_{s}^{\frac{1}{2}}X_{s},\quad s\in[t,T]. (17)

Also for HXH^{X} and H¯X\overline{H}^{X} we sometimes simply write HH and H¯\overline{H}, respectively. Note that, due to (14), it holds that H¯sX=γs12νs1(dγtxtsXrd(νrγr))\overline{H}^{X}_{s}=\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}(d-\gamma_{t}x-\int_{t}^{s}X_{r}d(\nu_{r}\gamma_{r})), s[t,T]s\in[t,T].

We next show that the scaled hidden deviation process satisfies a linear SDE and an L2L^{2}-bound. Moreover, we derive a representation of JpmJ^{pm} in terms of the scaled hidden deviation process.

Lemma 1.6.

Let t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d). Then it holds that

dH¯sX=(12(μs14σs2)H¯sX12(2(ρs+μs)σs2σsηsr¯s)γs12DsX)ds+(12σsH¯sX(σs+ηsr¯s)γs12DsX)dWs1ηs1r¯s2γs12DsXdWs2,s[t,T],H¯tX=dγtγtx,\begin{split}d\overline{H}^{X}_{s}&=\left(\frac{1}{2}\left(\mu_{s}-\frac{1}{4}\sigma_{s}^{2}\right)\overline{H}_{s}^{X}-\frac{1}{2}\left(2(\rho_{s}+\mu_{s})-\sigma_{s}^{2}-\sigma_{s}\eta_{s}\overline{r}_{s}\right)\gamma_{s}^{-\frac{1}{2}}D^{X}_{s}\right)ds\\ &\quad+\left(\frac{1}{2}\sigma_{s}\overline{H}^{X}_{s}-(\sigma_{s}+\eta_{s}\overline{r}_{s})\gamma_{s}^{-\frac{1}{2}}D^{X}_{s}\right)dW^{1}_{s}-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\gamma_{s}^{-\frac{1}{2}}D^{X}_{s}dW_{s}^{2},\quad s\in[t,T],\\ \overline{H}^{X}_{t}&=\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,\end{split} (18)

that E[sups[t,T](H¯sX)2]<E[\sup_{s\in[t,T]}(\overline{H}_{s}^{X})^{2}]<\infty, and that

Jtpm(x,d,X)=12Et[(H¯TX+γTξ)2+tT2(κs+λs)γs1(DsX)2𝑑s]d22γt+Et[tT(λs(H¯sX+γsζs)22λs(H¯sX+γsζs)γs12DsX)𝑑s].\begin{split}J^{pm}_{t}(x,d,X)&=\frac{1}{2}E_{t}\bigg{[}\big{(}\overline{H}^{X}_{T}+\sqrt{\gamma_{T}}\xi\big{)}^{2}+\int_{t}^{T}2(\kappa_{s}+\lambda_{s})\gamma_{s}^{-1}(D_{s}^{X})^{2}ds\bigg{]}-\frac{d^{2}}{2\gamma_{t}}\\ &\quad+E_{t}\bigg{[}\int_{t}^{T}\left(\lambda_{s}\left(\overline{H}^{X}_{s}+\sqrt{\gamma_{s}}\zeta_{s}\right)^{2}-2\lambda_{s}\left(\overline{H}^{X}_{s}+\sqrt{\gamma_{s}}\zeta_{s}\right)\gamma_{s}^{-\frac{1}{2}}D_{s}^{X}\right)ds\bigg{]}.\end{split} (19)

1.5 Continuous extension of the cost functional

Corollary˜1.5 states that for finite variation execution strategies, the cost functionals JfvJ^{fv} and JpmJ^{pm} are the same. In this subsection we show that JpmJ^{pm} can be considered as an extension of JfvJ^{fv} to progressively measurable strategies; i.e., we introduce a metric 𝐝\mathbf{d} on 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d) and show that Jtpm(x,d,X)J^{pm}_{t}(x,d,X) is continuous in the strategy X𝒜tpm(x,d)X\in\mathcal{A}_{t}^{pm}(x,d) (the first part of ˜1.7), that 𝒜tfv(x,d)\mathcal{A}^{fv}_{t}(x,d) is dense in 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) (the second part of ˜1.7) and that the metric space (𝒜tpm(x,d),𝐝)(\mathcal{A}_{t}^{pm}(x,d),\mathbf{d}) is complete (the third part of ˜1.7). The first and the second parts of ˜1.7 mean that, under the metric 𝐝\mathbf{d}, Jtpm(x,d,)J_{t}^{pm}(x,d,\cdot) is a unique continuous extension of Jtfv(x,d,)J_{t}^{fv}(x,d,\cdot) from 𝒜tfv(x,d)\mathcal{A}_{t}^{fv}(x,d) onto 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d). The third part of ˜1.7 means that, under the metric 𝐝\mathbf{d}, 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d) is the largest space where such a continuous extension is uniquely determined by Jtfv(x,d,)J_{t}^{fv}(x,d,\cdot) on 𝒜tfv(x,d)\mathcal{A}_{t}^{fv}(x,d). This is because the completeness of (𝒜tpm(x,d),𝐝)(\mathcal{A}_{t}^{pm}(x,d),\mathbf{d}) is equivalent to the following statement: For any metric space (𝒜^t(x,d),𝐝^)(\widehat{\mathcal{A}}_{t}(x,d),\widehat{\mathbf{d}}) containing 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d) and such that 𝐝^|𝒜tpm(x,d)2=𝐝\widehat{\mathbf{d}}|_{\mathcal{A}_{t}^{pm}(x,d)^{2}}=\mathbf{d}, it holds that the set 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d) is closed in 𝒜^t(x,d)\widehat{\mathcal{A}}_{t}(x,d).

For t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and X,Y𝒜tpm(x,d)X,Y\in\mathcal{A}_{t}^{pm}(x,d) with associated deviation processes DXD^{X}, DYD^{Y} defined by (14), we define

𝐝(X,Y)=(E[tT(DsXDsY)2γs1𝑑s])12.\mathbf{d}(X,Y)=\left(E\left[\int_{t}^{T}(D_{s}^{X}-D_{s}^{Y})^{2}\gamma_{s}^{-1}ds\right]\right)^{\frac{1}{2}}. (20)

Identifying any processes that are equal dP×ds|[t,T]dP\times ds|_{[t,T]}-a.e., this indeed is a metric on 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d), see Lemma˜5.2.

Note that, for fixed t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}, we may consider the cost functional (16) as a function Jtpm(x,d,):(𝒜tpm(x,d),𝐝)(L1(Ω,t,P),L1).J^{pm}_{t}(x,d,\cdot)\colon(\mathcal{A}_{t}^{pm}(x,d),\mathbf{d})\to(L^{1}(\Omega,\mathcal{F}_{t},P),\lVert\cdot\rVert_{L^{1}}). Indeed, using (A1), Lemma˜1.6, (5), and boundedness of the input processes, we see that Jtpm(x,d,X)L1(Ω,t,P)J^{pm}_{t}(x,d,X)\in L^{1}(\Omega,\mathcal{F}_{t},P) for all X𝒜tpm(x,d)X\in\mathcal{A}_{t}^{pm}(x,d).

Theorem 1.7.

Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}.

(i) Suppose that X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d). For every sequence (Xn)n(X^{n})_{n\in\mathbb{N}} in 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) with limn𝐝(Xn,X)=0\lim_{n\to\infty}\mathbf{d}(X^{n},X)=0 it holds that limnJtpm(x,d,Xn)Jtpm(x,d,X)L1=0\lim_{n\to\infty}\lVert J^{pm}_{t}(x,d,X^{n})-J^{pm}_{t}(x,d,X)\rVert_{L^{1}}=0.

(ii) For any X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) there exists a sequence (Xn)n(X^{n})_{n\in\mathbb{N}} in 𝒜tfv(x,d)\mathcal{A}^{fv}_{t}(x,d) such that limn𝐝(Xn,X)=0\lim_{n\to\infty}\mathbf{d}(X^{n},X)=0. In particular, it holds that

essinfX𝒜tfv(x,d)Jtfv(x,d,X)=essinfX𝒜tpm(x,d)Jtpm(x,d,X).\operatorname*{ess\,inf}_{X\in\mathcal{A}^{fv}_{t}(x,d)}J^{fv}_{t}(x,d,X)=\operatorname*{ess\,inf}_{X\in\mathcal{A}^{pm}_{t}(x,d)}J^{pm}_{t}(x,d,X). (21)

(iii) For any Cauchy sequence (Xn)n(X^{n})_{n\in\mathbb{N}} in (𝒜tpm(x,d),𝐝)(\mathcal{A}_{t}^{pm}(x,d),\mathbf{d}) there exists some X0𝒜tpm(x,d)X^{0}\in\mathcal{A}_{t}^{pm}(x,d) such that limn𝐝(Xn,X0)=0\lim_{n\to\infty}\mathbf{d}(X^{n},X^{0})=0.

In Corollary˜3.4 below we provide sufficient conditions that ensure that the infimum on the right-hand side of (21) is indeed a minimum.

2 Reduction to a standard LQ stochastic control problem

In this section we recast the problem of minimizing JpmJ^{pm} over X𝒜tpm(x,d)X\in\mathcal{A}_{t}^{pm}(x,d) as a standard LQ stochastic control problem. All proofs of this section are given in Section˜5.

2.1 The first reduction

Note that (19) in Lemma˜1.6 shows that for t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) the costs Jtpm(x,d,X)J^{pm}_{t}(x,d,X) depend in a quadratic way on (H¯X,γ12DX)(\overline{H}^{X},\gamma^{-\frac{1}{2}}D^{X}). Moreover, (18) in Lemma˜1.6 ensures that the dynamics of H¯X\overline{H}^{X} depend linearly on (H¯X,γ12DX)(\overline{H}^{X},\gamma^{-\frac{1}{2}}D^{X}). These two observations suggest to view the minimization problem of JpmJ^{pm} over X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) as a standard LQ stochastic control problem with state process H¯X\overline{H}^{X} and control γ12DX\gamma^{-\frac{1}{2}}D^{X}. This motivates the following definitions. For every t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and ut2u\in\mathcal{L}_{t}^{2}, we consider the state process H~u=(H~su)s[t,T]\widetilde{H}^{u}=(\widetilde{H}^{u}_{s})_{s\in[t,T]} defined by

dH~su=(12(μs14σs2)H~su12(2(ρs+μs)σs2σsηsr¯s)us)ds+(12σsH~su(σs+ηsr¯s)us)dWs1ηs1r¯s2usdWs2,s[t,T],H~tu=dγtγtx,\begin{split}d\widetilde{H}^{u}_{s}&=\left(\frac{1}{2}\left(\mu_{s}-\frac{1}{4}\sigma_{s}^{2}\right)\widetilde{H}^{u}_{s}-\frac{1}{2}\left(2(\rho_{s}+\mu_{s})-\sigma_{s}^{2}-\sigma_{s}\eta_{s}\overline{r}_{s}\right)u_{s}\right)ds\\ &\quad+\left(\frac{1}{2}\sigma_{s}\widetilde{H}^{u}_{s}-(\sigma_{s}+\eta_{s}\overline{r}_{s})u_{s}\right)dW^{1}_{s}-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}u_{s}dW_{s}^{2},\quad s\in[t,T],\\ \widetilde{H}^{u}_{t}&=\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,\end{split} (22)

and the cost functional JJ defined by

Jt(dγtγtx,u)=12Et[(H~Tu+γTξ)2+tT2(κs+λs)us2ds+tT(2λs(H~su+γsζs)24λs(H~su+γsζs)us)ds].\begin{split}J_{t}\left(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,u\right)&=\frac{1}{2}E_{t}\bigg{[}\big{(}\widetilde{H}^{u}_{T}+\sqrt{\gamma_{T}}\xi\big{)}^{2}+\int_{t}^{T}2(\kappa_{s}+\lambda_{s})u_{s}^{2}ds\\ &\quad\quad\quad+\int_{t}^{T}\left(2\lambda_{s}\left(\widetilde{H}^{u}_{s}+\sqrt{\gamma_{s}}\zeta_{s}\right)^{2}-4\lambda_{s}\left(\widetilde{H}^{u}_{s}+\sqrt{\gamma_{s}}\zeta_{s}\right)u_{s}\right)ds\bigg{]}.\end{split} (23)

Once again we sometimes simply write H~\widetilde{H} instead of H~u\widetilde{H}^{u}. The LQ stochastic control problem is to minimize (23) over the set of admissible controls t2\mathcal{L}_{t}^{2}.

It holds that for every progressively measurable execution strategy X𝒜tpm(x,d)X\in\mathcal{A}_{t}^{pm}(x,d) there exists a control ut2u\in\mathcal{L}_{t}^{2} such that the cost functional JpmJ^{pm} can be rewritten in terms of JJ (and d22γt-\frac{d^{2}}{2\gamma_{t}}). In fact, this is achieved by taking u=γ12DXu=\gamma^{-\frac{1}{2}}D^{X}, as outlined in the motivation above. We state this as Lemma˜2.1.

Lemma 2.1.

Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}. Suppose that X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) with associated deviation DXD^{X}. Define u=(us)s[t,T]u=(u_{s})_{s\in[t,T]} by us=γs12DsXu_{s}=\gamma_{s}^{-\frac{1}{2}}D^{X}_{s}, s[t,T]s\in[t,T]. It then holds that ut2u\in\mathcal{L}_{t}^{2} and that Jtpm(x,d,X)=Jt(dγtγtx,u)d22γtJ^{pm}_{t}(x,d,X)=J_{t}(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,u)-\frac{d^{2}}{2\gamma_{t}} a.s.

On the other hand, we may also start with ut2u\in\mathcal{L}_{t}^{2} and derive a progressively measurable execution strategy X𝒜tpm(x,d)X\in\mathcal{A}_{t}^{pm}(x,d) such that the expected costs match.

Lemma 2.2.

Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}. Suppose that u=(us)s[t,T]t2u=(u_{s})_{s\in[t,T]}\in\mathcal{L}_{t}^{2} and let H~u\widetilde{H}^{u} be the associated solution of (22). Define X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} by Xs=γs12(usH~su)X_{s}=\gamma_{s}^{-\frac{1}{2}}(u_{s}-\widetilde{H}^{u}_{s}), s[t,T)s\in[t,T), Xt=xX_{t-}=x, XT=ξX_{T}=\xi. It then holds that X𝒜tpm(x,d)X\in\mathcal{A}_{t}^{pm}(x,d) and that Jtpm(x,d,X)=Jt(dγtγtx,u)d22γtJ^{pm}_{t}(x,d,X)=J_{t}(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,u)-\frac{d^{2}}{2\gamma_{t}} a.s.

Lemma˜2.1 and Lemma˜2.2 together with ˜1.7 establish the following equivalence of the control problems pertaining to JfvJ^{fv}, JpmJ^{pm}, and JJ.

Corollary 2.3.

For t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R} it holds that

essinfX𝒜tfv(x,d)Jtfv(x,d,X)=essinfX𝒜tpm(x,d)Jtpm(x,d,X)=essinfut2Jt(dγtγtx,u)d22γt a.s.\operatorname*{ess\,inf}_{X\in\mathcal{A}^{fv}_{t}(x,d)}J^{fv}_{t}(x,d,X)=\operatorname*{ess\,inf}_{X\in\mathcal{A}^{pm}_{t}(x,d)}J^{pm}_{t}(x,d,X)=\operatorname*{ess\,inf}_{u\in\mathcal{L}_{t}^{2}}J_{t}\left(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,u\right)-\frac{d^{2}}{2\gamma_{t}}\text{ a.s.}

Furthermore, Lemma˜2.1, Lemma˜2.2, and Corollary˜2.3 provide a method to obtain an optimal progressively measurable execution strategy and potentially an optimal finite variation execution strategy from the standard optimal control problem and vice versa.

Corollary 2.4.

Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}.

(i) Suppose that X=(Xs)s[t,T]𝒜tpm(x,d)X^{*}=(X^{*}_{s})_{s\in[t-,T]}\in\mathcal{A}_{t}^{pm}(x,d) minimizes JpmJ^{pm} over 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d) and let DXD^{X^{*}} be the associated deviation process. Then, u=(us)s[t,T]u^{*}=(u^{*}_{s})_{s\in[t,T]} defined by us=γs12DsXu^{*}_{s}=\gamma_{s}^{-\frac{1}{2}}D^{X^{*}}_{s}, s[t,T]s\in[t,T], minimizes JJ over t2\mathcal{L}_{t}^{2}.

(ii) Suppose that u=(us)s[t,T]t2u^{*}=(u^{*}_{s})_{s\in[t,T]}\in\mathcal{L}_{t}^{2} minimizes JJ over t2\mathcal{L}_{t}^{2} and let H~u\widetilde{H}^{u^{*}} be the associated solution of (22) for uu^{*}. Then, X=(Xs)s[t,T]X^{*}=(X^{*}_{s})_{s\in[t,T]} defined by Xs=γs12(usH~su)X^{*}_{s}=\gamma_{s}^{-\frac{1}{2}}(u^{*}_{s}-\widetilde{H}^{u^{*}}_{s}), s[t,T)s\in[t,T), Xt=xX_{t-}^{*}=x, XT=ξX_{T}^{*}=\xi, minimizes JpmJ^{pm} over 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d).

Moreover, if X𝒜tfv(x,d)X^{*}\in\mathcal{A}_{t}^{fv}(x,d) (in the sense that there is an element of 𝒜tfv(x,d)\mathcal{A}_{t}^{fv}(x,d) within the equivalence class of XX^{*}, see (15)), then XX^{*} minimizes JfvJ^{fv} over 𝒜tfv(x,d)\mathcal{A}_{t}^{fv}(x,d).

2.2 Formulation without cross-terms

Note that the last integral in the definition (23) of the cost functional JJ involves a product between the state process H~u\widetilde{H}^{u} and the control process uu. A larger part of the literature on LQ optimal control considers cost functionals that do not contain such cross-terms. In particular, this applies to [34], whose results we apply in Section˜3 below. For this reason we provide in this subsection a reformulation of the control problem (22) and (23) that does not contain cross-terms. In order to carry out the transformation necessary for this, we need to impose a further condition on our model inputs. We assume that there exists a constant C[0,)C\in[0,\infty) such that for all s[0,T]s\in[0,T] we have PP-a.s. that

|λs|C|λs+κs|.\lvert\lambda_{s}\rvert\leq C\lvert\lambda_{s}+\kappa_{s}\rvert. (24)

Note that this assumption ensures that the set {λs+κs=0}\{\lambda_{s}+\kappa_{s}=0\} is a subset of {λs=0}\{\lambda_{s}=0\} (up to a PP-null set). For this reason we, in the sequel, use the following

Convention: Under (24) we always understand λsλs+κs=0\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}=0 on the set {λs+κs=0}\{\lambda_{s}+\kappa_{s}=0\}.

Now in order to get rid of the cross-term in (23) we transform for t[0,T]t\in[0,T] any control process ut2u\in\mathcal{L}_{t}^{2} in an affine way to u^s=usλsλs+κs(H~su+γsζs)\hat{u}_{s}=u_{s}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\widetilde{H}^{u}_{s}+\sqrt{\gamma_{s}}\zeta_{s}), s[t,T]s\in[t,T]. This leads to the new controlled state process H^u^=(H^su^)s[t,T]\widehat{H}^{\hat{u}}=(\widehat{H}^{\hat{u}}_{s})_{s\in[t,T]} which is defined for every t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and u^t2\hat{u}\in\mathcal{L}_{t}^{2} by

dH^su^=(μs218σs2λsλs+κs(ρs+μsσs2+σsηsr¯s2))H^su^ds(ρs+μsσs2+σsηsr¯s2)u^sdsλsλs+κs(ρs+μsσs2+σsηsr¯s2)γsζsds+(σs2λsλs+κs(σs+ηsr¯s))H^su^dWs1(σs+ηsr¯s)u^sdWs1λsλs+κs(σs+ηsr¯s)γsζsdWs1λsλs+κsηs1r¯s2H^su^dWs2ηs1r¯s2u^sdWs2λsλs+κsηs1r¯s2γsζsdWs2,s[t,T],H^tu^=dγtγtx.\begin{split}d\widehat{H}_{s}^{\hat{u}}&=\left(\frac{\mu_{s}}{2}-\frac{1}{8}\sigma_{s}^{2}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\left(\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\right)\right)\widehat{H}_{s}^{\hat{u}}ds\\ &\quad-\left(\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\right)\hat{u}_{s}ds-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\left(\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\right)\sqrt{\gamma_{s}}\zeta_{s}ds\\ &\quad+\left(\frac{\sigma_{s}}{2}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\sigma_{s}+\eta_{s}\overline{r}_{s})\right)\widehat{H}_{s}^{\hat{u}}dW_{s}^{1}-(\sigma_{s}+\eta_{s}\overline{r}_{s})\hat{u}_{s}dW_{s}^{1}\\ &\quad-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\sigma_{s}+\eta_{s}\overline{r}_{s})\sqrt{\gamma_{s}}\zeta_{s}dW_{s}^{1}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\widehat{H}_{s}^{\hat{u}}dW_{s}^{2}\\ &\quad-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\hat{u}_{s}dW_{s}^{2}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\sqrt{\gamma_{s}}\zeta_{s}dW_{s}^{2},\quad s\in[t,T],\\ \widehat{H}_{t}^{\hat{u}}&=\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x.\end{split} (25)

The meaning of (25) is that we only reparametrize the control (uu^u\to\hat{u}) but not the state variable (H^u^=H~u\widehat{H}^{\hat{u}}=\widetilde{H}^{u}), see Lemma˜2.5 for the formal statement. For t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, u^t2\hat{u}\in\mathcal{L}_{t}^{2} and associated H^u^\widehat{H}^{\hat{u}}, we define the cost functional J^\hat{J} by

J^t(dγtγtx,u^)=Et[12(H^Tu^+γTξ)2+tT(λsκsλs+κs(H^su^+γsζs)2+(λs+κs)u^s2)𝑑s].\begin{split}\hat{J}_{t}\!\left(\!\frac{d}{\sqrt{\gamma_{t}}}-\!\sqrt{\gamma_{t}}x,\hat{u}\!\right)&\!=\!E_{t}\bigg{[}\frac{1}{2}\big{(}\widehat{H}^{\hat{u}}_{T}+\!\sqrt{\gamma_{T}}\xi\big{)}^{2}\!+\!\!\int_{t}^{T}\!\!\!\left(\!\frac{\lambda_{s}\kappa_{s}}{\lambda_{s}+\kappa_{s}}\left(\!\widehat{H}^{\hat{u}}_{s}+\!\sqrt{\gamma_{s}}\zeta_{s}\right)^{2}\!\!+\!(\lambda_{s}\!+\!\kappa_{s})\hat{u}_{s}^{2}\right)ds\bigg{]}.\end{split} (26)

This cost functional does not exhibit cross-terms, but is equivalent to JJ of (23) in the sense of the following lemma.

Lemma 2.5.

Assume that (24) holds true. Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}.

(i) Suppose that ut2u\in\mathcal{L}_{t}^{2} with associated state process H~u\widetilde{H}^{u} defined by (22). Then, u^=(u^s)s[t,T]\hat{u}=(\hat{u}_{s})_{s\in[t,T]} defined by u^s=usλsλs+κs(H~su+γsζs)\hat{u}_{s}=u_{s}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\widetilde{H}^{u}_{s}+\sqrt{\gamma_{s}}\zeta_{s}), s[t,T]s\in[t,T], is in t2\mathcal{L}_{t}^{2}, and it holds that H^u^=H~u\widehat{H}^{\hat{u}}=\widetilde{H}^{u} and Jt(dγtγtx,u)=J^t(dγtγtx,u^)J_{t}(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,u)=\hat{J}_{t}(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,\hat{u}).

(ii) Suppose that u^t2\hat{u}\in\mathcal{L}_{t}^{2} with associated state process H^u^\widehat{H}^{\hat{u}} defined by (25). Then, u=(us)s[t,T]u=(u_{s})_{s\in[t,T]} defined by us=u^s+λsλs+κs(H^su^+γsζs)u_{s}=\hat{u}_{s}+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\widehat{H}^{\hat{u}}_{s}+\sqrt{\gamma_{s}}\zeta_{s}), s[t,T]s\in[t,T], is in t2\mathcal{L}_{t}^{2}, and it holds that H~u=H^u^\widetilde{H}^{u}=\widehat{H}^{\hat{u}} and Jt(dγtγtx,u)=J^t(dγtγtx,u^)J_{t}(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,u)=\hat{J}_{t}(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,\hat{u}).

As a corollary, we obtain the following link between an optimal control for J^\hat{J} and an optimal control for JJ.

Corollary 2.6.

Assume that (24) holds true. Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}.

(i) Suppose that u=(us)s[t,T]t2u^{*}=(u^{*}_{s})_{s\in[t,T]}\in\mathcal{L}_{t}^{2} is an optimal control for JJ, and let H~u\widetilde{H}^{u^{*}} be the solution of (22) for uu^{*}. Then, u^=(u^s)s[t,T]\hat{u}^{*}=(\hat{u}^{*}_{s})_{s\in[t,T]} defined by u^s=usλsλs+κs(H~su+γsζs)\hat{u}^{*}_{s}=u^{*}_{s}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\widetilde{H}^{u^{*}}_{s}+\sqrt{\gamma_{s}}\zeta_{s}), s[t,T]s\in[t,T], is an optimal control in t2\mathcal{L}_{t}^{2} for J^\hat{J}.

(ii) Suppose that u^=(u^s)s[t,T]t2\hat{u}^{*}=(\hat{u}^{*}_{s})_{s\in[t,T]}\in\mathcal{L}_{t}^{2} is an optimal control for J^\hat{J}, and let H^u^\widehat{H}^{\hat{u}^{*}} be the solution of (25) for u^\hat{u}^{*}. Then, u=(us)s[t,T]u^{*}=(u^{*}_{s})_{s\in[t,T]} defined by us=u^s+λsλs+κs(H^su^+γsζs)u^{*}_{s}=\hat{u}^{*}_{s}+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\widehat{H}^{\hat{u}^{*}}_{s}+\sqrt{\gamma_{s}}\zeta_{s}), s[t,T]s\in[t,T], is an optimal control in t2\mathcal{L}_{t}^{2} for JJ.

3 Solving the LQ control problem and the trade execution problem

We now solve the LQ control problem from Section˜2 and consequently obtain a solution of the trade execution problem.

Remark 3.1.

The solution approach of [34], which we are about to apply, is built on the tight connection between standard LQ stochastic control problems and Riccati-type BSDEs (BSRDEs). This connection is well known and dates back at least to Bismut (see, e.g., [20] and [21]). The central challenge in this approach is to establish the existence of a solution of the BSRDE. Kohlmann and Tang prove in [34] such results in a general framework which in particular covers our problem formulation in Section˜2.2 under appropriate assumptions.

There is a variety of further results in the literature on LQ stochastic control problems that provide existence results for BSRDEs under different sets of assumptions. A specific potential further possibility is, for example, to use the results of the recent article [40] by Sun et al. in our setting. The set-up of [40] allows for cross-terms in the cost functional and, more interestingly, the results in [40] hold under a uniform convexity assumption on the cost functional, which is a weaker requirement than the usually imposed nonnegativity and positivity assumptions on the coefficients of the cost functional. However, in general, the terminal costs and the running costs in (23) (and also in (26)) contain terms such as (H~Tu+γTξ)2(\widetilde{H}^{u}_{T}+\sqrt{\gamma_{T}}\xi)^{2} and λs(H~su+γsζs)2\lambda_{s}(\widetilde{H}^{u}_{s}+\sqrt{\gamma_{s}}\zeta_{s})^{2}, which are inhomogeneous. Therefore, the results of [40] are only directly applicable in the special case where ξ=0\xi=0 and at least one of λ\lambda and ζ\zeta vanishes. A possible route for future research could be to incorporate inhomogeneous control problems as presented in Section˜2 to the framework of [40].

Setting in Section˜3: In our general setting (see Section˜1.1) we additionally assume that the filtration (s)s[0,T](\mathcal{F}_{s})_{s\in[0,T]} for the filtered probability space (Ω,T,(s)s[0,T],P)(\Omega,\mathcal{F}_{T},(\mathcal{F}_{s})_{s\in[0,T]},P) is the augmented natural filtration of the Brownian motion (W1,,Wm)(W^{1},\ldots,W^{m})^{\top}. Furthermore, we set the initial time to t=0t=0. We also assume that λ\lambda and κ=12(2ρ+μσ2η22σηr¯)\kappa=\frac{1}{2}(2\rho+\mu-\sigma^{2}-\eta^{2}-2\sigma\eta\overline{r}) are nonnegative dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e.888We stress at this point that the results presented in Sections 1 and 2 are valid for more general filtrations and for processes λ\lambda and κ\kappa possibly taking negative values. This opens the way for applying Sections 1 and 2 in other settings in future research.

Remark 3.2.

Note that the assumption of nonnegativity of λ\lambda and κ\kappa is necessary to apply the results of [34]. Indeed, [34] requires that λ+κ\lambda+\kappa (the coefficient in front of u^2\hat{u}^{2} in (26)) and λκλ+κ\frac{\lambda\kappa}{\lambda+\kappa} (the coefficient in front of (H^su^+γsζs)2(\widehat{H}^{\hat{u}}_{s}+\sqrt{\gamma_{s}}\zeta_{s})^{2} in (26)) are nonnegative and bounded, which implies that λ\lambda and κ\kappa have to be nonnegative.

Moreover, we note that nonnegativity of λ\lambda and κ\kappa ensures that (24) is satisfied. Further, we observe that the mentioned coefficients λ+κ\lambda+\kappa and λκλ+κ\frac{\lambda\kappa}{\lambda+\kappa} are bounded, as required. Indeed, it clearly holds λκλ+κκ\frac{\lambda\kappa}{\lambda+\kappa}\leq\kappa, and it remains to recall that μ,σ,ρ,η\mu,\sigma,\rho,\eta, and λ\lambda are bounded and r¯\overline{r} is [1,1][-1,1]-valued (see Section˜1.1).

Note that the LQ control problem of Section˜2.2, which consists of minimizing J^\hat{J} in (26) with state dynamics given by (25), is of the form considered in [34, (79)-(81)]. The solution can be described by the two BSDEs [34, (9) and (85)]. The first one, [34, (9)], is a Riccati-type BSDE, which in our setting reads

dKs=[(μs+λsλs+κs(λsλs+κs(σs2+2σsηsr¯s+ηs2)2(ρs+μs)))Ks+(σsλsλs+κs2(σs+ηsr¯s))Ls1λsλs+κs2ηs1r¯s2Ls2+λsκsλs+κs((ρs+μsλsλs+κs(σs2+2σsηsr¯s+ηs2))Ks+(σs+ηsr¯s)Ls1+ηs1r¯s2Ls2)2λs+κs+(σs2+2σsηsr¯s+ηs2)Ks]ds+j=1mLsjdWsj,s[0,T],KT=12.\begin{split}dK_{s}&=-\Bigg{[}\left(\mu_{s}+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\bigg{(}\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})-2(\rho_{s}+\mu_{s})\bigg{)}\right)K_{s}\\ &\qquad+\left(\sigma_{s}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}2(\sigma_{s}+\eta_{s}\overline{r}_{s})\right)L^{1}_{s}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}2\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}L^{2}_{s}+\frac{\lambda_{s}\kappa_{s}}{\lambda_{s}+\kappa_{s}}\\[2.84526pt] &\qquad-\frac{\left(\left(\rho_{s}+\mu_{s}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})\right)K_{s}+(\sigma_{s}+\eta_{s}\overline{r}_{s})L^{1}_{s}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}L^{2}_{s}\right)^{2}}{\lambda_{s}+\kappa_{s}+(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})K_{s}}\Bigg{]}ds\\ &\quad+\sum_{j=1}^{m}L_{s}^{j}dW_{s}^{j},\quad s\in[0,T],\\ K_{T}&=\frac{1}{2}.\end{split} (27)

We call a pair (K,L)(K,L) with L=(L1,L2,,Lm)L=(L^{1},L^{2},\ldots,L^{m})^{\top} a solution to BSDE (27) if

  1. (i)

    KK is an adapted, continuous, nonnegative, and bounded process,

  2. (ii)

    λ+κ+(σ2+2σηr¯+η2)K>0\lambda+\kappa+(\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2})K>0 dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e.,

  3. (iii)

    L1,,Lm02L^{1},\ldots,L^{m}\in\mathcal{L}_{0}^{2}, and

  4. (iv)

    BSDE (27) is satisfied PP-a.s.

A discussion of this definition is in order. The requirement of nonnegativity and boundedness of KK can be explained at this point by the fact that, under mild conditions, such a solution exists (see ˜3.3 below). Condition (ii) ensures that there is no problem with division in the driver of (27), where the quantity λ+κ+(σ2+2σηr¯+η2)K\lambda+\kappa+(\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2})K appears in the denominator. Moreover, it is worth noting that, for a nonnegative KK, in our setting we always have λ+κ+(σ2+2σηr¯+η2)K0\lambda+\kappa+(\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2})K\geq 0, as σ2+2σηr¯+η2=(σ+ηr¯)2+η2(1r¯2)\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2}=(\sigma+\eta\overline{r})^{2}+\eta^{2}(1-\overline{r}^{2}). From this we also see that the quantity λ+κ+(σ2+2σηr¯+η2)K\lambda+\kappa+(\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2})K can vanish only in “very degenerate” situations. The conclusion is that condition (ii) is quite natural.

To shorten notation, we introduce, for a solution (K,L)(K,L) of BSDE (27), the process θ=(θs)s[0,T]\theta=(\theta_{s})_{s\in[0,T]} by, for s[0,T]s\in[0,T],

θs=(ρs+μsλsλs+κs(σs2+2σsηsr¯s+ηs2))Ks+(σs+ηsr¯s)Ls1+ηs1r¯s2Ls2λs+κs+(σs2+2σsηsr¯s+ηs2)Ks.\theta_{s}=\frac{\left(\rho_{s}+\mu_{s}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})\right)K_{s}+(\sigma_{s}+\eta_{s}\overline{r}_{s})L^{1}_{s}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}L^{2}_{s}}{\lambda_{s}+\kappa_{s}+(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})K_{s}}. (28)

Next, we consider the second BSDE [34, (85)], which is linear and reads in our setting

dψs=[(μs2σs28(ρs+μsσs2+σsηsr¯s2)(λsλs+κs+θs))ψs+(σs2(σs+ηsr¯s)(λsλs+κs+θs))(ϕs1+λsλs+κs(σs+ηsr¯s)γsζsKs)ηs1r¯s2(λsλs+κs+θs)(ϕs2+λsλs+κsηs1r¯s2γsζsKs)+λsλs+κsγsζs((ρs+μsσs2+σsηsr¯s2)Ks+(σs+ηsr¯s)Ls1+ηs1r¯s2Ls2)λsκsλs+κsγsζs]ds+j=1mϕsjdWsj,s[0,T],ψT=12γTξ.\begin{split}d\psi_{s}&=-\Bigg{[}\left(\frac{\mu_{s}}{2}-\frac{\sigma_{s}^{2}}{8}-\left(\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\right)\left(\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}+\theta_{s}\right)\right)\psi_{s}\\ &\qquad+\left(\frac{\sigma_{s}}{2}-(\sigma_{s}+\eta_{s}\overline{r}_{s})\left(\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}+\theta_{s}\right)\right)\left(\phi_{s}^{1}+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\sigma_{s}+\eta_{s}\overline{r}_{s})\sqrt{\gamma_{s}}\zeta_{s}K_{s}\right)\\ &\qquad-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\left(\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}+\theta_{s}\right)\left(\phi_{s}^{2}+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\sqrt{\gamma_{s}}\zeta_{s}K_{s}\right)\\ &\qquad+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\sqrt{\gamma_{s}}\zeta_{s}\left(\left(\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\right)K_{s}+(\sigma_{s}+\eta_{s}\overline{r}_{s})L_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}L_{s}^{2}\right)\\ &\qquad-\frac{\lambda_{s}\kappa_{s}}{\lambda_{s}+\kappa_{s}}\sqrt{\gamma_{s}}\zeta_{s}\Bigg{]}ds+\sum_{j=1}^{m}\phi_{s}^{j}dW_{s}^{j},\quad s\in[0,T],\\ \psi_{T}&=-\frac{1}{2}\sqrt{\gamma_{T}}\xi.\end{split} (29)

A pair (ψ,ϕ)(\psi,\phi) with ϕ=(ϕ1,ϕ2,,ϕm)\phi=(\phi^{1},\phi^{2},\ldots,\phi^{m})^{\top} is called a solution to BSDE (29) if

  1. (i)

    ψ\psi is an adapted continuous process with E[sups[0,T]ψs2]<E\left[\sup_{s\in[0,T]}\psi_{s}^{2}\right]<\infty,

  2. (ii)

    ϕ\phi is progressively measurable with 0Tϕs22𝑑s<\int_{0}^{T}\lVert\phi_{s}\rVert_{2}^{2}ds<\infty PP-a.s., and

  3. (iii)

    BSDE (29) is satisfied PP-a.s.

For a solution (K,L)(K,L) of BSDE (27) and a corresponding solution (ψ,ϕ)(\psi,\phi) of BSDE (29), we define θ0=(θs0)s[0,T]\theta^{0}=(\theta^{0}_{s})_{s\in[0,T]} by

θs0=((ρs+μsσs2+σsηsr¯s2)ψs+λsλs+κsγsζs(σs2+2σsηsr¯s+ηs2)Ks+(σs+ηsr¯s)ϕs1+ηs1r¯s2ϕs2)(λs+κs+(σs2+2σsηsr¯s+ηs2)Ks)1,\begin{split}\theta_{s}^{0}&=\Bigg{(}\bigg{(}\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\bigg{)}\psi_{s}+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\sqrt{\gamma_{s}}\zeta_{s}(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})K_{s}\\ &\qquad+(\sigma_{s}+\eta_{s}\overline{r}_{s})\phi^{1}_{s}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\phi^{2}_{s}\Bigg{)}\cdot\left(\lambda_{s}+\kappa_{s}+(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})K_{s}\right)^{-1},\end{split} (30)

for s[0,T]s\in[0,T]. We further introduce for x,dx,d\in\mathbb{R} and s[0,T]s\in[0,T] the SDE

dH^s=H^sd𝒴s+d𝒵s,H^0=dγ0γ0x,d\widehat{H}_{s}^{*}=\widehat{H}_{s}^{*}\,d\mathcal{Y}_{s}+d\mathcal{Z}_{s},\quad\widehat{H}_{0}^{*}=\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x, (31)

where for s[0,T]s\in[0,T]

d𝒴s=(μs2σs28(ρs+μsσs2+σsηsr¯s2)(λsλs+κs+θs))ds+(σs2(σs+ηsr¯s)(λsλs+κs+θs))dWs1ηs1r¯s2(λsλs+κs+θs)dWs2,\begin{split}d\mathcal{Y}_{s}&=\left(\frac{\mu_{s}}{2}-\frac{\sigma_{s}^{2}}{8}-\bigg{(}\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\bigg{)}\bigg{(}\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}+\theta_{s}\bigg{)}\right)ds\\ &\quad+\left(\frac{\sigma_{s}}{2}-(\sigma_{s}+\eta_{s}\overline{r}_{s})\bigg{(}\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}+\theta_{s}\bigg{)}\right)dW_{s}^{1}-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\bigg{(}\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}+\theta_{s}\bigg{)}dW_{s}^{2},\end{split}
d𝒵s=(ρs+μsσs2+σsηsr¯s2)(θs0γsζsλsλs+κs)ds+(σs+ηsr¯s)(θs0γsζsλsλs+κs)dWs1+ηs1r¯s2(θs0γsζsλsλs+κs)dWs2.\begin{split}d\mathcal{Z}_{s}&=\bigg{(}\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\bigg{)}\bigg{(}\theta_{s}^{0}-\sqrt{\gamma_{s}}\zeta_{s}\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\bigg{)}ds\\ &\quad+(\sigma_{s}+\eta_{s}\overline{r}_{s})\bigg{(}\theta_{s}^{0}-\sqrt{\gamma_{s}}\zeta_{s}\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\bigg{)}dW_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\bigg{(}\theta_{s}^{0}-\sqrt{\gamma_{s}}\zeta_{s}\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\bigg{)}dW_{s}^{2}.\end{split}

We will show that the solution H^\widehat{H}^{*} of (31) is the optimal state process in the stochastic control problem to minimize J^\hat{J} of (26). Notice that H^\widehat{H}^{*} can be easily expressed via 𝒴\mathcal{Y} and 𝒵\mathcal{Z} in closed form.

In the next theorem, we summarize consequences from [34] in our setting to obtain a minimizer of J^\hat{J} in (26) and a representation of the minimal costs.

Theorem 3.3.

Assume that there exists ε(0,)\varepsilon\in(0,\infty) such that λ+κε\lambda+\kappa\geq\varepsilon dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. or σ2+2σηr¯+η2ε\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2}\geq\varepsilon dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. We then have:

(i) There exists a unique solution (K,L)(K,L) of BSDE (27). If σ2+2σηr¯+η2ε\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2}\geq\varepsilon dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e., there exists c(0,)c\in(0,\infty) such that P(Ksc for all s[0,T])=1P(K_{s}\geq c\text{ for all }s\in[0,T])=1.

(ii) There exists a unique solution (ψ,ϕ)(\psi,\phi) of BSDE (29).

(iii) Let x,dx,d\in\mathbb{R}, and let H^\widehat{H}^{*} be the solution of SDE (31). Then, u^=(u^s)s[0,T]\hat{u}^{*}=(\hat{u}^{*}_{s})_{s\in[0,T]} defined by

u^s=θsH^sθs0,s[0,T],\hat{u}^{*}_{s}=\theta_{s}\widehat{H}_{s}^{*}-\theta_{s}^{0},\quad s\in[0,T], (32)

is the unique optimal control in 02\mathcal{L}_{0}^{2} for J^\hat{J}, and H^\widehat{H}^{*} is the corresponding state process (i.e., H^=H^u^\widehat{H}^{*}=\widehat{H}^{\hat{u}^{*}}).

(iv) Let x,dx,d\in\mathbb{R}. The costs associated to the optimal control (32) are given by

infu^02J^0(dγ0γ0x,u^)=J^0(dγ0γ0x,u^)=K0(dγ0γ0x)22ψ0(dγ0γ0x)+C0,\begin{split}\inf_{\hat{u}\in\mathcal{L}_{0}^{2}}\hat{J}_{0}\left(\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x,\hat{u}\right)&=\hat{J}_{0}\left(\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x,\hat{u}^{*}\right)\\ &=K_{0}\left(\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x\right)^{2}-2\psi_{0}\left(\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x\right)+C_{0},\end{split}

where

C0=12E0[γTξ2]+E0[0TKsλs2(λs+κs)2γsζs2(σs2+2σsηsr¯s+ηs2)𝑑s]+E0[0Tλsκsλs+κsγsζs2𝑑s]E0[0T(θs0)2(λs+κs+(σs2+2σsηsr¯s+ηs2)Ks)𝑑s]+E0[0T2λsλs+κsγsζsψs(ρs+μsσs2+σsηsr¯s2)𝑑s]+E0[0T2λsλs+κsγsζs(ϕs1(σs+ηsr¯s)+ϕs2ηs1r¯s2)𝑑s].\begin{split}C_{0}&=\frac{1}{2}E_{0}\left[\gamma_{T}\xi^{2}\right]+E_{0}\left[\int_{0}^{T}K_{s}\frac{\lambda_{s}^{2}}{\left(\lambda_{s}+\kappa_{s}\right)^{2}}\gamma_{s}\zeta_{s}^{2}(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})ds\right]\\ &\quad+E_{0}\left[\int_{0}^{T}\frac{\lambda_{s}\kappa_{s}}{\lambda_{s}+\kappa_{s}}\gamma_{s}\zeta_{s}^{2}ds\right]-E_{0}\left[\int_{0}^{T}(\theta_{s}^{0})^{2}(\lambda_{s}+\kappa_{s}+(\sigma_{s}^{2}+2\sigma_{s}\eta_{s}\overline{r}_{s}+\eta_{s}^{2})K_{s})ds\right]\\ &\quad+E_{0}\left[\int_{0}^{T}2\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\sqrt{\gamma_{s}}\zeta_{s}\psi_{s}\bigg{(}\rho_{s}+\mu_{s}-\frac{\sigma_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}}{2}\bigg{)}ds\right]\\ &\quad+E_{0}\left[\int_{0}^{T}2\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\sqrt{\gamma_{s}}\zeta_{s}\left(\phi_{s}^{1}(\sigma_{s}+\eta_{s}\overline{r}_{s})+\phi_{s}^{2}\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\right)ds\right].\end{split} (33)
Proof.

Observe that the problem in Section˜2.2 fits the problem considered in [34, Section 5]. In particular, note that the coefficients in SDE (25) for H^u^\widehat{H}^{\hat{u}} and in the cost functional J^\hat{J} (see (26)) are bounded, and that the inhomogeneities are in 02\mathcal{L}_{0}^{2}. Moreover, we have that 12\frac{1}{2}, λκλ+κ\frac{\lambda\kappa}{\lambda+\kappa}, and λ+κ\lambda+\kappa are nonnegative. Furthermore, the filtration by assumption in this section is generated by the Brownian motion (W1,,Wm)(W^{1},\ldots,W^{m})^{\top}.

(i) If λ+κε\lambda+\kappa\geq\varepsilon, this is an immediate consequence of [34, Theorem 2.1]. In the case σ2+2σηr¯+η2ε\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2}\geq\varepsilon, this is an application of [34, Theorem 2.2].

(ii) This is due to [34, Theorem 5.1].

(iii) The first part of [34, Theorem 5.2] yields the existence of a unique optimal control u^\hat{u}^{*}, which is given in feedback form by the formula u^=θH^u^θ0\hat{u}^{*}=\theta\widehat{H}^{\hat{u}^{*}}-\theta^{0}. We obtain (31) by plugging this into (25).

(iv) The second part of [34, Theorem 5.2] provides us with the optimal costs. ∎

By an application of Corollary˜2.6 and Corollary˜2.4, we obtain a solution to the trade execution problem of Section˜1.

Corollary 3.4.

Assume that there exists ε(0,)\varepsilon\in(0,\infty) such that λ+κε\lambda+\kappa\geq\varepsilon dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. or σ2+2σηr¯+η2ε\sigma^{2}+2\sigma\eta\overline{r}+\eta^{2}\geq\varepsilon dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. Let (K,L)(K,L) be the unique solution of BSDE (27), (ψ,ϕ)(\psi,\phi) the unique solution of BSDE (29), and recall definitions (28) of θ\theta and (30) of θ0\theta^{0}. Let x,dx,d\in\mathbb{R}. Then, X=(Xs)s[0,T]X^{*}=(X^{*}_{s})_{s\in[0-,T]} defined by

X0=x,XT=ξ,Xs=γs12((θs+λsλs+κs1)H^s+γs12ζsλsλs+κsθs0),s[0,T),X^{*}_{0-}\!=x,\,\,\,X^{*}_{T}=\xi,\,\,\,\,X_{s}^{*}=\gamma_{s}^{-\frac{1}{2}}\left(\!\bigg{(}\theta_{s}+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}-1\bigg{)}\widehat{H}_{s}^{*}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}\!-\theta_{s}^{0}\right)\!,\,s\in[0,T),

with H^\widehat{H}^{*} from (31), is the unique (up to dP×ds|[0,T]dP\times ds|_{[0,T]}-null sets) optimal execution strategy in 𝒜0pm(x,d)\mathcal{A}_{0}^{pm}(x,d) for JpmJ^{pm}. The associated costs are given by

infX𝒜0pm(x,d)J0pm(x,d,X)=J0pm(x,d,X)=K0(dγ0γ0x)22ψ0(dγ0γ0x)+C0d22γ0\inf_{X\in\mathcal{A}_{0}^{pm}(x,d)}\!\!J^{pm}_{0}(x,d,X)\!=\!J^{pm}_{0}(x,d,X^{*})\!=\!K_{0}\!\left(\!\frac{d}{\sqrt{\gamma_{0}}}\!-\!\sqrt{\gamma_{0}}x\!\right)^{\!\!2}\!\!-\!2\psi_{0}\!\left(\!\frac{d}{\sqrt{\gamma_{0}}}\!-\!\sqrt{\gamma_{0}}x\!\right)\!+\!C_{0}\!-\!\frac{d^{2}}{2\gamma_{0}}

with C0C_{0} from (33).

Remark 3.5.

(i) Note that BSDE (27) neither contains ξ\xi nor ζ\zeta. In particular, the solution component KK and the process θ\theta from (28) do not depend on the choice of ξ\xi or ζ\zeta (although they depend on the choice of λ\lambda). In contrast, BSDE (29) involves both ξ\xi and ζ\zeta. If ξ=0\xi=0 and at least one of λ\lambda and ζ\zeta is equivalent to 0, we have that (ψ,ϕ)(\psi,\phi) from (29), θ0\theta^{0} from (30), and C0C_{0} from (33) vanish.

(ii) Under the assumptions of Corollary˜3.4 it holds that K012K_{0}\leq\frac{1}{2}. This is a direct consequence of Corollary˜3.4 and (i) above. Indeed, choose ξ=0\xi=0 and ζ=0\zeta=0 (by (i) this choice does not affect KK). Then Corollary˜3.4 and (i) show that J0pm(1,0,X)=K0γ0J^{pm}_{0}(1,0,X^{*})=K_{0}\gamma_{0} for the optimal strategy XX^{*} from Corollary˜3.4. The suboptimal finite variation execution strategy X0=1X_{0-}=1, Xs=0X_{s}=0, s[0,T]s\in[0,T], in 𝒜0fv(1,0)\mathcal{A}_{0}^{fv}(1,0) incurs costs J0pm(1,0,X)=γ02J^{pm}_{0}(1,0,X)=\frac{\gamma_{0}}{2} and hence K012K_{0}\leq\frac{1}{2}.

(iii) Our present setting essentially999The word “essentially” relates to different integrability conditions and to the fact that in [1] the formulation is for a continuous local martingale and a general filtration instead of Brownian motion with Brownian filtration. includes the one in [1], where we have ξ=0\xi=0, λ=0\lambda=0, and η=0\eta=0 (and, therefore, the processes ζ\zeta and r¯\overline{r} are not needed, cf. (3) and (7)). In this subsetting the finite variation control problem associated with (6)–(7) is extended in [1] to a problem where the control XX is a càdlàg semimartingale that acts as integrator in the extended101010The word “extended” relates to the fact that (6) and (7) need to be extended with certain additional terms when allowing for general semimartingale strategies, see [1]. state dynamics of the form (6) and target functional of the form (7). In [1] the existence of an optimal semimartingale strategy as well as the form of the optimal strategy (when it exists) is characterized in terms of a certain process β~\widetilde{\beta}, which is in turn defined via a solution (Y,Z,M)(Y,Z,M^{\perp}) to a certain quadratic BSDE (see (3.2) in [1]). It is worth noting that, in the subsetting with ξ=0\xi=0, λ=0\lambda=0, and η=0\eta=0, all formulas in this section greatly simplify and, in particular, BSDE (27) above is equivalent111111For the sake of fair comparison, we consider the subsetting in [1] where the filtration is generated by (W1,,Wm)(W^{1},\ldots,W^{m})^{\top} and the continuous local martingale MM is W1W^{1}. to BSDE (3.2) in [1]. The relation is Y=KY=K, Z=L1Z=L^{1}, dMs=j=2mLsjdWsjdM^{\perp}_{s}=\sum_{j=2}^{m}L^{j}_{s}\,dW^{j}_{s}. Further, in that subsetting, our process θ\theta from (28) reduces to the above-mentioned process β~\widetilde{\beta} (see (3.5) in [1]), while (ψ,ϕ)(\psi,\phi) from (29), θ0\theta^{0} from (30), and C0C_{0} from (33) vanish.

(iv) It is also instructive to compare Corollary˜3.4 above, where we obtain that the extended to 𝒜0pm(x,d)\mathcal{A}_{0}^{pm}(x,d) control problem always admits a minimizer, with Theorem 3.4 in [1], where it turns out that an optimal semimartingale strategy can fail to exist. See the discussion in the end of Section˜4.3 for a specific example.

On the continuity of optimal position paths: In the setting of [37] optimal position paths XX^{*} exhibit jumps (so-called block trades) at times 0 and TT but are continuous on the interior (0,T)(0,T) (see also Section˜4.1 below). An interesting question is whether the continuity on (0,T)(0,T) prevails in the generalized setting considered in this paper. This is not reasonable to expect when we have the risk term with a “sufficiently irregular” process ζ\zeta. And, indeed, we see that the continuity of XX^{*} on (0,T)(0,T) can fail in ˜4.1 below (this is discussed in Remark˜4.2). More interestingly, such a continuity can already fail even without the risk term (i.e. λ=0\lambda=0) and with terminal target ξ=0\xi=0. Indeed, consider the setting with σ=0\sigma=0, λ=0\lambda=0, ξ=0\xi=0 and non-diffusive resilience process RR given by Rs=ρsR_{s}=\rho s (with ρ\rho being a deterministic constant). Then it follows from [1, Example 6.2] that continuity of the price impact process γ\gamma is not sufficient for continuity of optimal position paths XX^{*} on (0,T)(0,T). It is shown that if the paths of γ\gamma are absolutely continuous, then a jump of the weak derivative of γ\gamma on (0,T)(0,T) already causes XX^{*} to jump on (0,T)(0,T). Moreover, it is possible that the random terminal target position ξ\xi causes the optimal position path XX^{*} to jump in (0,T)(0,T) with all other input processes being continuous. We present an example for this phenomenon in Section˜4.2.

A way to obtain sufficient conditions for the continuity of XX^{*} on (0,T)(0,T) consists of combining Corollary˜3.4 with path regularity results for BSDEs. Indeed, if the coefficient processes ρ,μ,σ,η,r¯,λ,ζ\rho,\mu,\sigma,\eta,\overline{r},\lambda,\zeta are continuous and if one can ensure that the solution components L1,L2L^{1},L^{2} and ϕ1,ϕ2\phi^{1},\phi^{2} (which correspond to the martingale representation part of the solution) of the BSDE (27) resp. (29) have continuous sample paths, then Corollary˜3.4 ensures that XX^{*} also has continuous sample paths on (0,T)(0,T). Results that guarantee continuity of BSDE solutions in a Markovian framework, including the quadratic case, can for example be found in [32].

4 Examples

In this section we apply the results from the preceding sections in specific case studies.

4.1 The Obizhaeva-Wang model with random targets

The models developed by Obizhaeva and Wang [37] can be considered as special cases of the model set up in Section˜1. Indeed, we obtain the problem of [37, Section 6] by setting μ0\mu\equiv 0, σ0\sigma\equiv 0, η0\eta\equiv 0, r¯0\overline{r}\equiv 0, λ0\lambda\equiv 0 and choosing ρ(0,)\rho\in(0,\infty) and ξ\xi\in\mathbb{R} as deterministic constants.

Example 4.1.

In this example we apply our results (in particular, Corollary˜3.4) and provide closed-form solutions (see (38) below) for optimal progressively measurable execution strategies in versions of these problems which allow for general random terminal targets ξ\xi and general running targets ζ\zeta.

To this end let x,dx,d\in\mathbb{R}. Suppose that μ0\mu\equiv 0, σ0\sigma\equiv 0, η0\eta\equiv 0, and r¯0\overline{r}\equiv 0. Furthermore, assume that ρ(0,)\rho\in(0,\infty) and λ[0,)\lambda\in[0,\infty) are deterministic constants. We take some ξ\xi and ζ\zeta as specified in Section˜1.1 (in particular, see (5)). Note that the conditions of ˜3.3 and Corollary˜3.4 hold true, and that γs=γ0\gamma_{s}=\gamma_{0} for all s[0,T]s\in[0,T]. In the current setting, BSDE (27) reads

dKs=(ρ2ρ+λKs2+2λρρ+λKsλρρ+λ)ds+j=1mLsjdWsj,s[0,T],KT=12.\begin{split}dK_{s}&=\bigg{(}\frac{\rho^{2}}{\rho+\lambda}K_{s}^{2}+\frac{2\lambda\rho}{\rho+\lambda}K_{s}-\frac{\lambda\rho}{\rho+\lambda}\bigg{)}ds+\sum_{j=1}^{m}L_{s}^{j}dW_{s}^{j},\quad s\in[0,T],\quad K_{T}=\frac{1}{2}.\end{split} (34)

By ˜3.3, there exists a unique solution (K,L)(K,L). Since the driver and the terminal condition in (34) are deterministic, we obtain that L0L\equiv 0, and hence (34) is in fact a scalar Riccati ODE with constant coefficients. Such an equation can be solved explicitly, and in our situation we obtain in the case λ>0\lambda>0 that

Ks=12λtanh(λρ(Ts)λ+ρ)+λ(ρ+λ)(ρ2+λ)tanh(λρ(Ts)λ+ρ)+λ(ρ+λ),s[0,T],K_{s}=\frac{1}{2}\frac{\lambda\tanh\left(\frac{\sqrt{\lambda}\rho(T-s)}{\sqrt{\lambda+\rho}}\right)+\sqrt{\lambda(\rho+\lambda)}}{(\frac{\rho}{2}+\lambda)\tanh\left(\frac{\sqrt{\lambda}\rho(T-s)}{\sqrt{\lambda+\rho}}\right)+\sqrt{\lambda(\rho+\lambda)}},\quad s\in[0,T],

and in the case λ=0\lambda=0 that

Ks=12+(Ts)ρ,s[0,T].K_{s}=\frac{1}{2+(T-s)\rho},\quad s\in[0,T]. (35)

The process θ\theta from (28) here is given by θs=ρλ+ρKs\theta_{s}=\frac{\rho}{\lambda+\rho}K_{s}, s[0,T]s\in[0,T]. BSDE (29) becomes

dψs=(ρλλ+ρ+ρθs)ψsds+ρλλ+ργ0ζs(1Ks)ds+j=1mϕsjdWsj,s[0,T],ψT=12γ0ξ.d\psi_{s}=\!\left(\!\frac{\rho\lambda}{\lambda+\rho}\!+\!\rho\theta_{s}\!\right)\!\psi_{s}ds+\frac{\rho\lambda}{\lambda+\rho}\sqrt{\gamma_{0}}\zeta_{s}(1-K_{s})ds+\sum_{j=1}^{m}\!\phi_{s}^{j}dW_{s}^{j},\,\,\,s\in[0,T],\,\,\,\psi_{T}=\!-\frac{1}{2}\sqrt{\gamma_{0}}\xi. (36)

Again, by ˜3.3, there exists a unique solution (ψ,ϕ)(\psi,\phi). The solution component ψ\psi is given by

ψs=Γs1γ0(12ΓTEs[ξ]ρλλ+ρEs[sTΓr(1Kr)ζr𝑑r]),s[0,T],\begin{split}\psi_{s}&=\Gamma_{s}^{-1}\sqrt{\gamma_{0}}\left(-\frac{1}{2}\Gamma_{T}E_{s}[\xi]-\frac{\rho\lambda}{\lambda+\rho}E_{s}\left[\int_{s}^{T}\Gamma_{r}(1-K_{r})\zeta_{r}\,dr\right]\right),\quad s\in[0,T],\end{split}

where

Γs=exp(ρ0s(λλ+ρ+θr)𝑑r)=exp(ρλ+ρ(λs+ρ0sKr𝑑r)),s[0,T].\Gamma_{s}=\exp\left(-\rho\int_{0}^{s}\left(\frac{\lambda}{\lambda+\rho}+\theta_{r}\right)dr\right)=\exp\left(-\frac{\rho}{\lambda+\rho}\left(\lambda s+\rho\int_{0}^{s}K_{r}dr\right)\right),\quad s\in[0,T]. (37)

It holds for the process in (30) that θs0=ρλ+ρψs\theta_{s}^{0}=\frac{\rho}{\lambda+\rho}\psi_{s}, s[0,T]s\in[0,T]. Further, SDE (31) reads

dH^s=ρ(λλ+ρ+θs)H^sds+ρ(θs0γ0ζsλλ+ρ)ds,s[0,T],H^0=dγ0γ0x,d\widehat{H}_{s}^{*}=\!-\rho\left(\!\frac{\lambda}{\lambda+\rho}+\theta_{s}\!\right)\widehat{H}_{s}^{*}ds+\rho\left(\!\theta_{s}^{0}-\!\sqrt{\gamma_{0}}\zeta_{s}\frac{\lambda}{\lambda+\rho}\right)\!ds,\,\,\,s\in[0,T],\,\,\,\widehat{H}_{0}^{*}=\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x,

and has solution

H^s=Γs(dγ0γ0x+ρ0sΓr1(θr0γ0ζrλλ+ρ)𝑑r),s[0,T],\widehat{H}_{s}^{*}=\Gamma_{s}\left(\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x+\rho\int_{0}^{s}\Gamma_{r}^{-1}\left(\theta_{r}^{0}-\sqrt{\gamma_{0}}\zeta_{r}\frac{\lambda}{\lambda+\rho}\right)dr\right),\quad s\in[0,T],

with Γ\Gamma from (37). It then follows from Corollary˜3.4 that X=(Xs)s[0,T]X^{*}=(X^{*}_{s})_{s\in[0-,T]} defined by X0=xX_{0-}^{*}=x, XT=ξX_{T}^{*}=\xi, and, for s[0,T)s\in[0,T),

Xs=γ012((θsρλ+ρ)H^sθs0)+ζsλλ+ρ=ρλ+ρ(1Ks)Γs(xdγ0+ρλ+ρ0sΓr1(λζrργ0ψr)𝑑r)+ρλ+ρ(λρζs1γ0ψs)\begin{split}&X_{s}^{*}=\gamma_{0}^{-\frac{1}{2}}\left(\left(\theta_{s}-\frac{\rho}{\lambda+\rho}\right)\widehat{H}_{s}^{*}-\theta_{s}^{0}\right)+\zeta_{s}\frac{\lambda}{\lambda+\rho}\\ &=\frac{\rho}{\lambda+\rho}(1-K_{s})\Gamma_{s}\!\left(\!x-\!\frac{d}{\gamma_{0}}\!+\!\frac{\rho}{\lambda+\rho}\int_{0}^{s}\!\Gamma_{r}^{-1}\!\left(\!\lambda\zeta_{r}\!-\!\frac{\rho}{\sqrt{\gamma_{0}}}\psi_{r}\!\right)dr\!\right)\!+\!\frac{\rho}{\lambda+\rho}\left(\frac{\lambda}{\rho}\zeta_{s}\!-\!\frac{1}{\sqrt{\gamma_{0}}}\psi_{s}\!\right)\end{split} (38)

is the (up to dP×ds|[0,T]dP\!\times\!ds|_{[0,T]}-null sets unique) execution strategy in 𝒜0pm(x,d)\mathcal{A}_{0}^{pm}(x,d) that minimizes JpmJ^{pm}.

Remark 4.2.

From ˜4.1 we see that discontinuities of the target process ζ\zeta can cause jumps of the optimal position path XX^{*} in (0,T)(0,T). Indeed, as θ\theta, θ0\theta^{0} and H^\widehat{H}^{*} are continuous, it follows from (38) that, in the case λ>0\lambda>0, paths of the optimal strategy XX^{*} inherit discontinuities from ζ\zeta on (0,T)(0,T) (in particular, XX^{*} jumps on (0,T)(0,T) whenever ζ\zeta does).

In the next example we study the case λ0\lambda\equiv 0 in more detail.

Example 4.3.

In the setting of the previous example suppose that λ0\lambda\equiv 0. If the terminal target ξ\xi\in\mathbb{R} is a deterministic constant, then it follows from [37, Proposition 3] that the optimal finite variation execution strategy is given by

Xs=(xξdγ0)1+(Ts)ρ2+Tρ+ξ,s[0,T).X_{s}^{*}=\left(x-\xi-\frac{d}{\gamma_{0}}\right)\frac{1+(T-s)\rho}{2+T\rho}+\xi,\quad s\in[0,T). (39)

So the optimal strategy consists of potential block trades (jumps of XX^{*}) at times 0 and TT and a continuous linear trading program on [0,T)[0,T). In the following we analyze how this structure changes as we allow for a random terminal target ξ\xi.

First recall that the solution of BSDE (34) is given in this case by (35). It follows that Γ\Gamma from (37) simplifies to Γs=2+(Ts)ρ2+Tρ\Gamma_{s}=\frac{2+(T-s)\rho}{2+T\rho}, s[0,T]s\in[0,T]. For the solution component ψ\psi of BSDE (36), we thus obtain

ψs=γ02+(Ts)ρEs[ξ],s[0,T].\psi_{s}=-\frac{\sqrt{\gamma_{0}}}{2+(T-s)\rho}E_{s}[\xi],\quad s\in[0,T].

The optimal strategy from (38) on [0,T)[0,T) becomes, for s[0,T)s\in[0,T),

Xs=(1Ks)Γs(xdγ0ρ0sΓr11γ0ψr𝑑r)1γ0ψs=(xdγ0)1+(Ts)ρ2+Tρ+ρ(1+(Ts)ρ)0sEr[ξ](2+(Tr)ρ)2𝑑r+Es[ξ]2+(Ts)ρ.\begin{split}X_{s}^{*}&=(1-K_{s})\Gamma_{s}\left(x-\frac{d}{\gamma_{0}}-\rho\int_{0}^{s}\Gamma_{r}^{-1}\frac{1}{\sqrt{\gamma_{0}}}\psi_{r}dr\right)-\frac{1}{\sqrt{\gamma_{0}}}\psi_{s}\\ &=\left(x-\frac{d}{\gamma_{0}}\right)\frac{1+(T-s)\rho}{2+T\rho}+\rho(1+(T-s)\rho)\int_{0}^{s}\!\frac{E_{r}[\xi]}{(2+(T-r)\rho)^{2}}dr+\frac{E_{s}[\xi]}{2+(T-s)\rho}.\end{split} (40)

Integration by parts implies that (note that (Er[ξ])r[0,T](E_{r}[\xi])_{r\in[0,T]} is a continuous martingale)

0sEr[ξ](2+(Tr)ρ)2𝑑r=0sEr[ξ]𝑑1ρ(2+(Tr)ρ)=Es[ξ]ρ(2+(Ts)ρ)E0[ξ]ρ(2+Tρ)0s1ρ(2+(Tr)ρ)𝑑Er[ξ],s[0,T).\begin{split}&\int_{0}^{s}\frac{E_{r}[\xi]}{(2+(T-r)\rho)^{2}}dr=\int_{0}^{s}E_{r}[\xi]d\frac{1}{\rho(2+(T-r)\rho)}\\ &=\frac{E_{s}[\xi]}{\rho(2+(T-s)\rho)}-\frac{E_{0}[\xi]}{\rho(2+T\rho)}-\int_{0}^{s}\frac{1}{\rho(2+(T-r)\rho)}dE_{r}[\xi],\quad s\in[0,T).\end{split}

Substituting this into (40) yields, for s[0,T)s\in[0,T),

Xs\displaystyle X_{s}^{*} =(xE0[ξ]dγ0)1+(Ts)ρ2+Tρ+Es[ξ]0s1+(Ts)ρ2+(Tr)ρ𝑑Er[ξ]\displaystyle=\left(x-E_{0}[\xi]-\frac{d}{\gamma_{0}}\right)\frac{1+(T-s)\rho}{2+T\rho}+E_{s}[\xi]-\int_{0}^{s}\frac{1+(T-s)\rho}{2+(T-r)\rho}dE_{r}[\xi]
=(xE0[ξ]dγ0)1+(Ts)ρ2+Tρ+E0[ξ]+0s(11+(Ts)ρ2+(Tr)ρ)𝑑Er[ξ].\displaystyle=\left(x-E_{0}[\xi]-\frac{d}{\gamma_{0}}\right)\frac{1+(T-s)\rho}{2+T\rho}+E_{0}[\xi]+\int_{0}^{s}\left(1-\frac{1+(T-s)\rho}{2+(T-r)\rho}\right)dE_{r}[\xi].

We, finally, obtain the alternative representation

Xs=(xE0[ξ]dγ0)1+(Ts)ρ2+Tρ+E0[ξ]+0s1+(sr)ρ2+(Tr)ρ𝑑Er[ξ],s[0,T),X_{s}^{*}=\left(x-E_{0}[\xi]-\frac{d}{\gamma_{0}}\right)\frac{1+(T-s)\rho}{2+T\rho}+E_{0}[\xi]+\int_{0}^{s}\frac{1+(s-r)\rho}{2+(T-r)\rho}dE_{r}[\xi],\quad s\in[0,T),

for (40). We see that this optimal strategy X𝒜0pm(x,d)X^{*}\in\mathcal{A}_{0}^{pm}(x,d) consists of two additive parts: The first part exactly corresponds to the optimal deterministic strategy in (39) where the deterministic terminal target is replaced by the expected terminal target E0[ξ]E_{0}[\xi]. The second part represents fluctuations around this deterministic strategy which incorporate updates about the random terminal target ξ\xi. Note that this stochastic integral vanishes in expectation, although this is not a martingale (indeed, the time ss is not only the upper bound of integration but also appears in the integrand).

4.2 A discontinuous optimal position path for continuous inputs

We now show that the optimal strategy can have jumps inside (0,T)(0,T) even if all input processes, including ζ\zeta, are continuous. To this end, let x,dx,d\in\mathbb{R}. Take λ0\lambda\equiv 0, ζ0\zeta\equiv 0, η0\eta\equiv 0, r¯0\overline{r}\equiv 0, and μ0\mu\equiv 0, and assume that σ(0,)\sigma\in(0,\infty) and ρ(σ22,)\rho\in(\frac{\sigma^{2}}{2},\infty) are deterministic constants. Moreover, we will later consider an appropriate random terminal target ξ\xi, satisfying the assumptions of Section˜1.1, to produce a jump of the optimal strategy.

Note that the conditions of ˜3.3 and Corollary˜3.4 hold true. In particular, there exists a unique solution (K,L)(K,L) of BSDE (27), and it is given by (compare also with [1, Section 5.2]) L0L\equiv 0 and

Ks=ρσ22σ2𝒲(ρσ22σ2ec0ρ2σ2s)1,s[0,T],K_{s}=\frac{\rho-\frac{\sigma^{2}}{2}}{\sigma^{2}}\mathcal{W}\left(\frac{\rho-\frac{\sigma^{2}}{2}}{\sigma^{2}}e^{c_{0}-\frac{\rho^{2}}{\sigma^{2}}s}\right)^{-1},\quad s\in[0,T],

where 𝒲\mathcal{W} denotes the Lambert WW function and c0=ln(2)+1σ2(2ρσ2+ρ2T)c_{0}=\ln(2)+\frac{1}{\sigma^{2}}(2\rho-\sigma^{2}+\rho^{2}T). The process θ\theta from (28) becomes

θs=ρKsρσ22+σ2Ks,s[0,T],\theta_{s}=\frac{\rho K_{s}}{\rho-\frac{\sigma^{2}}{2}+\sigma^{2}K_{s}},\quad s\in[0,T],

and both θ\theta and KK are deterministic, increasing, continuous, (0,1/2](0,1/2]-valued functions.

For some t0(0,T)t_{0}\in(0,T), let

ξ=2γ012(σt0TΓsθs𝑑s+t0TΓs𝑑Ws1)exp(σ2WT1+38σ2T+(ρσ22)0Tθs𝑑s),\xi=-2\gamma_{0}^{-\frac{1}{2}}\left(\sigma\int_{t_{0}}^{T}\Gamma_{s}\theta_{s}ds+\int_{t_{0}}^{T}\Gamma_{s}dW_{s}^{1}\right)\,\exp\left(-\frac{\sigma}{2}W^{1}_{T}+\frac{3}{8}\sigma^{2}T+\left(\rho-\frac{\sigma^{2}}{2}\right)\int_{0}^{T}\theta_{s}ds\right),

where Γt=exp(σ28t(ρσ22)0tθs𝑑s)\Gamma_{t}=\exp(-\frac{\sigma^{2}}{8}t-(\rho-\frac{\sigma^{2}}{2})\int_{0}^{t}\theta_{s}ds), t[0,T]t\in[0,T]. Note that ξ\xi is T\mathcal{F}_{T}-measurable and that E[γTξ2]<E[\gamma_{T}\xi^{2}]<\infty. The terminal target ξ\xi here is defined in such a way that the unique solution (ψ,ϕ)(\psi,\phi) of BSDE (29) (cf. ˜3.3) is given by ϕ1=1[t0,T]\phi^{1}=1_{[t_{0},T]}, ϕj0\phi^{j}\equiv 0, j{2,,m}j\in\{2,\ldots,m\}, and

ψt={0,0t<t0,Γt1(σt0tΓsθs𝑑s+t0tΓs𝑑Ws1),t0tT.\psi_{t}=\begin{cases}0,&0\leq t<t_{0},\\ \Gamma_{t}^{-1}\left(\sigma\int_{t_{0}}^{t}\Gamma_{s}\theta_{s}ds+\int_{t_{0}}^{t}\Gamma_{s}dW_{s}^{1}\right),&t_{0}\leq t\leq T.\end{cases}

It follows for the process in (30) that

θt0={0,0t<t0,(ρσ22)ψt+σρσ22+σ2Kt,t0tT.\theta_{t}^{0}=\begin{cases}0,&0\leq t<t_{0},\\ \frac{\left(\rho-\frac{\sigma^{2}}{2}\right)\psi_{t}+\sigma}{\rho-\frac{\sigma^{2}}{2}+\sigma^{2}K_{t}},&t_{0}\leq t\leq T.\end{cases}

We thus have that

Δθt00=σρσ22+σ2Kt0>0.\Delta\theta_{t_{0}}^{0}=\frac{\sigma}{\rho-\frac{\sigma^{2}}{2}+\sigma^{2}K_{t_{0}}}>0.

From Corollary˜3.4 we obtain existence of a unique optimal strategy XX^{*} and that Xs=γs12((θs1)H^sθs0)X_{s}^{*}=\gamma_{s}^{-\frac{1}{2}}((\theta_{s}-1)\widehat{H}_{s}^{*}-\theta_{s}^{0}), s(0,T)s\in(0,T). Since γ\gamma, θ\theta, and H^s\widehat{H}_{s}^{*} (see (31)) are continuous and Δθt00>0\Delta\theta_{t_{0}}^{0}>0, it holds that ΔXt0=γt012Δθt00<0\Delta X_{t_{0}}^{*}=-\gamma_{t_{0}}^{-\frac{1}{2}}\Delta\theta_{t_{0}}^{0}<0. Hence, the optimal strategy has a jump at t0(0,T)t_{0}\in(0,T).

4.3 An example where JfvJ^{fv} does not admit a minimizer

Let x,dx,d\in\mathbb{R} with xdγ0x\neq\frac{d}{\gamma_{0}}. Suppose that σ0\sigma\equiv 0, η0\eta\equiv 0, λ0\lambda\equiv 0, r¯0\overline{r}\equiv 0, ζ0\zeta\equiv 0, ξ=0\xi=0. Choose μ\mu to be a bounded deterministic càdlàg function such that there exists δ(0,T)\delta\in(0,T) with μ\mu having infinite variation on [0,Tδ][0,T-\delta], and take ρ{0}\rho\in\mathbb{R}\setminus\{0\} such that there exists ε>0\varepsilon>0 with 2ρ+με2\rho+\mu\geq\varepsilon. Note that this corresponds to the setting in [1, Example 6.4]. Moreover, observe that the conditions of Corollary˜3.4 are satisfied. In the current setting, BSDE (27) becomes

dKs=(μsKs+2(ρ+μs)2Ks22ρ+μs)ds+j=1mLsjdWsj,s[0,T],KT=12.\begin{split}dK_{s}&=\left(-\mu_{s}K_{s}+\frac{2(\rho+\mu_{s})^{2}K_{s}^{2}}{2\rho+\mu_{s}}\right)ds+\sum_{j=1}^{m}L_{s}^{j}dW_{s}^{j},\quad s\in[0,T],\quad K_{T}=\frac{1}{2}.\end{split}

Its solution is given by (K,0)(K,0), where (see also YY in [1, Section 6])

Ks=esTμr𝑑r(sT2(ρ+μr)22ρ+μrerTμl𝑑l𝑑r+2)1,s[0,T],K_{s}=e^{\int_{s}^{T}\mu_{r}dr}\left(\int_{s}^{T}\frac{2(\rho+\mu_{r})^{2}}{2\rho+\mu_{r}}e^{\int_{r}^{T}\mu_{l}dl}dr+2\right)^{-1},\quad s\in[0,T],

is a deterministic continuous function of finite variation. We have that

θs=2(ρ+μs)2ρ+μsKs,s[0,T],\theta_{s}=\frac{2(\rho+\mu_{s})}{2\rho+\mu_{s}}K_{s},\quad s\in[0,T],

which is the same as β~\widetilde{\beta} in [1, Example 6.4]. The solution of BSDE (29) is given by (ψ,ϕ)=(0,0)(\psi,\phi)=(0,0), and it holds θ00\theta^{0}\equiv 0. Furthermore, (31) reads

dH^s=(μs2(ρ+μs)θs)H^sds,s[0,T],H^0=dγ0γ0x,d\widehat{H}^{*}_{s}=\left(\frac{\mu_{s}}{2}-(\rho+\mu_{s})\theta_{s}\right)\widehat{H}_{s}^{*}ds,\quad s\in[0,T],\quad\widehat{H}_{0}^{*}=\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x,

and is solved by the continuous deterministic finite-variation function

H^s=(dγ0γ0x)exp(0s(μr2(ρ+μr)θr)𝑑r),s[0,T],\widehat{H}_{s}^{*}=\left(\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x\right)\exp\left(\int_{0}^{s}\left(\frac{\mu_{r}}{2}-(\rho+\mu_{r})\theta_{r}\right)dr\right),\quad s\in[0,T],

which is nonvanishing due to our assumption xdγ0x\neq\frac{d}{\gamma_{0}}.121212At this point it is easy to explain why we exclude the case x=dγ0x=\frac{d}{\gamma_{0}} in this example. In the case x=dγ0x=\frac{d}{\gamma_{0}} we get that H^0\widehat{H}^{*}\equiv 0 and then the optimal strategy is to close the position immediately, i.e., X0=xX^{*}_{0-}=x, Xs=0X^{*}_{s}=0, s[0,T]s\in[0,T], which is always a finite-variation strategy. By Corollary˜3.4, there exists a (up to dP×ds|[0,T]dP\times ds|_{[0,T]}-null sets) unique minimizer X=(Xs)s[0,T]X^{*}=(X^{*}_{s})_{s\in[0-,T]} of JpmJ^{pm} in 𝒜0pm(x,d)\mathcal{A}_{0}^{pm}(x,d), namely

X0=x,XT=0,Xs=γs12(θs1)H^s,s[0,T).\begin{split}&X^{*}_{0-}=x,\quad X^{*}_{T}=0,\quad X_{s}^{*}=\gamma_{s}^{-\frac{1}{2}}\left(\theta_{s}-1\right)\widehat{H}_{s}^{*},\quad s\in[0,T).\end{split}

Assume by contradiction that there exists a minimizer X0=(Xs0)s[0,T]X^{0}=(X^{0}_{s})_{s\in[0-,T]} of JfvJ^{fv} in 𝒜0fv(x,d)\mathcal{A}_{0}^{fv}(x,d). We know from Corollary˜2.3 that X0X^{0} is then also a minimizer of JpmJ^{pm} in 𝒜0pm(x,d)\mathcal{A}_{0}^{pm}(x,d). It follows that X0=XX^{0}=X^{*} dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. Since H^\widehat{H}^{*} is nowhere 0, we obtain that

1+γ12X0H^=θdP×ds|[0,T]-a.e.1+\frac{\gamma^{\frac{1}{2}}X^{0}}{\widehat{H}^{*}}=\theta\quad dP\times ds|_{[0,T]}\text{-a.e.} (41)

Observe that the left-hand side is a process of finite variation. On the other hand, our assumption on μ\mu easily yields that θ\theta has infinite variation. This contradiction proves that in the setting of this example, JfvJ^{fv} does not admit a minimizer in 𝒜0fv(x,d)\mathcal{A}_{0}^{fv}(x,d).

We can say even more: In this example there does not exist a semimartingale optimal strategy.131313Under a semimartingale strategy we formally understand a semimartingale that is an element of 𝒜0pm(x,d)\mathcal{A}_{0}^{pm}(x,d). Indeed, if we had a semimartingale X0X^{0} as a minimizer, we would still get (41) (with a semimartingale X0X^{0}). The left-hand side would then be a semimartingale. On the other hand, it is shown in [1, Example 6.4] that there does not exist a semimartingale β\beta such that β=θ\beta=\theta dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. Thus, the cost functional does not have a minimizer in the set of semimartingales, but we are now able to find a minimizer in the set of progressively measurable execution strategies.

4.4 An example with a diffusive resilience

As already mentioned in the introduction, the literature on optimal trade execution in Obizhaeva-Wang type models typically assumes that RR is an increasing process. In [1] and [3] RR is allowed to have finite variation. Now we consider an example with a truly diffusive RR.

Let x,dx,d\in\mathbb{R} with xdγ0x\neq\frac{d}{\gamma_{0}}. Let ξ=0\xi=0, λ0\lambda\equiv 0, ζ0\zeta\equiv 0, and μ0\mu\equiv 0. Suppose that r¯[1,1]\overline{r}\in[-1,1] and η,ρ,σ\eta,\rho,\sigma\in\mathbb{R} are deterministic constants such that κ=12(2ρσ2η22σηr¯)>0\kappa=\frac{1}{2}(2\rho-\sigma^{2}-\eta^{2}-2\sigma\eta\overline{r})>0 and σ2+η2+2σηr¯>0\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}>0 (in particular, we thus need ρ>0\rho>0). Note that the assumptions of Corollary˜3.4 are satisfied. We moreover remark that the subsetting where η0\eta\equiv 0 corresponds to the setting in [1, Section 5.2]. That means, the difference to [1, Section 5.2] is that we now consider a more general resilience. The Riccati-BSDE (27) becomes

dKs=(ρKs+(σ+ηr¯)Ls1+η1r¯2Ls2)2(σ2+η2+2σηr¯)Ks+κdsσLs1ds+j=1mLsjdWsj,s[0,T],KT=12.dK_{s}=\frac{(\rho K_{s}+(\sigma+\eta\overline{r})L_{s}^{1}+\eta\sqrt{1-\overline{r}^{2}}L_{s}^{2})^{2}}{(\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r})K_{s}+\kappa}ds-\sigma L_{s}^{1}ds+\sum_{j=1}^{m}\!L_{s}^{j}dW_{s}^{j},\,\,\,s\in[0,T],\,\,\,K_{T}=\frac{1}{2}.

This has solution (K,L)=(K,0)(K,L)=(K,0) with

Ks=κσ2+η2+2σηr¯𝒲(κσ2+η2+2σηr¯exp(cρ2sσ2+η2+2σηr¯))1,s[0,T],K_{s}=\frac{\kappa}{\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}}\mathcal{W}\left(\frac{\kappa}{\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}}\exp\left(c-\frac{\rho^{2}s}{\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}}\right)\right)^{-1},\quad s\in[0,T],

and c=ln(2)+2κ+ρ2Tσ2+η2+2σηr¯c=\ln(2)+\frac{2\kappa+\rho^{2}T}{\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}} (compare also with [1, Section 5.2]). We further have that θs=ρKs(σ2+η2+2σηr¯)Ks+κ\theta_{s}=\frac{\rho K_{s}}{(\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r})K_{s}+\kappa}, s[0,T]s\in[0,T]. Observe that (ψ,ϕ)=(0,0)(\psi,\phi)=(0,0) is the solution of (29) in the present setting and that θ00\theta^{0}\equiv 0 in (30). Moreover, we have that SDE (31) reads

dH^s=(σ28(ρσ2+σηr¯2)θs)H^sds+(σ2(σ+ηr¯)θs)H^sdWs1η1r¯2θsH^sdWs2d\widehat{H}_{s}^{*}=\!\left(\!-\frac{\sigma^{2}}{8}\!-\!\left(\!\rho-\frac{\sigma^{2}+\sigma\eta\overline{r}}{2}\right)\theta_{s}\!\right)\!\widehat{H}_{s}^{*}ds+\left(\frac{\sigma}{2}\!-\!(\sigma+\eta\overline{r})\theta_{s}\right)\!\widehat{H}_{s}^{*}dW_{s}^{1}-\eta\sqrt{1-\overline{r}^{2}}\theta_{s}\widehat{H}_{s}^{*}dW_{s}^{2}

for s[0,T]s\in[0,T], with start in H^0=dγ0γ0x\widehat{H}_{0}^{*}=\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x; hence,

H^s=(dγ0γ0x)exp(σ2s4(ρσ2σηr¯)0sθr𝑑rσ2+η2+2σηr¯20sθr2𝑑r)exp(σ2Ws1(σ+ηr¯)0sθrdWr1η1r¯20sθrdWr2),s[0,T].\begin{split}\widehat{H}_{s}^{*}&=\left(\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x\right)\exp\!\left(-\frac{\sigma^{2}s}{4}-(\rho-\sigma^{2}-\sigma\eta\overline{r})\int_{0}^{s}\theta_{r}dr-\frac{\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}}{2}\int_{0}^{s}\theta_{r}^{2}dr\right)\\ &\quad\cdot\exp\left(\frac{\sigma}{2}W_{s}^{1}-(\sigma+\eta\overline{r})\int_{0}^{s}\theta_{r}dW_{r}^{1}-\eta\sqrt{1-\overline{r}^{2}}\int_{0}^{s}\theta_{r}dW_{r}^{2}\right),\quad s\in[0,T].\end{split}

It follows from Corollary˜3.4 that for s[0,T)s\in[0,T) the optimal execution strategy is given by

Xs=(xdγ0)(1θs)exp((ρσ2σηr¯)0sθr𝑑rσ2+η2+2σηr¯20sθr2𝑑r)exp((σ+ηr¯)0sθr𝑑Wr1η1r¯20sθr𝑑Wr2).\begin{split}X_{s}^{*}&=\left(x-\frac{d}{\gamma_{0}}\right)(1-\theta_{s})\exp\!\left(-(\rho-\sigma^{2}-\sigma\eta\overline{r})\int_{0}^{s}\theta_{r}dr-\frac{\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}}{2}\int_{0}^{s}\theta_{r}^{2}dr\right)\\ &\quad\cdot\exp\left(-(\sigma+\eta\overline{r})\int_{0}^{s}\theta_{r}dW_{r}^{1}-\eta\sqrt{1-\overline{r}^{2}}\int_{0}^{s}\theta_{r}dW_{r}^{2}\right).\end{split}

We can show that KK and θ\theta both are continuous, deterministic, increasing, (0,1/2](0,1/2]-valued functions of finite variation. Since θ<1\theta<1, the optimal strategy on [0,T)[0,T) always has the same sign as xdγ0x-\frac{d}{\gamma_{0}}. Moreover, the optimal strategy is stochastic and has infinite variation, as in [1, Section 5.2]. In contrast to [1, Section 5.2], where the price impact always has infinite variation, we can here set σ0\sigma\equiv 0 for a choice of η2(0,2ρ)\eta^{2}\in(0,2\rho). In this case, the price impact γγ0\gamma\equiv\gamma_{0} is a deterministic constant, yet the optimal strategy has infinite variation (due to the infinite variation in the resilience RR).

Observe furthermore that by making use of η\eta and r¯\overline{r}, we can choose the parameters in the current setting in such a way that κ>0\kappa>0 and σ2+η2+2σηr¯>0\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}>0 are satisfied, but condition (3.1) in [1], i.e., 2ρσ2>02\rho-\sigma^{2}>0, is violated.

With regard to Section˜4.3 we remark that in both sections there does not exist an optimal strategy in 𝒜0fv(x,d)\mathcal{A}_{0}^{fv}(x,d), but opposed to Section˜4.3, it holds in the current section that there exists a semimartingale optimal strategy.

4.5 Cancellation of infinite variation

We now present an example where the infinite variation in the price impact process γ\gamma is “cancelled” by the infinite variation in the resilience process RR and we obtain the optimal strategy XX^{*} of finite variation.

Let x,dx,d\in\mathbb{R}, ξ=0\xi=0, λ0\lambda\equiv 0, ζ0\zeta\equiv 0, and μ0\mu\equiv 0. Suppose that r¯=1\overline{r}=-1 and ρ>0\rho>0 are deterministic constants, and that η\eta and σ\sigma are progressively measurable, dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. bounded processes such that η=σ\eta=\sigma dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. It then holds dP×ds|[0,T]dP\times ds|_{[0,T]}-a.e. that σ2+η2+2σηr¯=0\sigma^{2}+\eta^{2}+2\sigma\eta\overline{r}=0 and κ=ρ>0\kappa=\rho>0. In particular, the assumptions of Corollary˜3.4 are satisfied. The BSDE

dKs=ρKs2dsσsLs1ds+j=1mLsjdWsj,s[0,T],KT=12,\begin{split}dK_{s}&=\rho K_{s}^{2}ds-\sigma_{s}L_{s}^{1}ds+\sum_{j=1}^{m}L_{s}^{j}dW_{s}^{j},\quad s\in[0,T],\quad K_{T}=\frac{1}{2},\end{split}

which is BSDE (27) in the present setting, has the solution (K,L)=(K,0)(K,L)=(K,0) with Ks=12+(Ts)ρK_{s}=\frac{1}{2+(T-s)\rho}, s[0,T]s\in[0,T] (cf. Section˜4.1). It holds that θK\theta\equiv K, that (ψ,ϕ)=(0,0)(\psi,\phi)=(0,0) is the solution of (29), and that θ00\theta^{0}\equiv 0. It follows that (31) has the solution

H^s=(dγ0γ0x)exp(140sσr2𝑑rρ0sKr𝑑r+120sσr𝑑Wr1),s[0,T].\begin{split}\widehat{H}_{s}^{*}&=\left(\frac{d}{\sqrt{\gamma_{0}}}-\sqrt{\gamma_{0}}x\right)\exp\left(-\frac{1}{4}\int_{0}^{s}\sigma_{r}^{2}dr-\rho\int_{0}^{s}K_{r}dr+\frac{1}{2}\int_{0}^{s}\sigma_{r}dW_{r}^{1}\right),\quad s\in[0,T].\end{split}

For the optimal execution strategy from Corollary˜3.4 we then compute that

Xs=(xdγ0)1+(Ts)ρ2+Tρ,s[0,T).X_{s}^{*}=\left(x-\frac{d}{\gamma_{0}}\right)\frac{1+(T-s)\rho}{2+T\rho},\quad s\in[0,T).

The optimal strategy in the current setting with general stochastic σ=η\sigma=\eta and negative correlation r¯=1\overline{r}=-1 is thus the same as in the Obizhaeva-Wang setting σ=0=η\sigma=0=\eta (cf. [37, Proposition 3]; see also [1, Section 4.2]). In particular, the optimal strategy is deterministic and of finite variation, although the price impact γ\gamma and the resilience RR are both stochastic and of infinite variation (at least if σ=η\sigma=\eta is nonvanishing).

We finally remark that this setting does not reduce to the Obizhaeva-Wang setting σ=0=η\sigma=0=\eta. Indeed, while the optimal strategies for σ=0=η\sigma=0=\eta and for general stochastic σ=η\sigma=\eta with correlation r¯=1\overline{r}=-1 coincide, this is not true for the associated deviation processes. In general, it holds that

DsX=γ0(xdγ0)12+Tρexp(0sηr𝑑Wr1120sηr2𝑑r),s[0,T),D_{s}^{X^{*}}=-\gamma_{0}\left(x-\frac{d}{\gamma_{0}}\right)\frac{1}{2+T\rho}\exp\left(\int_{0}^{s}\eta_{r}dW_{r}^{1}-\frac{1}{2}\int_{0}^{s}\eta_{r}^{2}dr\right),\quad s\in[0,T),

which for a nonvanishing η\eta and xdγ0x\neq\frac{d}{\gamma_{0}} has infinite variation, whereas in the Obizhaeva-Wang setting is constant (take η=0\eta=0).

5 Proofs

In this section, we provide the proofs for the results presented in Section˜1 and Section˜2. We furthermore state and prove some auxiliary results that are used in the proofs of the main results.

For reference in several proofs, note that the order book height, i.e., the inverse of the price impact, has dynamics

dγs1=γs1((μsσs2)dsσsdWs1),s[0,T].d\gamma_{s}^{-1}=\gamma_{s}^{-1}\left(-(\mu_{s}-\sigma_{s}^{2})ds-\sigma_{s}dW^{1}_{s}\right),\quad s\in[0,T]. (42)

We moreover observe that by Itô’s lemma it holds that

dγs12=γs12(12μs18σs2)ds+12γs12σsdWs1,s[0,T],\begin{split}d\gamma_{s}^{\frac{1}{2}}&=\gamma_{s}^{\frac{1}{2}}\left(\frac{1}{2}\mu_{s}-\frac{1}{8}\sigma_{s}^{2}\right)ds+\frac{1}{2}\gamma_{s}^{\frac{1}{2}}\sigma_{s}dW^{1}_{s},\quad s\in[0,T],\end{split} (43)
dγs12=γs12(12μs+38σs2)ds12γs12σsdWs1,s[0,T].\begin{split}d\gamma_{s}^{-\frac{1}{2}}&=\gamma_{s}^{-\frac{1}{2}}\left(-\frac{1}{2}\mu_{s}+\frac{3}{8}\sigma_{s}^{2}\right)ds-\frac{1}{2}\gamma_{s}^{-\frac{1}{2}}\sigma_{s}dW^{1}_{s},\quad s\in[0,T].\end{split} (44)
Proof of ˜1.3.

Observe that integration by parts implies that for all s[t,T]s\in[t,T]

d(νsDs)=νsdDs+Dsdνs+d[ν,D]s=νsDsdRs+νsγsdXs+νsDsdRs+νsDsd[R]s+d[ν,D]s=νsγsdXs+νsDsd[R]s+d[ν,D]s.\begin{split}d(\nu_{s}D_{s})&=\nu_{s}dD_{s}+D_{s}d\nu_{s}+d[\nu,D]_{s}\\ &=-\nu_{s}D_{s}dR_{s}+\nu_{s}\gamma_{s}dX_{s}+\nu_{s}D_{s}dR_{s}+\nu_{s}D_{s}d[R]_{s}+d[\nu,D]_{s}\\ &=\nu_{s}\gamma_{s}dX_{s}+\nu_{s}D_{s}d[R]_{s}+d[\nu,D]_{s}.\end{split}

Since d[ν,D]s=νsd[R,D]s=νsDsd[R]sd[\nu,D]_{s}=\nu_{s}d[R,D]_{s}=-\nu_{s}D_{s}d[R]_{s}, s[t,T]s\in[t,T], it follows that the process D~s=νsDs\widetilde{D}_{s}=\nu_{s}D_{s}, s[t,T]s\in[t,T], D~t=d\widetilde{D}_{t-}=d, satisfies

dD~s=d(νsDs)=νsγsdXs,s[t,T].\begin{split}d\widetilde{D}_{s}=d(\nu_{s}D_{s})&=\nu_{s}\gamma_{s}dX_{s},\quad s\in[t,T].\end{split} (45)

In particular, D~\widetilde{D} is of finite variation. The facts that ΔDs=γsΔXs\Delta D_{s}=\gamma_{s}\Delta X_{s}, s[t,T]s\in[t,T], and dD~s=νsγsdXsd\widetilde{D}_{s}=\nu_{s}\gamma_{s}dX_{s}, s[t,T]s\in[t,T], imply that

[t,T](2Ds+ΔXsγs)𝑑Xs=[t,T](2Ds+ΔDs)𝑑Xs=[t,T](2Ds+ΔDs)γs1νs1𝑑D~s=[t,T]νs2γs1(2D~s+ΔD~s)𝑑D~s=[t,T]φsd(D~s2),\begin{split}\int_{[t,T]}\left(2D_{s-}+\Delta X_{s}\gamma_{s}\right)dX_{s}&=\int_{[t,T]}\left(2D_{s-}+\Delta D_{s}\right)dX_{s}=\int_{[t,T]}\left(2D_{s-}+\Delta D_{s}\right)\gamma_{s}^{-1}\nu_{s}^{-1}d\widetilde{D}_{s}\\ &=\int_{[t,T]}\nu_{s}^{-2}\gamma_{s}^{-1}\left(2\widetilde{D}_{s-}+\Delta\widetilde{D}_{s}\right)d\widetilde{D}_{s}=\int_{[t,T]}\varphi_{s}\,d(\widetilde{D}^{2}_{s}),\end{split} (46)

where we denote φs=νs2γs1\varphi_{s}=\nu_{s}^{-2}\gamma_{s}^{-1}, s[t,T]s\in[t,T], and, in the last equality, we use that d(D~s2)=(2D~s+ΔD~s)dD~sd(\widetilde{D}^{2}_{s})=(2\widetilde{D}_{s-}+\Delta\widetilde{D}_{s})\,d\widetilde{D}_{s}, as D~\widetilde{D} has finite variation. Summing up, (46) yields

[t,T]Ds𝑑Xs+12[t,T]ΔXsγs𝑑Xs=12(D~T2φTD~t2φttTD~s2𝑑φs)=12(γT1DT2γt1d2tTDs2νs2d(νs2γs1)).\begin{split}\int_{[t,T]}D_{s-}dX_{s}+\frac{1}{2}\int_{[t,T]}\Delta X_{s}\gamma_{s}dX_{s}&=\frac{1}{2}\left(\widetilde{D}_{T}^{2}\varphi_{T}-\widetilde{D}_{t-}^{2}\varphi_{t}-\int_{t}^{T}\widetilde{D}_{s}^{2}d\varphi_{s}\right)\\ &=\frac{1}{2}\left(\gamma_{T}^{-1}D_{T}^{2}-\gamma_{t}^{-1}d^{2}-\int_{t}^{T}D_{s}^{2}\nu_{s}^{2}d\left(\nu_{s}^{-2}\gamma_{s}^{-1}\right)\right).\end{split}

In order to show (11), we first obtain from (45) and integration by parts that

νrDrd=νrγrXrγtx[t,r]Xsd(νsγs)[t,r]d[νγ,X]s,r[t,T].\begin{split}\nu_{r}D_{r}-d&=\nu_{r}\gamma_{r}X_{r}-\gamma_{t}x-\int_{[t,r]}X_{s}d(\nu_{s}\gamma_{s})-\int_{[t,r]}d[\nu\gamma,X]_{s},\quad r\in[t,T].\end{split}

This implies that Dr=γrXr+νr1(dγtxtrXsd(νsγs))D_{r}=\gamma_{r}X_{r}+\nu_{r}^{-1}(d-\gamma_{t}x-\int_{t}^{r}X_{s}d(\nu_{s}\gamma_{s})), r[t,T]r\in[t,T]. ∎

Proof of ˜1.4.

We first consider the integrator ν2γ1\nu^{-2}\gamma^{-1} on the right hand side of (10). It holds by integration by parts and (9) that for all s[t,T]s\in[t,T]

d(νs2γs1)=νs1d(γs1νs1)+γs1νs1dνs1+d[ν1,γ1ν1]s=2νs1γs1dνs1+νs2dγs1+νs1d[γ1,ν1]s+d[ν1,γ1ν1]s=2νs2γs1dRs+νs2dγs1νs2d[γ1,R]s+d[ν1,γ1ν1]s.\begin{split}d(\nu_{s}^{-2}\gamma_{s}^{-1})&=\nu_{s}^{-1}d(\gamma_{s}^{-1}\nu_{s}^{-1})+\gamma_{s}^{-1}\nu_{s}^{-1}d\nu_{s}^{-1}+d[\nu^{-1},\gamma^{-1}\nu^{-1}]_{s}\\ &=2\nu_{s}^{-1}\gamma_{s}^{-1}d\nu_{s}^{-1}+\nu_{s}^{-2}d\gamma_{s}^{-1}+\nu_{s}^{-1}d[\gamma^{-1},\nu^{-1}]_{s}+d[\nu^{-1},\gamma^{-1}\nu^{-1}]_{s}\\ &=-2\nu_{s}^{-2}\gamma_{s}^{-1}dR_{s}+\nu_{s}^{-2}d\gamma_{s}^{-1}-\nu_{s}^{-2}d[\gamma^{-1},R]_{s}+d[\nu^{-1},\gamma^{-1}\nu^{-1}]_{s}.\end{split}

Note that for all s[t,T]s\in[t,T] we have

d[ν1,γ1ν1]s=νs1d[R,γ1ν1]s=νs1d[R,tγ1𝑑ν1+tν1𝑑γ1]s=νs1γs1d[R,ν1]sνs2d[R,γ1]s=νs2γs1d[R]sνs2d[R,γ1]s.\begin{split}d[\nu^{-1},\gamma^{-1}\nu^{-1}]_{s}&=-\nu_{s}^{-1}d[R,\gamma^{-1}\nu^{-1}]_{s}=-\nu_{s}^{-1}d\left[R,\int_{t}^{\cdot}\gamma^{-1}d\nu^{-1}+\int_{t}^{\cdot}\nu^{-1}d\gamma^{-1}\right]_{s}\\ &=-\nu_{s}^{-1}\gamma_{s}^{-1}d[R,\nu^{-1}]_{s}-\nu_{s}^{-2}d[R,\gamma^{-1}]_{s}=\nu_{s}^{-2}\gamma_{s}^{-1}d[R]_{s}-\nu_{s}^{-2}d[R,\gamma^{-1}]_{s}.\end{split}

It hence follows for all s[t,T]s\in[t,T] that

d(νs2γs1)=2νs2γs1dRs+νs2dγs12νs2d[γ1,R]s+νs2γs1d[R]s.\begin{split}d(\nu_{s}^{-2}\gamma_{s}^{-1})&=-2\nu_{s}^{-2}\gamma_{s}^{-1}dR_{s}+\nu_{s}^{-2}d\gamma_{s}^{-1}-2\nu_{s}^{-2}d[\gamma^{-1},R]_{s}+\nu_{s}^{-2}\gamma_{s}^{-1}d[R]_{s}.\end{split}

Plugged into (10) from ˜1.3, we obtain that

[t,T]Ds𝑑Xs+12[t,T]ΔXsγs𝑑Xs=12(γT1DT2γt1d2tTDs2(dγs1+γs1d[R]s2γs1dRs2d[γ1,R]s)).\begin{split}&\int_{[t,T]}D_{s-}dX_{s}+\frac{1}{2}\int_{[t,T]}\Delta X_{s}\gamma_{s}dX_{s}\\ &=\frac{1}{2}\left(\gamma_{T}^{-1}D_{T}^{2}-\gamma_{t}^{-1}d^{2}-\int_{t}^{T}D_{s}^{2}\left(d\gamma_{s}^{-1}+\gamma_{s}^{-1}d[R]_{s}-2\gamma_{s}^{-1}dR_{s}-2d[\gamma^{-1},R]_{s}\right)\right).\end{split} (47)

We further have by (3) and (42) that for all s[t,T]s\in[t,T]

dγs1+γs1d[R]s2γs1dRs2d[γ1,R]s=γs1(μsσs2)dsγs1σsdWs1+γs1ηs2ds2γs1ρsds2γs1ηsdWsR+2γs1σsηsr¯sds=γs1(2ρs+μsσs2ηs22σsηsr¯s)dsγs1σsdWs12γs1ηsdWsR.\begin{split}&d\gamma_{s}^{-1}+\gamma_{s}^{-1}d[R]_{s}-2\gamma_{s}^{-1}dR_{s}-2d[\gamma^{-1},R]_{s}\\ &=-\gamma_{s}^{-1}(\mu_{s}-\sigma_{s}^{2})ds-\gamma_{s}^{-1}\sigma_{s}dW^{1}_{s}+\gamma_{s}^{-1}\eta_{s}^{2}ds-2\gamma_{s}^{-1}\rho_{s}ds-2\gamma_{s}^{-1}\eta_{s}dW_{s}^{R}+2\gamma_{s}^{-1}\sigma_{s}\eta_{s}\overline{r}_{s}ds\\ &=-\gamma_{s}^{-1}\left(2\rho_{s}+\mu_{s}-\sigma_{s}^{2}-\eta_{s}^{2}-2\sigma_{s}\eta_{s}\overline{r}_{s}\right)ds-\gamma_{s}^{-1}\sigma_{s}dW^{1}_{s}-2\gamma_{s}^{-1}\eta_{s}dW_{s}^{R}.\end{split} (48)

It follows from assumption (A1) and the boundedness of the input processes that

E[|tTDs2γs1(2ρs+μsσs2ηs22σsηsr¯s)𝑑s|]<.\begin{split}E\left[\left\lvert\int_{t}^{T}D_{s}^{2}\gamma_{s}^{-1}\left(2\rho_{s}+\mu_{s}-\sigma_{s}^{2}-\eta_{s}^{2}-2\sigma_{s}\eta_{s}\overline{r}_{s}\right)ds\right\rvert\right]<\infty.\end{split}

The Burkholder-Davis-Gundy inequality together with assumption (A3) shows that it holds for some constant c(0,)c\in(0,\infty) that

E[supr[t,T]|trDs2γs1σs𝑑Ws1|]cE[(tTDs4γs2σs2𝑑s)12]<.\begin{split}E\left[\sup_{r\in[t,T]}\left\lvert\int_{t}^{r}D_{s}^{2}\gamma_{s}^{-1}\sigma_{s}dW^{1}_{s}\right\rvert\right]&\leq cE\left[\left(\int_{t}^{T}D_{s}^{4}\gamma_{s}^{-2}\sigma_{s}^{2}ds\right)^{\frac{1}{2}}\right]<\infty.\end{split}

We therefore have that Et[tTDs2γs1σs𝑑Ws1]=0.E_{t}[\int_{t}^{T}D_{s}^{2}\gamma_{s}^{-1}\sigma_{s}dW^{1}_{s}]=0. Similarly, assumption (A2) implies that Et[tT2Ds2γs1ηs𝑑WsR]=0.E_{t}[\int_{t}^{T}2D_{s}^{2}\gamma_{s}^{-1}\eta_{s}dW_{s}^{R}]=0. It thus follows from (47), (48), and (12) that

Et[[t,T]Ds𝑑Xs+12[t,T]ΔXsγs𝑑Xs]=12Et[γT1DT2+tTDs2γs12κs𝑑s]d22γt.E_{t}\left[\int_{[t,T]}D_{s-}dX_{s}+\frac{1}{2}\int_{[t,T]}\Delta X_{s}\gamma_{s}dX_{s}\right]=\frac{1}{2}E_{t}\left[\gamma_{T}^{-1}D_{T}^{2}+\int_{t}^{T}D_{s}^{2}\gamma_{s}^{-1}2\kappa_{s}ds\right]-\frac{d^{2}}{2\gamma_{t}}.

By definition (7) of JfvJ^{fv} this proves (13). ∎

The dynamics that we compute in the following lemma are used in the proofs of Lemma˜1.6 and Lemma˜5.5.

Lemma 5.1.

Let t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}. Assume that X=(Xs)s[t,T]X=(X_{s})_{s\in[t-,T]} is a progressively measurable process such that tTXs2𝑑s<\int_{t}^{T}X_{s}^{2}ds<\infty a.s. For αs=γs12νs1\alpha_{s}=\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}, s[t,T]s\in[t,T], and βs=dγtxtsXrd(νrγr)\beta_{s}=d-\gamma_{t}x-\int_{t}^{s}X_{r}d(\nu_{r}\gamma_{r}), s[t,T]s\in[t,T], it then holds for all s[t,T]s\in[t,T] that

d(αsβs)=γs12Xs((μs+ρs+ηs2+σsηsr¯s)ds+(σs+ηsr¯s)dWs1+ηs1r¯s2dWs2)+αsβs((ρs12μs+38σs2+12σsηsr¯s)ds+(ηsr¯s12σs)dWs1ηs1r¯s2dWs2)+γs12Xs(32ηsσsr¯s+12σs2+ηs2)ds.\begin{split}&d(\alpha_{s}\beta_{s})\\ &=-\gamma_{s}^{\frac{1}{2}}X_{s}\Bigg{(}\big{(}\mu_{s}+\rho_{s}+\eta_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}\big{)}ds+\big{(}\sigma_{s}+\eta_{s}\overline{r}_{s}\big{)}dW_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}\\ &\quad+\alpha_{s}\beta_{s}\Bigg{(}\Big{(}-\rho_{s}-\frac{1}{2}\mu_{s}+\frac{3}{8}\sigma_{s}^{2}+\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}\Big{)}ds+\Big{(}-\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}\Big{)}dW_{s}^{1}-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}\\ &\quad+\gamma_{s}^{\frac{1}{2}}X_{s}\left(\frac{3}{2}\eta_{s}\sigma_{s}\overline{r}_{s}+\frac{1}{2}\sigma_{s}^{2}+\eta_{s}^{2}\right)ds.\end{split} (49)
Proof.

Integration by parts implies that

d(αsβs)=αsXsd(νsγs)+βsd(γs12νs1)Xsd[γ12ν1,νγ]s,s[t,T].d(\alpha_{s}\beta_{s})=-\alpha_{s}X_{s}d(\nu_{s}\gamma_{s})+\beta_{s}d(\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1})-X_{s}d[\gamma^{-\frac{1}{2}}\nu^{-1},\nu\gamma]_{s},\quad s\in[t,T]. (50)

Furthermore, it holds by integration by parts, (8), (3) and (4) that for all s[t,T]s\in[t,T]

d(νsγs)=νsdγs+γsνsdRs+γsνsd[R]s+νsd[R,γ]s=νsγsμsds+νsγsσsdWs1+νsγsρsds+νsγsηsr¯sdWs1+νsγsηs1r¯s2dWs2+νsγsηs2ds+νsγsσsηsr¯sds=νsγs((μs+ρs+ηs2+σsηsr¯s)ds+(σs+ηsr¯s)dWs1+ηs1r¯s2dWs2).\begin{split}d(\nu_{s}\gamma_{s})&=\nu_{s}d\gamma_{s}+\gamma_{s}\nu_{s}dR_{s}+\gamma_{s}\nu_{s}d[R]_{s}+\nu_{s}d[R,\gamma]_{s}\\ &=\nu_{s}\gamma_{s}\mu_{s}ds+\nu_{s}\gamma_{s}\sigma_{s}dW^{1}_{s}+\nu_{s}\gamma_{s}\rho_{s}ds+\nu_{s}\gamma_{s}\eta_{s}\overline{r}_{s}dW_{s}^{1}+\nu_{s}\gamma_{s}\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\\ &\quad+\nu_{s}\gamma_{s}\eta_{s}^{2}ds+\nu_{s}\gamma_{s}\sigma_{s}\eta_{s}\overline{r}_{s}ds\\ &=\nu_{s}\gamma_{s}\Bigg{(}\big{(}\mu_{s}+\rho_{s}+\eta_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}\big{)}ds+\big{(}\sigma_{s}+\eta_{s}\overline{r}_{s}\big{)}dW_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}.\end{split} (51)

Also by integration by parts, and using (9), (3) and (44), we obtain for all s[t,T]s\in[t,T] that

d(γs12νs1)=γs12νs1dRs+νs1dγs12νs1d[R,γ12]s=γs12νs1ρsdsγs12νs1ηsr¯sdWs1γs12νs1ηs1r¯s2dWs2+γs12νs1(12μs+38σs2)ds12γs12νs1σsdWs1+12γs12νs1σsηsr¯sds=αs((ρs12μs+38σs2+12σsηsr¯s)ds+(ηsr¯s12σs)dWs1ηs1r¯s2dWs2).\begin{split}&d(\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1})=-\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}dR_{s}+\nu_{s}^{-1}d\gamma_{s}^{-\frac{1}{2}}-\nu_{s}^{-1}d[R,\gamma^{-\frac{1}{2}}]_{s}\\ &=-\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}\rho_{s}ds-\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}\eta_{s}\overline{r}_{s}dW_{s}^{1}-\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\\ &\quad+\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}\left(-\frac{1}{2}\mu_{s}+\frac{3}{8}\sigma_{s}^{2}\right)ds-\frac{1}{2}\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}\sigma_{s}dW_{s}^{1}+\frac{1}{2}\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}\sigma_{s}\eta_{s}\overline{r}_{s}ds\\ &=\alpha_{s}\Bigg{(}\Big{(}-\rho_{s}-\frac{1}{2}\mu_{s}+\frac{3}{8}\sigma_{s}^{2}+\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}\Big{)}ds+\Big{(}-\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}\Big{)}dW_{s}^{1}-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}.\end{split} (52)

It follows from (51) and (52) for all s[t,T]s\in[t,T] that

d[γ12ν1,νγ]s=γs12(ηsr¯s12σs)(σs+ηsr¯s)dsγs12ηs2(1r¯s2)ds=γs12(32ηsσsr¯s+12σs2+ηs2)ds.\begin{split}d[\gamma^{-\frac{1}{2}}\nu^{-1},\nu\gamma]_{s}&=\gamma_{s}^{\frac{1}{2}}\Big{(}-\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}\Big{)}\big{(}\sigma_{s}+\eta_{s}\overline{r}_{s}\big{)}ds-\gamma_{s}^{\frac{1}{2}}\eta_{s}^{2}(1-\overline{r}_{s}^{2})ds\\ &=-\gamma_{s}^{\frac{1}{2}}\left(\frac{3}{2}\eta_{s}\sigma_{s}\overline{r}_{s}+\frac{1}{2}\sigma_{s}^{2}+\eta_{s}^{2}\right)ds.\end{split} (53)

We then plug (51), (52) and (53) into (50), which yields (49). ∎

Proof of Lemma˜1.6.

We denote αs=γs12νs1\alpha_{s}=\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}, s[t,T]s\in[t,T], and βs=dγtxtsXrd(νrγr)\beta_{s}=d-\gamma_{t}x-\int_{t}^{s}X_{r}d(\nu_{r}\gamma_{r}), s[t,T]s\in[t,T]. It then holds that H¯s=αsβs\overline{H}_{s}=\alpha_{s}\beta_{s}, s[t,T]s\in[t,T]. We use Lemma˜5.1 and substitute γ12X=H¯γ12D-\gamma^{\frac{1}{2}}X=\overline{H}-\gamma^{-\frac{1}{2}}D in (49) to obtain for all s[t,T]s\in[t,T] that

dH¯s=(H¯sγs12Ds)((μs+ρs12σsηsr¯s12σs2)ds+(σs+ηsr¯s)dWs1+ηs1r¯s2dWs2)+H¯s((ρs12μs+38σs2+12σsηsr¯s)ds+(ηsr¯s12σs)dWs1ηs1r¯s2dWs2)=γs12Ds((μs+ρs12σsηsr¯s12σs2)ds+(σs+ηsr¯s)dWs1+ηs1r¯s2dWs2)+H¯s((12μs18σs2)ds+12σsdWs1).\begin{split}&d\overline{H}_{s}\\ &=\big{(}\overline{H}_{s}-\gamma_{s}^{-\frac{1}{2}}D_{s}\big{)}\Bigg{(}\left(\mu_{s}+\rho_{s}-\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}^{2}\right)ds+\big{(}\sigma_{s}+\eta_{s}\overline{r}_{s}\big{)}dW_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}\\ &\quad+\overline{H}_{s}\Bigg{(}\Big{(}-\rho_{s}-\frac{1}{2}\mu_{s}+\frac{3}{8}\sigma_{s}^{2}+\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}\Big{)}ds+\Big{(}-\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}\Big{)}dW_{s}^{1}-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}\\ &=-\gamma_{s}^{-\frac{1}{2}}D_{s}\Bigg{(}\left(\mu_{s}+\rho_{s}-\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}^{2}\right)ds+\big{(}\sigma_{s}+\eta_{s}\overline{r}_{s}\big{)}dW_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}\\ &\quad+\overline{H}_{s}\Bigg{(}\left(\frac{1}{2}\mu_{s}-\frac{1}{8}\sigma_{s}^{2}\right)ds+\frac{1}{2}\sigma_{s}dW_{s}^{1}\Bigg{)}.\end{split}

This proves the dynamics in (18).

In particular, H¯\overline{H} satisfies an SDE that is linear in H¯\overline{H} and γ12D\gamma^{-\frac{1}{2}}D. Furthermore, boundedness of ρ,μ,σ,η,r¯\rho,\mu,\sigma,\eta,\overline{r} implies that the coefficients of the SDE are bounded. Since moreover E[tT(γs12Ds)2𝑑s]<E[\int_{t}^{T}\big{(}\gamma_{s}^{-\frac{1}{2}}D_{s}\big{)}^{2}ds]<\infty by assumption (A1) and H¯t=γt12dγt12x\overline{H}_{t}=\gamma_{t}^{-\frac{1}{2}}d-\gamma_{t}^{\frac{1}{2}}x (cf. (17)) is square integrable, we have that E[sups[t,T]H¯s2]<E[\sup_{s\in[t,T]}\overline{H}_{s}^{2}]<\infty (see, e.g., [41, Theorem 3.2.2 and Theorem 3.3.1]).

We next prove that cost functional (16) admits representation (19). To this end, note that by (17) it holds for all s[t,T]s\in[t,T] that

γs(Xsζs)2=(γs12DsH¯sγs12ζs)2=γs1Ds22γs12Ds(H¯s+γs12ζs)+(H¯s+γs12ζs)2.\gamma_{s}\!\left(X_{s}\!-\zeta_{s}\right)^{2}=\left(\gamma^{-\frac{1}{2}}_{s}D_{s}-\overline{H}_{s}-\gamma_{s}^{\frac{1}{2}}\zeta_{s}\right)^{\!2}\!=\gamma_{s}^{-1}D_{s}^{2}-2\gamma_{s}^{-\frac{1}{2}}D_{s}\left(\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\right)+\left(\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\right)^{\!2}\!.

Due to assumption (5) on ζ\zeta and E[sups[t,T]H¯s2]<E[\sup_{s\in[t,T]}\overline{H}_{s}^{2}]<\infty, we have that Et[tT(H¯s+γs12ζs)2𝑑s]<E_{t}[\int_{t}^{T}(\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s})^{2}ds]<\infty. This, assumption (A1), and the Cauchy–Schwarz inequality imply that also Et[tT|γs12Ds(H¯s+γs12ζs)|𝑑s]<E_{t}[\int_{t}^{T}\lvert\gamma_{s}^{-\frac{1}{2}}D_{s}(\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s})\rvert ds]<\infty. Since λ\lambda is bounded, we conclude that

Et[tTλsγs(Xsζs)2𝑑s]=Et[tTλsγs1Ds2𝑑s]+Et[tTλs(H¯s+γs12ζs)2𝑑s]2Et[tTλsγs12Ds(H¯s+γs12ζs)𝑑s],\begin{split}E_{t}\left[\int_{t}^{T}\lambda_{s}\gamma_{s}\left(X_{s}-\zeta_{s}\right)^{2}ds\right]&=E_{t}\left[\int_{t}^{T}\lambda_{s}\gamma_{s}^{-1}D_{s}^{2}ds\right]+E_{t}\left[\int_{t}^{T}\lambda_{s}\left(\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\right)^{2}ds\right]\\ &\quad-2E_{t}\left[\int_{t}^{T}\lambda_{s}\gamma_{s}^{-\frac{1}{2}}D_{s}\left(\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\right)ds\right],\end{split} (54)

where all conditional expectations are well-defined and finite. Moreover, (17) implies that γT12DT=H¯T+γT12XT\gamma_{T}^{-\frac{1}{2}}D_{T}=\overline{H}_{T}+\gamma_{T}^{\frac{1}{2}}X_{T}, and thus γT1DT2=(H¯T+γTξ)2\gamma_{T}^{-1}D_{T}^{2}=(\overline{H}_{T}+\sqrt{\gamma_{T}}\xi)^{2}. Inserting this and (54) into (16), we obtain (19). ∎

Lemma 5.2.

Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}. Then, (20) defines a metric on 𝒜tpm(x,d)\mathcal{A}_{t}^{pm}(x,d) (identifying any processes that are equal dP×ds|[t,T]dP\times ds|_{[t,T]}-a.e.).

Proof.

Note first that it holds for all X,Y𝒜tpm(x,d)X,Y\in\mathcal{A}_{t}^{pm}(x,d) that 𝐝(X,Y)0\mathbf{d}(X,Y)\geq 0, and that 𝐝(X,Y)\mathbf{d}(X,Y) is finite due to (A1). Symmetry of 𝐝\mathbf{d} is obvious. The triangle inequality follows from the Cauchy–Schwarz inequality.

Let X,Y𝒜tpm(x,d)X,Y\in\mathcal{A}_{t}^{pm}(x,d) with associated deviation processes DX,DYD^{X},D^{Y}.

If X=YX=Y dP×ds|[t,T]dP\times ds|_{[t,T]}-a.e., then γ12DX=γ12DY\gamma^{-\frac{1}{2}}D^{X}=\gamma^{-\frac{1}{2}}D^{Y} dP×ds|[t,T]dP\times ds|_{[t,T]}-a.e., and thus 𝐝(X,Y)=(E[tT(γs12DsXγs12DsY)2𝑑s])12=0\mathbf{d}(X,Y)=(E[\int_{t}^{T}(\gamma_{s}^{-\frac{1}{2}}D_{s}^{X}-\gamma_{s}^{-\frac{1}{2}}D_{s}^{Y})^{2}ds])^{\frac{1}{2}}=0.

For the other direction, suppose that 𝐝(X,Y)=0\mathbf{d}(X,Y)=0. This implies that γ12DXγ12DY=0\gamma^{-\frac{1}{2}}D^{X}-\gamma^{-\frac{1}{2}}D^{Y}=0 dP×ds|[t,T]dP\times ds|_{[t,T]}-a.e. By definition of DXD^{X} and DYD^{Y} it further follows from a multiplication by νγ12\nu\gamma^{\frac{1}{2}} that νsγs(XsYs)=ts(XrYr)d(νrγr)\nu_{s}\gamma_{s}(X_{s}-Y_{s})=\int_{t}^{s}(X_{r}-Y_{r})d(\nu_{r}\gamma_{r}) dP×ds|[t,T]-a.e.\,dP\times ds|_{[t,T]}\text{-a.e.} Observe that νγ>0\nu\gamma>0 and consider the stochastic integral equation

Ks=tsKrνr1γr1d(νrγr),s[t,T].K_{s}=\int_{t}^{s}K_{r}\nu_{r}^{-1}\gamma_{r}^{-1}d(\nu_{r}\gamma_{r}),\quad s\in[t,T]. (55)

Define L=(Ls)s[0,T]L=(L_{s})_{s\in[0,T]} by L0=0L_{0}=0,

dLs=(μs+ρs+ηs2+σsηsr¯s)ds+(σs+ηsr¯s)dWs1+ηs1r¯s2dWs2,s[0,T].dL_{s}=\big{(}\mu_{s}+\rho_{s}+\eta_{s}^{2}+\sigma_{s}\eta_{s}\overline{r}_{s}\big{)}ds+\big{(}\sigma_{s}+\eta_{s}\overline{r}_{s}\big{)}dW_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2},\quad s\in[0,T].

It then follows from (51) that (55) can be written as Ks=tsKr𝑑LrK_{s}=\int_{t}^{s}K_{r}dL_{r}, s[t,T]s\in[t,T]. This has the unique solution K=0K=0. We therefore conclude that X=YX=Y dP×ds|[t,T]dP\times ds|_{[t,T]}-a.e. ∎

We now prepare the proof of ˜1.7. The next result on the scaled hidden deviation is helpful in ˜1.7 in order to show convergence of the cost functional.

Lemma 5.3.

Let t[0,T]t\in[0,T], x,dx,d\in\mathbb{R}, and X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d) with associated deviation DD and scaled hidden deviation H¯\overline{H}. Suppose in addition that (Xn)n(X^{n})_{n\in\mathbb{N}} is a sequence in 𝒜tpm(x,d)\mathcal{A}^{pm}_{t}(x,d) such that limnE[tT(DsnDs)2γs1𝑑s]=0\lim_{n\to\infty}E[\int_{t}^{T}(D_{s}^{n}-D_{s})^{2}\gamma_{s}^{-1}ds]=0. for the associated deviation processes DnD^{n}, nn\in\mathbb{N}. It then holds for the associated scaled hidden deviation processes H¯n\overline{H}^{n}, nn\in\mathbb{N}, that limnE[sups[t,T](H¯snH¯s)2]=0.\lim_{n\to\infty}E[\sup_{s\in[t,T]}(\overline{H}_{s}^{n}-\overline{H}_{s})^{2}]=0.

Proof.

Define δH¯n=H¯nH¯\delta\overline{H}^{n}=\overline{H}^{n}-\overline{H}, nn\in\mathbb{N}, and let for nn\in\mathbb{N}, s[t,T]s\in[t,T], zz\in\mathbb{R}

bsn(z)=12(2(ρs+μs)σs2σsηsr¯s)(γs12Dsnγs12Ds)+12(μs14σs2)z,asn(z)=((σs+ηsr¯s)(γs12Dsnγs12Ds)+12σsz,ηs1r¯s2(γs12Dsnγs12Ds)).\begin{split}b^{n}_{s}(z)&=-\frac{1}{2}\Big{(}2(\rho_{s}+\mu_{s})-\sigma_{s}^{2}-\sigma_{s}\eta_{s}\overline{r}_{s}\Big{)}\big{(}\gamma^{-\frac{1}{2}}_{s}D^{n}_{s}-\gamma^{-\frac{1}{2}}_{s}D_{s}\big{)}+\frac{1}{2}\left(\mu_{s}-\frac{1}{4}\sigma_{s}^{2}\right)z,\\ a_{s}^{n}(z)&=\bigg{(}-(\sigma_{s}+\eta_{s}\overline{r}_{s})\big{(}\gamma^{-\frac{1}{2}}_{s}D^{n}_{s}-\gamma^{-\frac{1}{2}}_{s}D_{s}\big{)}+\frac{1}{2}\sigma_{s}z,-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\big{(}\gamma^{-\frac{1}{2}}_{s}D^{n}_{s}-\gamma^{-\frac{1}{2}}_{s}D_{s}\big{)}\bigg{)}.\end{split}

In view of (18) it then holds for all nn\in\mathbb{N} that

d(δH¯sn)=bsn(δH¯sn)ds+asn(δH¯sn)d(Ws1Ws2),s[t,T],δH¯tn=0.d(\delta\overline{H}^{n}_{s})=b_{s}^{n}(\delta\overline{H}_{s}^{n})ds+a_{s}^{n}(\delta\overline{H}_{s}^{n})d\begin{pmatrix}W_{s}^{1}\\ W_{s}^{2}\end{pmatrix},\quad s\in[t,T],\quad\delta\overline{H}_{t}^{n}=0.

Linearity of bnb^{n}, ana^{n}, nn\in\mathbb{N}, and boundedness of μ,ρ,σ,η,r¯\mu,\rho,\sigma,\eta,\overline{r} imply that there exists c1(0,)c_{1}\in(0,\infty) such that for all nn\in\mathbb{N} and all z1,z2z_{1},z_{2}\in\mathbb{R} it holds dP×ds|[t,T]dP\times ds|_{[t,T]}-a.e. that

|bn(z1)bn(z2)|+an(z1)an(z2)212|μ14σ2||z1z2|+12|σ||z1z2|c1|z1z2|.\lvert b^{n}(z_{1})-b^{n}(z_{2})\rvert+\lVert a^{n}(z_{1})-a^{n}(z_{2})\rVert_{2}\leq\frac{1}{2}\left\lvert\mu-\frac{1}{4}\sigma^{2}\right\rvert\lvert z_{1}-z_{2}\rvert+\frac{1}{2}\lvert\sigma\rvert\lvert z_{1}-z_{2}\rvert\leq c_{1}\lvert z_{1}-z_{2}\rvert.

By boundedness of μ,ρ,σ,η,r¯\mu,\rho,\sigma,\eta,\overline{r} and Jensen’s inequality, we have some c2(0,)c_{2}\in(0,\infty) such that for all nn\in\mathbb{N},

E[(tT|bsn(0)|𝑑s)2]+E[tTasn(0)22𝑑s]c2E[tT(DsnDs)2γs1𝑑s].E\left[\left(\int_{t}^{T}\lvert b_{s}^{n}(0)\rvert ds\right)^{2}\right]+E\left[\int_{t}^{T}\lVert a^{n}_{s}(0)\rVert_{2}^{2}ds\right]\leq c_{2}E\left[\int_{t}^{T}(D_{s}^{n}-D_{s})^{2}\gamma_{s}^{-1}ds\right].

E.g., [41, Theorem 3.2.2] (see also [41, Theorem 3.4.2]) now implies that there exists c3(0,)c_{3}\in(0,\infty) such that for all nn\in\mathbb{N}

E[sups[t,T]|H¯snH¯s|2]c3E[(tT|bsn(0)|𝑑s)2+tTasn(0)22𝑑s]c2c3E[tT(DsnDs)2γs1𝑑s].\begin{split}E\left[\sup_{s\in[t,T]}\lvert\overline{H}_{s}^{n}-\overline{H}_{s}\rvert^{2}\right]&\leq c_{3}E\left[\left(\int_{t}^{T}\lvert b_{s}^{n}(0)\rvert ds\right)^{2}+\int_{t}^{T}\lVert a^{n}_{s}(0)\rVert_{2}^{2}ds\right]\\ &\leq c_{2}c_{3}E\left[\int_{t}^{T}(D_{s}^{n}-D_{s})^{2}\gamma_{s}^{-1}ds\right].\end{split}

The claim follows from the assumption that limnE[tT(DsnDs)2γs1𝑑s]=0\lim_{n\to\infty}E[\int_{t}^{T}\left(D_{s}^{n}-D_{s}\right)^{2}\gamma_{s}^{-1}ds]=0. ∎

In order to establish existence of an appropriate approximating sequence in ˜1.7, we rely on Lemma˜5.4 below. For its statement and the proof of the second part of ˜1.7, we introduce a process Z=(Zs)s[0,T]Z=(Z_{s})_{s\in[0,T]} defined by

Zs=exp(0s(12σr+ηrr¯r)𝑑Wr10sηr1r¯r2𝑑Wr2),s[0,T].Z_{s}=\exp\left(-\int_{0}^{s}\left(\frac{1}{2}\sigma_{r}+\eta_{r}\overline{r}_{r}\right)dW^{1}_{r}-\int_{0}^{s}\eta_{r}\sqrt{1-\overline{r}_{r}^{2}}dW_{r}^{2}\right),\quad s\in[0,T]. (56)

Observe that by Itô’s lemma, ZZ solves the SDE

dZs=Zs2((12σs+ηsr¯s)2+ηs2(1r¯s2))dsZs(12σs+ηsr¯s,ηs1r¯s2)d(Ws1Ws2),s[0,T],Z0=1.\begin{split}&dZ_{s}=\frac{Z_{s}}{2}\left(\left(\frac{1}{2}\sigma_{s}+\eta_{s}\overline{r}_{s}\right)^{2}+\eta_{s}^{2}(1-\overline{r}_{s}^{2})\right)ds-Z_{s}\left(\frac{1}{2}\sigma_{s}+\eta_{s}\overline{r}_{s},\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}\right)d\begin{pmatrix}W^{1}_{s}\\ W_{s}^{2}\end{pmatrix},\\ &s\in[0,T],\quad Z_{0}=1.\end{split} (57)
Lemma 5.4.

Let t[0,T]t\in[0,T] and let u=(us)s[t,T]t2u=(u_{s})_{s\in[t,T]}\in\mathcal{L}_{t}^{2}. Then there exists a sequence of bounded càdlàg finite variation processes (vn)n(v^{n})_{n\in\mathbb{N}} such that

limnE[tT(usZsvsn)2Zs2𝑑s]=0.\lim_{n\to\infty}E\left[\int_{t}^{T}\left(\frac{u_{s}}{Z_{s}}-v_{s}^{n}\right)^{2}Z_{s}^{2}ds\right]=0.

In particular, for the sequence of processes (un)n(u^{n})_{n\in\mathbb{N}} defined by un=vnZu^{n}=v^{n}Z, nn\in\mathbb{N}, it holds for all nn\in\mathbb{N} that unu^{n} is a càdlàg semimartingale and E[sups[t,T]|usn|p]<E[\sup_{s\in[t,T]}\lvert u_{s}^{n}\rvert^{p}]<\infty for any p2p\geq 2 (in particular, unt2u^{n}\in\mathcal{L}_{t}^{2}), and that limnE[tT(ususn)2𝑑s]=0.\lim_{n\to\infty}E[\int_{t}^{T}\left(u_{s}-u_{s}^{n}\right)^{2}ds]=0.

Proof.

Define A=(As)s[0,T]A=(A_{s})_{s\in[0,T]} by As=0sZr2𝑑rA_{s}=\int_{0}^{s}Z_{r}^{2}dr, s[0,T]s\in[0,T]. Moreover, let v=(vs)s[t,T]v=(v_{s})_{s\in[t,T]} be defined by vs=usZsv_{s}=\frac{u_{s}}{Z_{s}}, s[t,T]s\in[t,T]. We verify the assumptions of Lemma 2.7 in Section 3.2 of [33]. The process AA is continuous, adapted and nondecreasing. Note that boundedness of σ\sigma, η\eta and r¯\overline{r} implies that the coefficients of (57) are bounded. It follows for any p2p\geq 2 that E[sups[0,T]|Zs|p]<E[\sup_{s\in[0,T]}\lvert Z_{s}\rvert^{p}]<\infty (see, e.g., [41, Theorem 3.4.3]), and hence E[AT]=E[0TZr2𝑑r]<E[A_{T}]=E[\int_{0}^{T}Z_{r}^{2}dr]<\infty. Since ut2u\in\mathcal{L}_{t}^{2}, we have that vv is progressively measurable and satisfies E[tTvs2𝑑As]=E[tTus2𝑑s]<E[\int_{t}^{T}v_{s}^{2}dA_{s}]=E[\int_{t}^{T}u_{s}^{2}ds]<\infty. Thus, Lemma 2.7 in Section 3.2 of [33] applies and yields that there exists a sequence (v^n)n(\hat{v}^{n})_{n\in\mathbb{N}} of (càglàd) simple processes v^n=(v^sn)s[t,T]\hat{v}^{n}=(\hat{v}_{s}^{n})_{s\in[t,T]}, nn\in\mathbb{N}, such that limnE[tT(vsv^sn)2𝑑As]=0\lim_{n\to\infty}E[\int_{t}^{T}(v_{s}-\hat{v}_{s}^{n})^{2}dA_{s}]=0. Define vsn(ω)=limrsv^rn(ω)v_{s}^{n}(\omega)=\lim_{r\downarrow s}\hat{v}_{r}^{n}(\omega), s[t,T)s\in[t,T), ωΩ\omega\in\Omega, nn\in\mathbb{N}, and vTn=0v_{T}^{n}=0, nn\in\mathbb{N}. Then, (vn)n(v^{n})_{n\in\mathbb{N}} is a sequence of bounded càdlàg finite variation processes such that limnE[tT(vsvsn)2𝑑As]=0\lim_{n\to\infty}E[\int_{t}^{T}(v_{s}-v_{s}^{n})^{2}dA_{s}]=0. Note that for each nn\in\mathbb{N}, un=(usn)s[t,T]u^{n}=(u_{s}^{n})_{s\in[t,T]} defined by usn=vsnZsu^{n}_{s}=v^{n}_{s}Z_{s}, s[t,T]s\in[t,T], is càdlàg. Since vnv^{n} is bounded for all nn\in\mathbb{N} and E[sups[0,T]|Zs|p]<E[\sup_{s\in[0,T]}\lvert Z_{s}\rvert^{p}]<\infty for any p2p\geq 2, we have that E[sups[t,T]|usn|p]E[\sup_{s\in[t,T]}\lvert u_{s}^{n}\rvert^{p}] is finite for all nn\in\mathbb{N} and any p2p\geq 2. It furthermore holds that E[tT(ususn)2𝑑s]=E[tT(vsvsn)2𝑑As]0E[\int_{t}^{T}(u_{s}-u_{s}^{n})^{2}ds]=E[\int_{t}^{T}(v_{s}-v_{s}^{n})^{2}dA_{s}]\to 0 as nn\to\infty. ∎

For the part in ˜1.7 on completeness of (𝒜tpm(x,d),𝐝)(\mathcal{A}_{t}^{pm}(x,d),\mathbf{d}) we show how to construct an execution strategy X0𝒜tpm(x,d)X^{0}\in\mathcal{A}_{t}^{pm}(x,d) based on a square integrable process u0u^{0} and a process H0H^{0} that satisfies SDE (18) (with u0u^{0} instead of γ12D\gamma^{-\frac{1}{2}}D). This result is also crucial for Lemma˜2.2.

Lemma 5.5.

Let t[0,T]t\in[0,T] and x,dx,d\in\mathbb{R}. Suppose that u0=(us0)s[t,T]t2u^{0}=(u^{0}_{s})_{s\in[t,T]}\in\mathcal{L}_{t}^{2}, and let H0=(Hs0)s[t,T]H^{0}=(H^{0}_{s})_{s\in[t,T]} be given by Ht0=dγtγtxH^{0}_{t}=\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,

dHs0=(12(μs14σs2)Hs012(2(ρs+μs)σs2σsηsr¯s)us0)ds+(12σsHs0(σs+ηsr¯s)us0)dWs1ηs1r¯s2us0dWs2,s[t,T].\begin{split}dH^{0}_{s}&=\left(\frac{1}{2}\left(\mu_{s}-\frac{1}{4}\sigma_{s}^{2}\right)H^{0}_{s}-\frac{1}{2}\left(2(\rho_{s}+\mu_{s})-\sigma_{s}^{2}-\sigma_{s}\eta_{s}\overline{r}_{s}\right)u^{0}_{s}\right)ds\\ &\quad+\left(\frac{1}{2}\sigma_{s}H^{0}_{s}-(\sigma_{s}+\eta_{s}\overline{r}_{s})u^{0}_{s}\right)dW^{1}_{s}-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}u^{0}_{s}dW_{s}^{2},\quad s\in[t,T].\end{split} (58)

Define X0=(Xs0)s[t,T]X^{0}=(X^{0}_{s})_{s\in[t-,T]} by Xs0=γs12(us0Hs0)X^{0}_{s}=\gamma_{s}^{-\frac{1}{2}}(u^{0}_{s}-H^{0}_{s}), s[t,T)s\in[t,T), Xt0=xX^{0}_{t-}=x, XT0=ξX^{0}_{T}=\xi. Then, X0𝒜tpm(x,d)X^{0}\in\mathcal{A}_{t}^{pm}(x,d), and for the associated deviation process D0=(Ds0)s[t,T]D^{0}=(D^{0}_{s})_{s\in[t-,T]} it holds D0=γX0+γ12H0D^{0}=\gamma X^{0}+\gamma^{\frac{1}{2}}H^{0}.

Proof.

First, X0X^{0} is progressively measurable and has initial value Xt0=xX^{0}_{t-}=x and terminal value XT0=ξX^{0}_{T}=\xi. Furthermore, it holds that

tT(Xs0)2𝑑s2tTγs1(us0)2𝑑s+2tTγs1(Hs0)2𝑑s< a.s.\begin{split}\int_{t}^{T}(X_{s}^{0})^{2}ds&\leq 2\int_{t}^{T}\gamma_{s}^{-1}(u_{s}^{0})^{2}ds+2\int_{t}^{T}\gamma_{s}^{-1}(H^{0}_{s})^{2}ds<\infty\text{ a.s.}\end{split}

since γ\gamma and H0H^{0} have a.s. continuous paths and E[tT(us0)2𝑑s]<E[\int_{t}^{T}(u_{s}^{0})^{2}ds]<\infty. We are therefore able to define D0D^{0} by (14). Moreover, denote αs=γs12νs1\alpha_{s}=\gamma_{s}^{-\frac{1}{2}}\nu_{s}^{-1}, s[t,T]s\in[t,T], and βs=dγtxtsXr0d(νrγr)\beta_{s}=d-\gamma_{t}x-\int_{t}^{s}X^{0}_{r}d(\nu_{r}\gamma_{r}), s[t,T]s\in[t,T]. It follows from Lemma˜5.1 and γs12Xs0=Hs0us0-\gamma_{s}^{\frac{1}{2}}X^{0}_{s}=H^{0}_{s}-u^{0}_{s}, s[t,T)s\in[t,T), that for all s[t,T]s\in[t,T]

d(αsβs)=(Hs0us0)((μs+ρs12σsηsr¯s12σs2)ds+(σs+ηsr¯s)dWs1+ηs1r¯s2dWs2)+αsβs((ρs12μs+38σs2+12σsηsr¯s)ds+(ηsr¯s12σs)dWs1ηs1r¯s2dWs2).\begin{split}&d(\alpha_{s}\beta_{s})\\ &=(H^{0}_{s}-u^{0}_{s})\Bigg{(}\Big{(}\mu_{s}+\rho_{s}-\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}^{2}\Big{)}ds+\big{(}\sigma_{s}+\eta_{s}\overline{r}_{s}\big{)}dW_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}\\ &\quad+\alpha_{s}\beta_{s}\Bigg{(}\Big{(}-\rho_{s}-\frac{1}{2}\mu_{s}+\frac{3}{8}\sigma_{s}^{2}+\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}\Big{)}ds+\Big{(}-\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}\Big{)}dW_{s}^{1}-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}.\end{split}

We combine this with

dHs0=us0((μs+ρs12σsηsr¯s12σs2)ds+(σs+ηsr¯s)dWs1+ηs1r¯s2dWs2)+Hs0((12μs18σs2)ds+12σsdWs1),s[t,T],\begin{split}dH^{0}_{s}&=-u^{0}_{s}\Bigg{(}\Big{(}\mu_{s}+\rho_{s}-\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}^{2}\Big{)}ds+(\sigma_{s}+\eta_{s}\overline{r}_{s})dW_{s}^{1}+\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}\\ &\quad+H^{0}_{s}\Bigg{(}\Big{(}\frac{1}{2}\mu_{s}-\frac{1}{8}\sigma_{s}^{2}\Big{)}ds+\frac{1}{2}\sigma_{s}dW_{s}^{1}\Bigg{)},\quad s\in[t,T],\end{split}

to obtain for all s[t,T]s\in[t,T] that

d(αsβsHs0)=(αsβsHs0)((ρs12μs+38σs2+12σsηsr¯s)ds+(ηsr¯s12σs)dWs1ηs1r¯s2dWs2).\begin{split}d(\alpha_{s}\beta_{s}-H^{0}_{s})&=(\alpha_{s}\beta_{s}-H^{0}_{s})\Bigg{(}\Big{(}-\rho_{s}-\frac{1}{2}\mu_{s}+\frac{3}{8}\sigma_{s}^{2}+\frac{1}{2}\sigma_{s}\eta_{s}\overline{r}_{s}\Big{)}ds+\Big{(}-\eta_{s}\overline{r}_{s}-\frac{1}{2}\sigma_{s}\Big{)}dW_{s}^{1}\\ &\quad\quad\quad\qquad\qquad-\eta_{s}\sqrt{1-\overline{r}_{s}^{2}}dW_{s}^{2}\Bigg{)}.\end{split} (59)

Note that αtβt=γt12dγt12x=Ht0\alpha_{t}\beta_{t}=\gamma_{t}^{-\frac{1}{2}}d-\gamma_{t}^{\frac{1}{2}}x=H^{0}_{t}. We thus conclude that 0 is the unique solution of (59), and hence Hs0=γs12νs1(dγtxtsXr0d(νrγr))H^{0}_{s}=\gamma^{-\frac{1}{2}}_{s}\nu_{s}^{-1}(d-\gamma_{t}x-\int_{t}^{s}X^{0}_{r}d(\nu_{r}\gamma_{r})), s[t,T]s\in[t,T]. This implies that D0=γX0+γ12H0D^{0}=\gamma X^{0}+\gamma^{\frac{1}{2}}H^{0}, i.e., Ds0=γs12us0D^{0}_{s}=\gamma_{s}^{\frac{1}{2}}u^{0}_{s}, s[t,T)s\in[t,T), and DT0=γTξ+γT12HT0D^{0}_{T}=\gamma_{T}\xi+\gamma_{T}^{\frac{1}{2}}H^{0}_{T}. The fact that E[tT(us0)2𝑑s]<E[\int_{t}^{T}(u_{s}^{0})^{2}ds]<\infty then immediately yields that (A1) holds. This proves that X0𝒜tpm(x,d)X^{0}\in\mathcal{A}_{t}^{pm}(x,d). ∎

We finally are able to prove ˜1.7.

Proof of ˜1.7.

(i) Denote by DD, DnD^{n}, nn\in\mathbb{N}, the deviation processes associated to XX, XnX^{n}, nn\in\mathbb{N}, and let H¯\overline{H} and H¯n\overline{H}^{n}, nn\in\mathbb{N}, be the scaled hidden deviation processes. By Lemma˜1.6 it holds for all nn\in\mathbb{N} that

|Jtpm(x,d,Xn)Jtpm(x,d,X)|=|12Et[tTγs1((Dsn)2Ds2)2(κs+λs)ds]2Et[tTλsγs12(Dsn(H¯sn+γs12ζs)Ds(H¯s+γs12ζs))𝑑s]+Et[tTλs((H¯sn+γs12ζs)2(H¯s+γs12ζs)2)𝑑s]+12Et[(H¯Tn+γT12ξ)2(H¯T+γT12ξ)2]|.\begin{split}\left\lvert J^{pm}_{t}(x,d,X^{n})-J^{pm}_{t}(x,d,X)\right\rvert&=\Bigg{\lvert}\frac{1}{2}E_{t}\left[\int_{t}^{T}\gamma_{s}^{-1}\left((D_{s}^{n})^{2}-D_{s}^{2}\right)2(\kappa_{s}+\lambda_{s})ds\right]\\ &\quad-2E_{t}\!\left[\int_{t}^{T}\!\lambda_{s}\gamma_{s}^{-\frac{1}{2}}\!\left(D_{s}^{n}\big{(}\overline{H}_{s}^{n}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}\!-\!D_{s}\big{(}\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}\!\right)\!ds\right]\\ &\quad+E_{t}\left[\int_{t}^{T}\lambda_{s}\left(\big{(}\overline{H}_{s}^{n}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}^{2}-\big{(}\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}^{2}\right)ds\right]\\ &\quad+\frac{1}{2}E_{t}\left[(\overline{H}_{T}^{n}+\gamma_{T}^{\frac{1}{2}}\xi)^{2}-(\overline{H}_{T}+\gamma_{T}^{\frac{1}{2}}\xi)^{2}\right]\Bigg{\rvert}.\end{split}

Boundedness of λ,r¯,ρ,μ,η\lambda,\overline{r},\rho,\mu,\eta and σ\sigma implies (recall also (12)) that there exists some c(0,)c\in(0,\infty) such that for all nn\in\mathbb{N} it holds that

E[|Jtpm(x,d,Xn)Jtpm(x,d,X)|]E[|(H¯Tn+γT12ξ)2(H¯T+γT12ξ)2|]+cE[tT|γs1((Dsn)2Ds2)|𝑑s]+cE[tT|γs12(Dsn(H¯sn+γs12ζs)Ds(H¯s+γs12ζs))|𝑑s]+cE[tT|(H¯sn+γs12ζs)2(H¯s+γs12ζs)2|𝑑s].\begin{split}&E\left[\left\lvert J^{pm}_{t}(x,d,X^{n})-J^{pm}_{t}(x,d,X)\right\rvert\right]\\ &\leq E\left[\left\lvert(\overline{H}_{T}^{n}+\gamma_{T}^{\frac{1}{2}}\xi)^{2}-(\overline{H}_{T}+\gamma_{T}^{\frac{1}{2}}\xi)^{2}\right\rvert\right]+cE\left[\int_{t}^{T}\left\lvert\gamma_{s}^{-1}\left((D_{s}^{n})^{2}-D_{s}^{2}\right)\right\rvert ds\right]\\ &\quad+cE\left[\int_{t}^{T}\left\lvert\gamma_{s}^{-\frac{1}{2}}\left(D_{s}^{n}\big{(}\overline{H}_{s}^{n}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}-D_{s}\big{(}\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}\right)\right\rvert ds\right]\\ &\quad+cE\left[\int_{t}^{T}\left\lvert\big{(}\overline{H}_{s}^{n}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}^{2}-\big{(}\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}^{2}\right\rvert ds\right].\end{split} (60)

We treat the terminal costs first. It holds for all nn\in\mathbb{N} that

E[|(H¯Tn+γT12ξ)2(H¯T+γT12ξ)2|]=E[|(H¯Tn)2+2H¯TnγT12ξH¯T22H¯TγT12ξ|]E[|(H¯Tn)2H¯T2|]+2E[|(H¯TnH¯T)γT12ξ|]E[|(H¯Tn)2H¯T2|]+2(E[(H¯TnH¯T)2])12(E[γTξ2])12.\begin{split}&E\left[\left\lvert(\overline{H}_{T}^{n}+\gamma_{T}^{\frac{1}{2}}\xi)^{2}-(\overline{H}_{T}+\gamma_{T}^{\frac{1}{2}}\xi)^{2}\right\rvert\right]=E\left[\left\lvert(\overline{H}_{T}^{n})^{2}+2\overline{H}_{T}^{n}\gamma_{T}^{\frac{1}{2}}\xi-\overline{H}_{T}^{2}-2\overline{H}_{T}\gamma_{T}^{\frac{1}{2}}\xi\right\rvert\right]\\ &\qquad\qquad\qquad\qquad\leq E\left[\left\lvert(\overline{H}_{T}^{n})^{2}-\overline{H}_{T}^{2}\right\rvert\right]+2E\left[\left\lvert(\overline{H}^{n}_{T}-\overline{H}_{T})\gamma_{T}^{\frac{1}{2}}\xi\right\rvert\right]\\ &\qquad\qquad\qquad\qquad\leq E\left[\left\lvert(\overline{H}_{T}^{n})^{2}-\overline{H}_{T}^{2}\right\rvert\right]+2\left(E\left[(\overline{H}^{n}_{T}-\overline{H}_{T})^{2}\right]\right)^{\frac{1}{2}}\left(E\left[\gamma_{T}\xi^{2}\right]\right)^{\frac{1}{2}}.\end{split}

From

limnE[tT(DsnDs)2γs1𝑑s]=0\lim_{n\to\infty}E\left[\int_{t}^{T}(D_{s}^{n}-D_{s})^{2}\gamma_{s}^{-1}ds\right]=0 (61)

(cf. (20)) and Lemma˜5.3 we have that

limnE[sups[t,T]|H¯snH¯s|2]=0.\lim_{n\to\infty}E\left[\sup_{s\in[t,T]}\lvert\overline{H}_{s}^{n}-\overline{H}_{s}\rvert^{2}\right]=0. (62)

Since furthermore E[γTξ2]<E[\gamma_{T}\xi^{2}]<\infty, we obtain that limnE[|(H¯Tn+γT12ξ)2(H¯T+γT12ξ)2|]=0.\lim_{n\to\infty}E[\lvert(\overline{H}_{T}^{n}+\gamma_{T}^{\frac{1}{2}}\xi)^{2}-(\overline{H}_{T}+\gamma_{T}^{\frac{1}{2}}\xi)^{2}\rvert]=0. The second term in (60) converges to 0 using (61). For the third term in (60) we have for all nn\in\mathbb{N} that

E[tT|γs12(Dsn(H¯sn+γs12ζs)Ds(H¯s+γs12ζs))|𝑑s]E[tT(|H¯s+γs12ζs||DsnDs|γs12+γs12|Dsn||H¯snH¯s|)𝑑s](E[tT(H¯s+γs12ζs)2𝑑s])12(E[tT(DsnDs)2γs1𝑑s])12+(E[tTγs1(Dsn)2𝑑s])12T12(E[sups[t,T]|H¯snH¯s|2])12.\begin{split}&E\left[\int_{t}^{T}\left\lvert\gamma_{s}^{-\frac{1}{2}}\left(D_{s}^{n}\big{(}\overline{H}_{s}^{n}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}-D_{s}\big{(}\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{)}\right)\right\rvert ds\right]\\ &\leq E\left[\int_{t}^{T}\left(\big{\lvert}\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s}\big{\rvert}\,\lvert D_{s}^{n}-D_{s}\rvert\gamma_{s}^{-\frac{1}{2}}+\gamma_{s}^{-\frac{1}{2}}\lvert D_{s}^{n}\rvert\,\lvert\overline{H}_{s}^{n}-\overline{H}_{s}\rvert\right)ds\right]\\ &\leq\left(E\bigg{[}\int_{t}^{T}\big{(}\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s})^{2}ds\bigg{]}\right)^{\frac{1}{2}}\left(E\bigg{[}\int_{t}^{T}(D_{s}^{n}-D_{s})^{2}\gamma_{s}^{-1}ds\bigg{]}\right)^{\frac{1}{2}}\\ &\quad\,+\left(E\bigg{[}\int_{t}^{T}\gamma_{s}^{-1}(D_{s}^{n})^{2}ds\bigg{]}\right)^{\frac{1}{2}}T^{\frac{1}{2}}\left(E\bigg{[}\sup_{s\in[t,T]}\lvert\overline{H}_{s}^{n}-\overline{H}_{s}\rvert^{2}\bigg{]}\right)^{\frac{1}{2}}.\end{split} (63)

By Lemma˜1.6 and (5) it holds that E[tT(H¯s+γs12ζs)2𝑑s]<E[\int_{t}^{T}(\overline{H}_{s}+\gamma_{s}^{\frac{1}{2}}\zeta_{s})^{2}ds]<\infty. Moreover, due to (61), we have that E[tTγs1(Dsn)2𝑑s]E[\int_{t}^{T}\gamma_{s}^{-1}(D_{s}^{n})^{2}ds] is uniformly bounded in nn\in\mathbb{N}. It thus follows from (61), (62) and (63) that the third term in (60) converges to 0 as nn\to\infty. The last term in (60) converges to 0 using (5) and (62). This proves claim (i).

(ii) Suppose that X𝒜tpm(x,d)X\in\mathcal{A}^{pm}_{t}(x,d). Let u=(us)s[t,T]u=(u_{s})_{s\in[t,T]} be defined by us=γs12Dsu_{s}=\gamma_{s}^{-\frac{1}{2}}D_{s}, s[t,T]s\in[t,T], where DD denotes the deviation associated to XX. Then, uu is a progressively measurable process, and due to assumption (A1) it holds that E[tTus2𝑑s]<E[\int_{t}^{T}u_{s}^{2}ds]<\infty. By Lemma˜5.4 there exists a sequence of bounded càdlàg finite variation processes (vn)n(v^{n})_{n\in\mathbb{N}} such that limnE[tT(usZsvsn)2Zs2𝑑s]=0\lim_{n\to\infty}E[\int_{t}^{T}(\frac{u_{s}}{Z_{s}}-v_{s}^{n})^{2}Z_{s}^{2}ds]=0, where ZZ is defined in (56). Set un=vnZu^{n}=v^{n}Z, nn\in\mathbb{N}. This is a sequence of càdlàg semimartingales in t2\mathcal{L}_{t}^{2} that satisfies limnuunt2=0\lim_{n\to\infty}\lVert u-u^{n}\rVert_{\mathcal{L}_{t}^{2}}=0. Moreover, it holds for all nn\in\mathbb{N} and any p2p\geq 2 that E[sups[t,T]|usn|p]<E[\sup_{s\in[t,T]}\lvert u_{s}^{n}\rvert^{p}]<\infty. For each unu^{n}, nn\in\mathbb{N}, let Hn=(Hsn)s[t,T]H^{n}=(H^{n}_{s})_{s\in[t,T]} be the solution of (58). We then define a sequence of càdlàg semimartingales Xn=(Xsn)s[t,T]X^{n}=(X^{n}_{s})_{s\in[t-,T]}, nn\in\mathbb{N}, by Xsn=γs12(usnHsn)X^{n}_{s}=\gamma_{s}^{-\frac{1}{2}}(u_{s}^{n}-H_{s}^{n}), s[t,T)s\in[t,T), Xtn=xX_{t-}^{n}=x, XTn=ξX_{T}^{n}=\xi. By Lemma˜5.5 we have for all nn\in\mathbb{N} that Xn𝒜tpm(x,d)X^{n}\in\mathcal{A}_{t}^{pm}(x,d) and that Dn=γXn+γ12HnD^{n}=\gamma X^{n}+\gamma^{\frac{1}{2}}H^{n} for the associated deviation process Dn=(Dsn)s[t,T]D^{n}=(D^{n}_{s})_{s\in[t-,T]}. It follows for all nn\in\mathbb{N} that Dsn=γs12usnD_{s}^{n}=\gamma_{s}^{\frac{1}{2}}u_{s}^{n}, s[t,T)s\in[t,T). Therefore, it holds for all nn\in\mathbb{N} that

𝐝(Xn,X)=(E[tT(DsnDs)2γs1𝑑s])12=(E[tT(usnus)2𝑑s])12.\mathbf{d}(X^{n},X)=\left(E\left[\int_{t}^{T}(D_{s}^{n}-D_{s})^{2}\gamma_{s}^{-1}ds\right]\right)^{\frac{1}{2}}=\left(E\left[\int_{t}^{T}(u_{s}^{n}-u_{s})^{2}ds\right]\right)^{\frac{1}{2}}.

Due to limnuunt2=0\lim_{n\to\infty}\lVert u-u^{n}\rVert_{\mathcal{L}_{t}^{2}}=0, we thus have that limn𝐝(Xn,X)=0\lim_{n\to\infty}\mathbf{d}(X^{n},X)=0. We next show that for all nn\in\mathbb{N}, XnX^{n} has finite variation. To this end, we observe that for all nn\in\mathbb{N} and s[t,T)s\in[t,T) it holds by integration by parts that

dXsn=γs12d(usnHsn)+(usnHsn)dγs12+d[γ12,unHn]s.dX_{s}^{n}=\gamma_{s}^{-\frac{1}{2}}d(u_{s}^{n}-H_{s}^{n})+(u_{s}^{n}-H_{s}^{n})d\gamma_{s}^{-\frac{1}{2}}+d[\gamma^{-\frac{1}{2}},u^{n}-H^{n}]_{s}. (64)

Again by integration by parts, and using (57), we have for all nn\in\mathbb{N} and s[t,T]s\in[t,T] that

dusn=vsndZs+Zsdvsn+d[vn,Z]s=12usn((12σs+ηsr¯s)2+ηs2(1r¯s2))dsusn(12σs+ηsr¯s)dWs1usnηs1r¯s2dWs2+Zsdvsn.\begin{split}du_{s}^{n}&=v_{s}^{n}dZ_{s}+Z_{s}dv_{s}^{n}+d[v^{n},Z]_{s}\\ &=\!\frac{1}{2}u_{s}^{n}\bigg{(}\!\Big{(}\frac{1}{2}\sigma_{s}\!+\!\eta_{s}\overline{r}_{s}\!\Big{)}^{\!2}\!\!+\!\eta_{s}^{2}(1-\overline{r}_{s}^{2})\!\bigg{)}\!ds-\!u_{s}^{n}\Big{(}\frac{1}{2}\sigma_{s}\!+\!\eta_{s}\overline{r}_{s}\!\Big{)}\!dW_{s}^{1}\!-\!u_{s}^{n}\eta_{s}\sqrt{1\!-\!\overline{r}_{s}^{2}}dW_{s}^{2}\!+\!Z_{s}dv_{s}^{n}.\end{split}

This and (58) yield for all nn\in\mathbb{N} and s[t,T]s\in[t,T] that

γs12d(usnHsn)=γs12(ρs+μs+12ηs238σs2)usndsγs12(12μs18σs2)Hsnds+γs1212σs(usnHsn)dWs1+γs12Zsdvsn.\begin{split}\gamma_{s}^{-\frac{1}{2}}d(u_{s}^{n}-H_{s}^{n})&=\gamma_{s}^{-\frac{1}{2}}\left(\rho_{s}+\mu_{s}+\frac{1}{2}\eta_{s}^{2}-\frac{3}{8}\sigma_{s}^{2}\right)u_{s}^{n}ds-\gamma_{s}^{-\frac{1}{2}}\left(\frac{1}{2}\mu_{s}-\frac{1}{8}\sigma_{s}^{2}\right)H_{s}^{n}ds\\ &\quad+\gamma_{s}^{-\frac{1}{2}}\frac{1}{2}\sigma_{s}(u_{s}^{n}-H_{s}^{n})dW_{s}^{1}+\gamma_{s}^{-\frac{1}{2}}Z_{s}dv_{s}^{n}.\end{split} (65)

Moreover, it follows from (44) for all nn\in\mathbb{N} and s[t,T]s\in[t,T] that

(usnHsn)dγs12=(usnHsn)γs12(12μs+38σs2)ds(usnHsn)γs1212σsdWs1.\begin{split}(u_{s}^{n}-H_{s}^{n})d\gamma_{s}^{-\frac{1}{2}}&=(u_{s}^{n}-H_{s}^{n})\gamma_{s}^{-\frac{1}{2}}\left(-\frac{1}{2}\mu_{s}+\frac{3}{8}\sigma_{s}^{2}\right)ds-(u_{s}^{n}-H_{s}^{n})\gamma_{s}^{-\frac{1}{2}}\frac{1}{2}\sigma_{s}dW_{s}^{1}.\end{split} (66)

We combine (64), (65), and (66) to obtain for all nn\in\mathbb{N} and s(t,T)s\in(t,T) that

dXsn=γs12usn(ρs+12μs+12ηs2)dsγs12Hsn14σs2ds+γs12Zsdvsn+d[γ12,unHn]s.\begin{split}dX_{s}^{n}&=\gamma_{s}^{-\frac{1}{2}}u_{s}^{n}\left(\rho_{s}+\frac{1}{2}\mu_{s}+\frac{1}{2}\eta_{s}^{2}\right)ds-\gamma_{s}^{-\frac{1}{2}}H_{s}^{n}\frac{1}{4}\sigma_{s}^{2}ds+\gamma_{s}^{-\frac{1}{2}}Z_{s}dv_{s}^{n}+d[\gamma^{-\frac{1}{2}},u^{n}-H^{n}]_{s}.\end{split}

Since vnv^{n} has finite variation for all nn\in\mathbb{N}, this representation shows that also XnX^{n} has finite variation for all nn\in\mathbb{N}. Note that for all nn\in\mathbb{N}, by ˜1.3, the process (6) associated to the càdlàg finite variation process XnX^{n} is nothing but DnD^{n}. Since η\eta is bounded, there exists c(0,)c\in(0,\infty) such that for all nn\in\mathbb{N}

E[(tT(Dsn)4γs2ηs2𝑑s)12]=E[(tT(usn)4ηs2𝑑s)12]cE[sups[t,T](usn)2]<.\begin{split}E\left[\left(\int_{t}^{T}(D_{s}^{n})^{4}\gamma_{s}^{-2}\eta_{s}^{2}ds\right)^{\frac{1}{2}}\right]&=E\left[\left(\int_{t}^{T}(u_{s}^{n})^{4}\eta_{s}^{2}ds\right)^{\frac{1}{2}}\right]\leq cE[\sup_{s\in[t,T]}(u_{s}^{n})^{2}]<\infty.\end{split}

This implies (A2). Similarly, by boundedness of σ\sigma, we obtain (A3). We thus conclude that Xn𝒜tfv(x,d)X^{n}\in\mathcal{A}_{t}^{fv}(x,d) for all nn\in\mathbb{N}.

(iii) Let (Xn)n(X^{n})_{n\in\mathbb{N}} be a Cauchy sequence in (𝒜tpm(x,d),𝐝)(\mathcal{A}_{t}^{pm}(x,d),\mathbf{d}). For nn\in\mathbb{N} we denote by DnD^{n} the deviation process associated to XnX^{n}. It then holds that (γ12Dn)n(\gamma^{-\frac{1}{2}}D^{n})_{n\in\mathbb{N}} is a Cauchy sequence in (t2,t2)(\mathcal{L}_{t}^{2},\lVert\cdot\rVert_{\mathcal{L}_{t}^{2}}). Since (t2,t2)(\mathcal{L}_{t}^{2},\lVert\cdot\rVert_{\mathcal{L}_{t}^{2}}) is complete (see, e.g., Lemma 2.2 in Section 3.2 of [33]), there exists u0t2u^{0}\in\mathcal{L}_{t}^{2} such that limnγ12Dnu0t2=0\lim_{n\to\infty}\lVert\gamma^{-\frac{1}{2}}D^{n}-u^{0}\rVert_{\mathcal{L}_{t}^{2}}=0. Define X0=(Xs0)s[t,T]X^{0}=(X^{0}_{s})_{s\in[t-,T]} by Xt0=xX_{t-}^{0}=x, XT0=ξX_{T}^{0}=\xi, Xs0=γs12(us0Hs0)X_{s}^{0}=\gamma_{s}^{-\frac{1}{2}}(u_{s}^{0}-H_{s}^{0}), s[t,T)s\in[t,T), where H0H^{0} is given by (58). By Lemma˜5.5 it holds that X0𝒜tpm(x,d)X^{0}\in\mathcal{A}_{t}^{pm}(x,d). We furthermore obtain from Lemma˜5.5 that, for the associated deviation, D0=γX0+γ12H0D^{0}=\gamma X^{0}+\gamma^{\frac{1}{2}}H^{0}. By definition of X0X^{0}, this yields γs12Ds0=us0\gamma_{s}^{-\frac{1}{2}}D_{s}^{0}=u_{s}^{0}, s[t,T)s\in[t,T). It follows that

𝐝(Xn,X0)=(E[tT(γs12Dsnγs12Ds0)2𝑑s])12=γ12Dnu0t2,\begin{split}\mathbf{d}(X^{n},X^{0})&=\left(E\left[\int_{t}^{T}(\gamma_{s}^{-\frac{1}{2}}D^{n}_{s}-\gamma_{s}^{-\frac{1}{2}}D_{s}^{0})^{2}ds\right]\right)^{\frac{1}{2}}=\lVert\gamma^{-\frac{1}{2}}D^{n}-u^{0}\rVert_{\mathcal{L}_{t}^{2}},\end{split}

and hence limn𝐝(Xn,X0)=0\lim_{n\to\infty}\mathbf{d}(X^{n},X^{0})=0. ∎

Proof of Lemma˜2.1.

By definition of uu we have that uu is progressively measurable and, due to assumption (A1), satisfies E[tTus2𝑑s]<E[\int_{t}^{T}u_{s}^{2}ds]<\infty; hence, ut2u\in\mathcal{L}_{t}^{2}.

Let H¯s=γs12Dsγs12Xs\overline{H}_{s}=\gamma_{s}^{-\frac{1}{2}}D_{s}-\gamma_{s}^{\frac{1}{2}}X_{s}, s[t,T]s\in[t,T], be the scaled hidden deviation (17) associated to XX. We can substitute u=γ12Du=\gamma^{-\frac{1}{2}}D in the cost functional (19) and also in the dynamics (18) of H¯\overline{H}. Observe that H¯\overline{H} follows the same dynamics as the state process H~\widetilde{H} associated to uu (see (22)), and that H¯t=dγtγtx=H~t\overline{H}_{t}=\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x=\widetilde{H}_{t}. Therefore, H¯\overline{H} and H~\widetilde{H} coincide, which completes the proof. ∎

Proof of Lemma˜2.2.

It follows from Lemma˜5.5 that X𝒜tpm(x,d)X\in\mathcal{A}_{t}^{pm}(x,d). Moreover, we have from Lemma˜5.5 that the associated deviation satisfies D=γX+γ12H~D=\gamma X+\gamma^{\frac{1}{2}}\widetilde{H}, i.e., Ds=γs12usD_{s}=\gamma_{s}^{\frac{1}{2}}u_{s}, s[t,T)s\in[t,T), and H~\widetilde{H} is the scaled hidden deviation of XX. It thus holds that Jtpm(x,d,X)J^{pm}_{t}(x,d,X) is given by (19). In the definition (23) of JJ, we may replace uu under the integrals with respect to the Lebesgue measure by γ12D\gamma^{-\frac{1}{2}}D. This shows that Jtpm(x,d,X)=Jt(dγtγtx,u)d22γtJ^{pm}_{t}(x,d,X)=J_{t}(\frac{d}{\sqrt{\gamma_{t}}}-\sqrt{\gamma_{t}}x,u)-\frac{d^{2}}{2\gamma_{t}}. ∎

Proof of Lemma˜2.5.

(i) We have that u^\hat{u} is progressively measurable. Furthermore, the facts that E[tTus2𝑑s]<E[\int_{t}^{T}u_{s}^{2}ds]<\infty, E[sups[t,T]H~s2]<E[\sup_{s\in[t,T]}\widetilde{H}_{s}^{2}]<\infty, E[0Tγsζs2𝑑s]<E[\int_{0}^{T}\gamma_{s}\zeta_{s}^{2}ds]<\infty, and (24) imply that E[tTu^s2𝑑s]<E[\int_{t}^{T}\hat{u}_{s}^{2}ds]<\infty. Hence, u^t2\hat{u}\in\mathcal{L}_{t}^{2}. Substituting us=u^s+λsλs+κs(H~s+γsζs)u_{s}=\hat{u}_{s}+\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\widetilde{H}_{s}+\sqrt{\gamma_{s}}\zeta_{s}), s[t,T]s\in[t,T], in (22) leads to (25). For the cost functional, observe that

12(2λs(H~s+γsζs)24λs(H~s+γsζs)us)+(κs+λs)us2=λs(H~s+γsζs)2(λs+κs)λs2(λs+κs)2(H~s+γsζs)2+(λs+κs)(usλsλs+κs(H~s+γsζs))2=λsκsλs+κs(H^s+γsζs)2+(λs+κs)u^s2,s[t,T].\begin{split}&\frac{1}{2}\Big{(}2\lambda_{s}(\widetilde{H}_{s}+\sqrt{\gamma_{s}}\zeta_{s})^{2}-4\lambda_{s}(\widetilde{H}_{s}+\sqrt{\gamma_{s}}\zeta_{s})u_{s}\Big{)}+(\kappa_{s}+\lambda_{s})u_{s}^{2}\\ &=\lambda_{s}(\widetilde{H}_{s}+\sqrt{\gamma_{s}}\zeta_{s})^{2}-(\lambda_{s}+\kappa_{s})\frac{\lambda_{s}^{2}}{\left(\lambda_{s}+\kappa_{s}\right)^{2}}(\widetilde{H}_{s}+\sqrt{\gamma_{s}}\zeta_{s})^{2}\\ &\quad+(\lambda_{s}+\kappa_{s})\left(u_{s}-\frac{\lambda_{s}}{\lambda_{s}+\kappa_{s}}(\widetilde{H}_{s}+\sqrt{\gamma_{s}}\zeta_{s})\right)^{2}\\ &=\frac{\lambda_{s}\kappa_{s}}{\lambda_{s}+\kappa_{s}}\left(\widehat{H}_{s}+\sqrt{\gamma_{s}}\zeta_{s}\right)^{2}+(\lambda_{s}+\kappa_{s})\hat{u}_{s}^{2},\quad s\in[t,T].\end{split} (67)

(ii) Note that (25) is an SDE that is linear in H^\widehat{H}, u^\hat{u}, and γζ\sqrt{\gamma}\zeta. Furthermore, boundedness of ρ,μ,σ,η,r¯\rho,\mu,\sigma,\eta,\overline{r} and (24) imply that the coefficients of the SDE are bounded. Since moreover E[tT(u^s)2+γsζs2ds]<E[\int_{t}^{T}(\hat{u}_{s})^{2}+\gamma_{s}\zeta_{s}^{2}ds]<\infty and H^t\widehat{H}_{t} is square integrable, we know that E[sups[t,T]H^s2]<E[\sup_{s\in[t,T]}\widehat{H}_{s}^{2}]<\infty (see, e.g., [41, Theorem 3.2.2 and Theorem 3.3.1]). We can thus argue similar to (i) that ut2u\in\mathcal{L}_{t}^{2}. A substitution of u^\hat{u} in (25) yields (22). A reverse version of the argument in (67) proves equality of the cost functionals. ∎


Acknowledgement: We thank Dirk Becherer, Tiziano De Angelis, Miryana Grigorova, Martin Herdegen, and Yuri Kabanov for inspiring discussions. We are grateful to the associate editor and two anonymous referees for constructive comments and suggestions that helped us improve the manuscript.

References

  • [1] J. Ackermann, T. Kruse, and M. Urusov. Càdlàg semimartingale strategies for optimal trade execution in stochastic order book models. Finance Stoch., 25(4):757–810, 2021.
  • [2] J. Ackermann, T. Kruse, and M. Urusov. Optimal trade execution in an order book model with stochastic liquidity parameters. SIAM Journal on Financial Mathematics, 12(2):788–822, 2021.
  • [3] J. Ackermann, T. Kruse, and M. Urusov. Self-exciting price impact via negative resilience in stochastic order books. Annals of Operations Research, DOI:10.1007/s10479-022-04973-00, 2022.
  • [4] A. Alfonsi and J. I. Acevedo. Optimal execution and price manipulations in time-varying limit order books. Applied Mathematical Finance, 21(3):201–237, 2014.
  • [5] A. Alfonsi, A. Fruth, and A. Schied. Constrained portfolio liquidation in a limit order book model. Banach Center Publ, 83:9–25, 2008.
  • [6] A. Alfonsi, A. Fruth, and A. Schied. Optimal execution strategies in limit order books with general shape functions. Quantitative Finance, 10(2):143–157, 2010.
  • [7] A. Alfonsi and A. Schied. Optimal trade execution and absence of price manipulations in limit order book models. SIAM Journal on Financial Mathematics, 1(1):490–522, 2010.
  • [8] A. Alfonsi, A. Schied, and A. Slynko. Order book resilience, price manipulation, and the positive portfolio problem. SIAM Journal on Financial Mathematics, 3(1):511–533, 2012.
  • [9] S. Ankirchner, M. Jeanblanc, and T. Kruse. BSDEs with singular terminal condition and a control problem with constraints. SIAM Journal on Control and Optimization, 52(2):893–913, 2014.
  • [10] S. Ankirchner and T. Kruse. Optimal position targeting with stochastic linear-quadratic costs. Advances in mathematics of finance, 104:9–24, 2015.
  • [11] P. Bank and Y. Dolinsky. Continuous-time duality for superreplication with transient price impact. Ann. Appl. Probab., 29(6):3893–3917, 2019.
  • [12] P. Bank and Y. Dolinsky. Scaling limits for super-replication with transient price impact. Bernoulli, 26(3):2176–2201, 2020.
  • [13] P. Bank and A. Fruth. Optimal order scheduling for deterministic liquidity patterns. SIAM Journal on Financial Mathematics, 5(1):137–152, 2014.
  • [14] P. Bank, H. M. Soner, and M. Voß. Hedging with temporary price impact. Math. Financ. Econ., 11(2):215–239, 2017.
  • [15] P. Bank and M. Voß. Linear quadratic stochastic control problems with stochastic terminal constraint. SIAM Journal on Control and Optimization, 56(2):672–699, 2018.
  • [16] P. Bank and M. Voß. Optimal investment with transient price impact. SIAM J. Financial Math., 10(3):723–768, 2019.
  • [17] D. Becherer, T. Bilarev, and P. Frentrup. Optimal asset liquidation with multiplicative transient price impact. Applied Mathematics & Optimization, 78(3):643–676, 2018.
  • [18] D. Becherer, T. Bilarev, and P. Frentrup. Optimal liquidation under stochastic liquidity. Finance and Stochastics, 22(1):39–68, 2018.
  • [19] D. Becherer, T. Bilarev, and P. Frentrup. Stability for gains from large investors’ strategies in M1M_{1}/J1J_{1} topologies. Bernoulli, 25(2):1105–1140, 2019.
  • [20] J.-M. Bismut. Linear quadratic optimal stochastic control with random coefficients. SIAM Journal on Control and Optimization, 14(3):419–444, 1976.
  • [21] J.-M. Bismut. Contrôle des systemes linéaires quadratiques: applications de l’intégrale stochastique. In Séminaire de Probabilités XII, pages 180–264. Springer, 1978.
  • [22] R. Carmona and K. Webster. The self-financing equation in limit order book markets. Finance Stoch., 23(3):729–759, 2019.
  • [23] Y. Dolinsky, B. Gottesman, and O. Gurel-Gurevich. A note on costs minimization with stochastic target constraints. Electronic Communications in Probability, 25, 2020.
  • [24] A. Fruth, T. Schöneborn, and M. Urusov. Optimal trade execution and price manipulation in order books with time-varying liquidity. Math. Finance, 24(4):651–695, 2014.
  • [25] A. Fruth, T. Schöneborn, and M. Urusov. Optimal trade execution in order books with stochastic liquidity. Math. Finance, 29(2):507–541, 2019.
  • [26] N. Gârleanu and L. H. Pedersen. Dynamic portfolio choice with frictions. Journal of Economic Theory, 165:487–516, 2016.
  • [27] J. Gatheral, A. Schied, and A. Slynko. Transient linear price impact and Fredholm integral equations. Mathematical Finance: An International Journal of Mathematics, Statistics and Financial Economics, 22(3):445–474, 2012.
  • [28] P. Graewe and U. Horst. Optimal trade execution with instantaneous price impact and stochastic resilience. SIAM Journal on Control and Optimization, 55(6):3707–3725, 2017.
  • [29] U. Horst and E. Kivman. Optimal trade execution under small market impact and portfolio liquidation with semimartingale strategies. Preprint, arXiv:2103.05957, 2021.
  • [30] U. Horst and F. Naujokat. When to cross the spread? Trading in two-sided limit order books. SIAM J. Financial Math., 5(1):278–315, 2014.
  • [31] U. Horst and X. Xia. Multi-dimensional optimal trade execution under stochastic resilience. Finance and Stochastics, 23(4):889–923, 2019.
  • [32] P. Imkeller and G. Dos Reis. Path regularity and explicit convergence rate for BSDE with truncated quadratic growth. Stochastic Processes and their Applications, 120(3):348–379, 2010.
  • [33] I. Karatzas and S. E. Shreve. Brownian motion and stochastic calculus. Graduate texts in mathematics; 113. Springer, New York, 2nd edition, 1991.
  • [34] M. Kohlmann and S. Tang. Global adapted solution of one-dimensional backward stochastic Riccati equations, with application to the mean–variance hedging. Stochastic Process. Appl., 97(2):255–288, 2002.
  • [35] C. Lorenz and A. Schied. Drift dependence of optimal trade execution strategies under transient price impact. Finance and Stochastics, 17(4):743–770, 2013.
  • [36] F. Naujokat and N. Westray. Curve following in illiquid markets. Math. Financ. Econ., 4(4):299–335, 2011.
  • [37] A. A. Obizhaeva and J. Wang. Optimal trading strategy and supply/demand dynamics. Journal of Financial Markets, 16:1–32, 2013.
  • [38] S. Predoiu, G. Shaikhet, and S. Shreve. Optimal execution in a general one-sided limit-order book. SIAM Journal on Financial Mathematics, 2(1):183–212, 2011.
  • [39] C. C. Siu, I. Guo, S.-P. Zhu, and R. J. Elliott. Optimal execution with regime-switching market resilience. J. Econom. Dynam. Control, 101:17–40, 2019.
  • [40] J. Sun, J. Xiong, and J. Yong. Indefinite stochastic linear-quadratic optimal control problems with random coefficients: Closed-loop representation of open-loop optimal controls. Ann. Appl. Probab., 31(1):460 – 499, 2021.
  • [41] J. Zhang. Backward Stochastic Differential Equations. Probability Theory and Stochastic Modelling; 86. Springer, New York, 1st edition, 2017.