Reducing Obizhaeva-Wang type trade execution problems to LQ stochastic control problems
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 and let be a filtered probability space satisfying the usual conditions. Let be an -measurable random variable and let be a progressively measurable process both satisfying suitable integrability assumptions (see (5) below). Further, let be a bounded progressively measurable process. Let be a positive Itô process driven by some Brownian motion and an Itô process driven by a (stochastically) correlated Brownian motion (see (3) and (4) below). Throughout the introduction we fix , and denote by the set of all adapted, càdlàg, finite variation processes satisfying , , and appropriate integrability assumptions (see (A1)–(A3) below). To each we associate a process satisfying
(1) |
We consider the finite variation stochastic control problem of minimizing the cost functional
(2) |
over , where is a shorthand notation for .
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 a position ( meaning a long position of shares of a stock and a short position of shares) of a certain financial asset. The investor trades the asset during the period in such a way that at each time the position is given by the value of the adapted, càdlàg, finite variation process (satisfying ). More precisely, represents the position immediately prior to the trade at time , while is the position immediately after that trade. The investor’s goal is to reach the target position
during the course of the trading period . Note that we allow 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 . Such situations may for example be faced by airline companies buying on forward markets the kerosene they need in 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 consists of an unaffected price plus a deviation that is caused by the investor’s trades during . We assume that the unaffected price process is a càdlàg martingale satisfying appropriate integrability conditions. Then integration by parts and the martingale property of ensure that expected trading costs due to are given by
Thus, these costs do not depend on the investor’s trading strategy and are therefore neglected in the sequel (we refer to Remark 2.2 in [1] for a more detailed discussion in the case ). The deviation process associated to is given by (1). Informally speaking, we see from (1) that a trade of size at time impacts by . So, the factor determines how strongly the price reacts to trades, and the process is therefore called the price impact process. In particular, the fact that is nonnegative entails that a buy trade leads to higher prices whereas a sell trade leads to smaller prices. The second component in the dynamics (1) describes the behavior of when the investor is not trading. Typically, it is assumed that is an increasing process such that in the absence of trades is reverting to with relative rate . Therefore, is called the resilience process. We refer to [3] for a discussion of the effects of “negative” resilience, where might also be decreasing. We highlight that in the present paper we allow to have a diffusive part. In summary, we note that the deviation prior to a trade of the investor at time is given by whereas it is equal to afterwards. We take the mean of these two values as the realized price per unit so that the investor’s overall trading costs due to amount to . This describes the first integral on the right-hand side of (2). Under the assumption that is nonnegative, the second integral can be understood as a risk term that penalizes any deviation of the position from the moving target in a quadratic way222The parametrization , , 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 , , as a weight and replace by in the subsequent assumptions and results.. A possible and natural choice would be , , so that the risk term ensures that any optimal strategy does not deviate too much from the (expected) target position 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 (see Section˜4.3 below for a specific example). In [1] the functional (2) was extended to a set of càdlàg semimartingales 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 of progressively measurable processes satisfying appropriate integrability conditions (see (A1) below) and the boundary conditions and . 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 as an integrator (see ˜1.3). It follows that the resulting alternative representation of (see ˜1.4) is not only well-defined on but even on , and we denote this extended functional by (see Section˜1.3). We next introduce a metric on and prove that is the unique continuous extension of from to (see ˜1.7). In particular, it follows that the infimum of over and the infimum of over coincide.
Next, for a given we identify the process , , as a useful tool in our analysis. Despite and having discontinuous paths in general, the process , which we call the scaled hidden deviation process, is always continuous. Moreover, we show that can be expressed in feedback form as an Itô process with coefficients that are linear in and (see Lemma˜1.6). Subsequently, we reinterpret the process as a control process and as the associated state process. Since the cost functional is quadratic in and , we arrive at a standard LQ stochastic control problem (see (22) and (23)) whose minimal costs coincide with the infimum of over (see Corollary˜2.3). Importantly, there is a one-to-one correspondence between square integrable controls for this standard problem and strategies , which allows to recover the minimizer of from a minimizer 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 exists (and is unique). This optimal control 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 of appropriate progressively measurable processes .
(b) Problem (1)–(2) is rather general. In particular, it includes the following features:
-
•
Presence of random terminal and moving targets and ;
-
•
Price impact is a positive Itô process ;
-
•
Resilience444To expand on this point, it is worth noting that in our current parametrization, only processes with dynamics without a diffusive component were considered by now in the literature on optimal trade execution in Obizhaeva-Wang type models. Moreover, in most papers is assumed to be positive, that is, only the case of an increasing was extensively studied previously. is an Itô process acting as an integrator in (1).
(c) Via introducing the mentioned scaled hidden deviation process and reinterpreting the process as a control in an (a priori, different) stochastic control problem, the extended to problem is reduced to an explicitly solvable LQ stochastic control problem. Thus, a unique optimal execution strategy in 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 that is a positive constant and the resilience process given by with some positive constant . 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 ). 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 and resilience 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 , a possibly non-zero (random) terminal target , and a larger class of resilience processes (in [1], as in many previous papers, is assumed to have the dynamics , and 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 with 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 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 acts as integrator there. In contrast, our continuous extension needs to employ essentially different ideas since we want to consider the set 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 and 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 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 . 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 and , . We fix a filtered probability space satisfying the usual conditions and supporting an -dimensional Brownian motion with respect to the filtration .
We first fix some notation. For conditional expectations with respect to are denoted by . For and a càdlàg process a jump at time is denoted by . We follow the convention that, for , and a càdlàg semimartingale , jumps of the càdlàg integrator at time contribute to integrals of the form . In contrast, we write when we do not include jumps of at time into the integral. The notation is sometimes used for continuous integrators . For and let . For every we mean by the space of all real-valued -measurable random variables such that . For , let denote the space of all (equivalence classes of) real-valued progressively measurable processes such that .
The control problem we are about to set up requires as input the real-valued, -measurable random variable and the real-valued, progressively measurable processes , , , , , and . We suppose that , , , and are -a.e. bounded. Moreover, we assume that is -valued. We define by , , and refer to as the correlation process. The processes and give rise to the continuous semimartingale with
(3) |
which is called the resilience process. We use the processes and to define the positive continuous semimartingale by
(4) |
with deterministic initial value . We refer to as the price impact process. Finally, we assume that and satisfy the integrability conditions
(5) |
Remark 1.1.
Note that the components 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 is generated by . The components 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 and we associate to an adapted, càdlàg, finite variation process a process defined by
(6) |
If there is no risk of confusion we sometimes simply write instead of in the sequel. For , we denote by the set of all adapted, càdlàg, finite variation processes satisfying , , and
-
(A1)
,
-
(A2)
,
-
(A3)
.
Any element is called a finite variation execution strategy. The process defined via (6) is called the associated deviation process.
1.2 Alternative representations for the cost functional and the deviation process
For we introduce an auxiliary process . It is defined to be the solution of
(8) |
Observe that the inverse is given by
(9) |
Remark 1.2.
Proposition 1.3.
Let and . Suppose that is an adapted, càdlàg, finite variation process with and with associated process defined by (6). It then holds that
(10) |
and
(11) |
As a consequence of ˜1.3, and relying on (A1)–(A3), we can rewrite the cost functional 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 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 , hence , can have nonvanishing quadratic variation. Here we, essentially, express everything not through but rather through , which has finite variation by Remark 1.2 (as 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 defined by
(12) |
1.3 Progressively measurable execution strategies
We point out that the right-hand side of (13) is also well-defined for progressively measurable processes satisfying an appropriate integrability condition and with associated deviation defined by (11) for which one assumes (A1). This motivates the following extension of the setting from Section˜1.1.
For , and a progressively measurable process such that a.s. and , we define the process by
(14) |
(recall from (8)). Notice that the condition a.s. ensures that the stochastic integral in (14) is well-defined. Again, we sometimes write instead of . Further, for , , let be the set of (equivalence classes of) progressively measurable processes with and that satisfy a.s. and such that condition (A1) holds true for defined by (14). To be precise, we stress that the equivalence classes for are understood with respect to the equivalence relation
(15) |
Any element is called a progressively measurable execution strategy. Again the process now defined via (14) is called the associated deviation process. Clearly, we have that .
Given , , and with associated (see (14)), we define the cost functional by
(16) |
1.4 The hidden deviation process
For , , and with associated deviation process , we define by , . Observe that if the investor followed a finite variation execution strategy until time and then decided to sell units of the asset ( means buying) at time , then by (6) the resulting deviation at time would equal . The value of hence represents the hypothetical deviation if the investor decides to close the position at time . We therefore call the hidden deviation process. Despite and in general being discontinuous, the hidden deviation process is always continuous. This can be seen from (14) and the fact that (hence also ) and are continuous. In the case of a finite variation execution strategy , it holds that , . In particular, the infinitesimal change of the hidden deviation is driven by the changes of the resilience process and the price impact process.
For , , and , 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 is quantity (of shares), while both and are measured in $. Thus, the scaled hidden deviation is measured in . defined by
(17) |
Also for and we sometimes simply write and , respectively. Note that, due to (14), it holds that , .
We next show that the scaled hidden deviation process satisfies a linear SDE and an -bound. Moreover, we derive a representation of in terms of the scaled hidden deviation process.
Lemma 1.6.
Let , , and . Then it holds that
(18) |
that , and that
(19) |
1.5 Continuous extension of the cost functional
Corollary˜1.5 states that for finite variation execution strategies, the cost functionals and are the same. In this subsection we show that can be considered as an extension of to progressively measurable strategies; i.e., we introduce a metric on and show that is continuous in the strategy (the first part of ˜1.7), that is dense in (the second part of ˜1.7) and that the metric space is complete (the third part of ˜1.7). The first and the second parts of ˜1.7 mean that, under the metric , is a unique continuous extension of from onto . The third part of ˜1.7 means that, under the metric , is the largest space where such a continuous extension is uniquely determined by on . This is because the completeness of is equivalent to the following statement: For any metric space containing and such that , it holds that the set is closed in .
For , , and with associated deviation processes , defined by (14), we define
(20) |
Identifying any processes that are equal -a.e., this indeed is a metric on , see Lemma˜5.2.
Note that, for fixed and , we may consider the cost functional (16) as a function Indeed, using (A1), Lemma˜1.6, (5), and boundedness of the input processes, we see that for all .
Theorem 1.7.
Let and .
(i) Suppose that . For every sequence in with it holds that .
(ii) For any there exists a sequence in such that . In particular, it holds that
(21) |
(iii) For any Cauchy sequence in there exists some such that .
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 over 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 , , and the costs depend in a quadratic way on . Moreover, (18) in Lemma˜1.6 ensures that the dynamics of depend linearly on . These two observations suggest to view the minimization problem of over as a standard LQ stochastic control problem with state process and control . This motivates the following definitions. For every , , and , we consider the state process defined by
(22) |
and the cost functional defined by
(23) |
Once again we sometimes simply write instead of . The LQ stochastic control problem is to minimize (23) over the set of admissible controls .
It holds that for every progressively measurable execution strategy there exists a control such that the cost functional can be rewritten in terms of (and ). In fact, this is achieved by taking , as outlined in the motivation above. We state this as Lemma˜2.1.
Lemma 2.1.
Let and . Suppose that with associated deviation . Define by , . It then holds that and that a.s.
On the other hand, we may also start with and derive a progressively measurable execution strategy such that the expected costs match.
Lemma 2.2.
Let and . Suppose that and let be the associated solution of (22). Define by , , , . It then holds that and that a.s.
Lemma˜2.1 and Lemma˜2.2 together with ˜1.7 establish the following equivalence of the control problems pertaining to , , and .
Corollary 2.3.
For and it holds that
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 and .
(i) Suppose that minimizes over and let be the associated deviation process. Then, defined by , , minimizes over .
(ii) Suppose that minimizes over and let be the associated solution of (22) for . Then, defined by , , , , minimizes over .
Moreover, if (in the sense that there is an element of within the equivalence class of , see (15)), then minimizes over .
2.2 Formulation without cross-terms
Note that the last integral in the definition (23) of the cost functional involves a product between the state process and the control process . 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 such that for all we have -a.s. that
(24) |
Note that this assumption ensures that the set is a subset of (up to a -null set). For this reason we, in the sequel, use the following
Convention: Under (24) we always understand on the set .
Now in order to get rid of the cross-term in (23) we transform for any control process in an affine way to , . This leads to the new controlled state process which is defined for every , , and by
(25) |
The meaning of (25) is that we only reparametrize the control () but not the state variable (), see Lemma˜2.5 for the formal statement. For , , and associated , we define the cost functional by
(26) |
This cost functional does not exhibit cross-terms, but is equivalent to of (23) in the sense of the following lemma.
Lemma 2.5.
Assume that (24) holds true. Let and .
(i) Suppose that with associated state process defined by (22). Then, defined by , , is in , and it holds that and .
(ii) Suppose that with associated state process defined by (25). Then, defined by , , is in , and it holds that and .
As a corollary, we obtain the following link between an optimal control for and an optimal control for .
Corollary 2.6.
Assume that (24) holds true. Let and .
(i) Suppose that is an optimal control for , and let be the solution of (22) for . Then, defined by , , is an optimal control in for .
(ii) Suppose that is an optimal control for , and let be the solution of (25) for . Then, defined by , , is an optimal control in for .
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 and , which are inhomogeneous. Therefore, the results of [40] are only directly applicable in the special case where and at least one of and 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 for the filtered probability space is the augmented natural filtration of the Brownian motion . Furthermore, we set the initial time to . We also assume that and are nonnegative -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 and 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 and is necessary to apply the results of [34]. Indeed, [34] requires that (the coefficient in front of in (26)) and (the coefficient in front of in (26)) are nonnegative and bounded, which implies that and have to be nonnegative.
Moreover, we note that nonnegativity of and ensures that (24) is satisfied. Further, we observe that the mentioned coefficients and are bounded, as required. Indeed, it clearly holds , and it remains to recall that , and are bounded and is -valued (see Section˜1.1).
Note that the LQ control problem of Section˜2.2, which consists of minimizing 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
(27) |
We call a pair with a solution to BSDE (27) if
-
(i)
is an adapted, continuous, nonnegative, and bounded process,
-
(ii)
-a.e.,
-
(iii)
, and
-
(iv)
BSDE (27) is satisfied -a.s.
A discussion of this definition is in order. The requirement of nonnegativity and boundedness of 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 appears in the denominator. Moreover, it is worth noting that, for a nonnegative , in our setting we always have , as . From this we also see that the quantity can vanish only in “very degenerate” situations. The conclusion is that condition (ii) is quite natural.
To shorten notation, we introduce, for a solution of BSDE (27), the process by, for ,
(28) |
Next, we consider the second BSDE [34, (85)], which is linear and reads in our setting
(29) |
A pair with is called a solution to BSDE (29) if
-
(i)
is an adapted continuous process with ,
-
(ii)
is progressively measurable with -a.s., and
-
(iii)
BSDE (29) is satisfied -a.s.
For a solution of BSDE (27) and a corresponding solution of BSDE (29), we define by
(30) |
for . We further introduce for and the SDE
(31) |
where for
We will show that the solution of (31) is the optimal state process in the stochastic control problem to minimize of (26). Notice that can be easily expressed via and in closed form.
In the next theorem, we summarize consequences from [34] in our setting to obtain a minimizer of in (26) and a representation of the minimal costs.
Theorem 3.3.
Assume that there exists such that -a.e. or -a.e. We then have:
(i) There exists a unique solution of BSDE (27). If -a.e., there exists such that .
(ii) There exists a unique solution of BSDE (29).
(iii) Let , and let be the solution of SDE (31). Then, defined by
(32) |
is the unique optimal control in for , and is the corresponding state process (i.e., ).
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 and in the cost functional (see (26)) are bounded, and that the inhomogeneities are in . Moreover, we have that , , and are nonnegative. Furthermore, the filtration by assumption in this section is generated by the Brownian motion .
(i) If , this is an immediate consequence of [34, Theorem 2.1]. In the case , 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 , which is given in feedback form by the formula . 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 such that -a.e. or -a.e. Let be the unique solution of BSDE (27), the unique solution of BSDE (29), and recall definitions (28) of and (30) of . Let . Then, defined by
with from (31), is the unique (up to -null sets) optimal execution strategy in for . The associated costs are given by
with from (33).
Remark 3.5.
(i) Note that BSDE (27) neither contains nor . In particular, the solution component and the process from (28) do not depend on the choice of or (although they depend on the choice of ). In contrast, BSDE (29) involves both and . If and at least one of and is equivalent to , we have that from (29), from (30), and from (33) vanish.
(ii) Under the assumptions of Corollary˜3.4 it holds that . This is a direct consequence of Corollary˜3.4 and (i) above. Indeed, choose and (by (i) this choice does not affect ). Then Corollary˜3.4 and (i) show that for the optimal strategy from Corollary˜3.4. The suboptimal finite variation execution strategy , , , in incurs costs and hence .
(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 , , and (and, therefore, the processes and 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 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 , which is in turn defined via a solution to a certain quadratic BSDE (see (3.2) in [1]). It is worth noting that, in the subsetting with , , and , 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 and the continuous local martingale is . to BSDE (3.2) in [1]. The relation is , , . Further, in that subsetting, our process from (28) reduces to the above-mentioned process (see (3.5) in [1]), while from (29), from (30), and from (33) vanish.
(iv) It is also instructive to compare Corollary˜3.4 above, where we obtain that the extended to 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 exhibit jumps (so-called block trades) at times and but are continuous on the interior (see also Section˜4.1 below). An interesting question is whether the continuity on 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 . And, indeed, we see that the continuity of on 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. ) and with terminal target . Indeed, consider the setting with , , and non-diffusive resilience process given by (with being a deterministic constant). Then it follows from [1, Example 6.2] that continuity of the price impact process is not sufficient for continuity of optimal position paths on . It is shown that if the paths of are absolutely continuous, then a jump of the weak derivative of on already causes to jump on . Moreover, it is possible that the random terminal target position causes the optimal position path to jump in 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 on consists of combining Corollary˜3.4 with path regularity results for BSDEs. Indeed, if the coefficient processes are continuous and if one can ensure that the solution components and (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 also has continuous sample paths on . 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 , , , , and choosing and 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 and general running targets .
To this end let . Suppose that , , , and . Furthermore, assume that and are deterministic constants. We take some and 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 for all . In the current setting, BSDE (27) reads
(34) |
By ˜3.3, there exists a unique solution . Since the driver and the terminal condition in (34) are deterministic, we obtain that , 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 that
and in the case that
(35) |
The process from (28) here is given by , . BSDE (29) becomes
(36) |
Again, by ˜3.3, there exists a unique solution . The solution component is given by
where
(37) |
It holds for the process in (30) that , . Further, SDE (31) reads
and has solution
with from (37). It then follows from Corollary˜3.4 that defined by , , and, for ,
(38) |
is the (up to -null sets unique) execution strategy in that minimizes .
Remark 4.2.
In the next example we study the case in more detail.
Example 4.3.
In the setting of the previous example suppose that . If the terminal target is a deterministic constant, then it follows from [37, Proposition 3] that the optimal finite variation execution strategy is given by
(39) |
So the optimal strategy consists of potential block trades (jumps of ) at times and and a continuous linear trading program on . In the following we analyze how this structure changes as we allow for a random terminal target .
First recall that the solution of BSDE (34) is given in this case by (35). It follows that from (37) simplifies to , . For the solution component of BSDE (36), we thus obtain
The optimal strategy from (38) on becomes, for ,
(40) |
Integration by parts implies that (note that is a continuous martingale)
Substituting this into (40) yields, for ,
We, finally, obtain the alternative representation
for (40). We see that this optimal strategy 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 . The second part represents fluctuations around this deterministic strategy which incorporate updates about the random terminal target . Note that this stochastic integral vanishes in expectation, although this is not a martingale (indeed, the time 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 even if all input processes, including , are continuous. To this end, let . Take , , , , and , and assume that and are deterministic constants. Moreover, we will later consider an appropriate random terminal target , 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 of BSDE (27), and it is given by (compare also with [1, Section 5.2]) and
where denotes the Lambert function and . The process from (28) becomes
and both and are deterministic, increasing, continuous, -valued functions.
For some , let
where , . Note that is -measurable and that . The terminal target here is defined in such a way that the unique solution of BSDE (29) (cf. ˜3.3) is given by , , , and
It follows for the process in (30) that
We thus have that
From Corollary˜3.4 we obtain existence of a unique optimal strategy and that , . Since , , and (see (31)) are continuous and , it holds that . Hence, the optimal strategy has a jump at .
4.3 An example where does not admit a minimizer
Let with . Suppose that , , , , , . Choose to be a bounded deterministic càdlàg function such that there exists with having infinite variation on , and take such that there exists with . 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
Its solution is given by , where (see also in [1, Section 6])
is a deterministic continuous function of finite variation. We have that
which is the same as in [1, Example 6.4]. The solution of BSDE (29) is given by , and it holds . Furthermore, (31) reads
and is solved by the continuous deterministic finite-variation function
which is nonvanishing due to our assumption .121212At this point it is easy to explain why we exclude the case in this example. In the case we get that and then the optimal strategy is to close the position immediately, i.e., , , , which is always a finite-variation strategy. By Corollary˜3.4, there exists a (up to -null sets) unique minimizer of in , namely
Assume by contradiction that there exists a minimizer of in . We know from Corollary˜2.3 that is then also a minimizer of in . It follows that -a.e. Since is nowhere , we obtain that
(41) |
Observe that the left-hand side is a process of finite variation. On the other hand, our assumption on easily yields that has infinite variation. This contradiction proves that in the setting of this example, does not admit a minimizer in .
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 . Indeed, if we had a semimartingale as a minimizer, we would still get (41) (with a semimartingale ). 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 such that -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 is an increasing process. In [1] and [3] is allowed to have finite variation. Now we consider an example with a truly diffusive .
Let with . Let , , , and . Suppose that and are deterministic constants such that and (in particular, we thus need ). Note that the assumptions of Corollary˜3.4 are satisfied. We moreover remark that the subsetting where 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
This has solution with
and (compare also with [1, Section 5.2]). We further have that , . Observe that is the solution of (29) in the present setting and that in (30). Moreover, we have that SDE (31) reads
for , with start in ; hence,
It follows from Corollary˜3.4 that for the optimal execution strategy is given by
We can show that and both are continuous, deterministic, increasing, -valued functions of finite variation. Since , the optimal strategy on always has the same sign as . 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 for a choice of . In this case, the price impact is a deterministic constant, yet the optimal strategy has infinite variation (due to the infinite variation in the resilience ).
Observe furthermore that by making use of and , we can choose the parameters in the current setting in such a way that and are satisfied, but condition (3.1) in [1], i.e., , is violated.
With regard to Section˜4.3 we remark that in both sections there does not exist an optimal strategy in , 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 is “cancelled” by the infinite variation in the resilience process and we obtain the optimal strategy of finite variation.
Let , , , , and . Suppose that and are deterministic constants, and that and are progressively measurable, -a.e. bounded processes such that -a.e. It then holds -a.e. that and . In particular, the assumptions of Corollary˜3.4 are satisfied. The BSDE
which is BSDE (27) in the present setting, has the solution with , (cf. Section˜4.1). It holds that , that is the solution of (29), and that . It follows that (31) has the solution
For the optimal execution strategy from Corollary˜3.4 we then compute that
The optimal strategy in the current setting with general stochastic and negative correlation is thus the same as in the Obizhaeva-Wang setting (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 and the resilience are both stochastic and of infinite variation (at least if is nonvanishing).
We finally remark that this setting does not reduce to the Obizhaeva-Wang setting . Indeed, while the optimal strategies for and for general stochastic with correlation coincide, this is not true for the associated deviation processes. In general, it holds that
which for a nonvanishing and has infinite variation, whereas in the Obizhaeva-Wang setting is constant (take ).
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
(42) |
We moreover observe that by Itô’s lemma it holds that
(43) |
(44) |
Proof of ˜1.3.
Observe that integration by parts implies that for all
Since , , it follows that the process , , , satisfies
(45) |
In particular, is of finite variation. The facts that , , and , , imply that
(46) |
where we denote , , and, in the last equality, we use that , as has finite variation. Summing up, (46) yields
Proof of ˜1.4.
We first consider the integrator on the right hand side of (10). It holds by integration by parts and (9) that for all
Note that for all we have
It hence follows for all that
Plugged into (10) from ˜1.3, we obtain that
(47) |
We further have by (3) and (42) that for all
(48) |
It follows from assumption (A1) and the boundedness of the input processes that
The Burkholder-Davis-Gundy inequality together with assumption (A3) shows that it holds for some constant that
We therefore have that Similarly, assumption (A2) implies that It thus follows from (47), (48), and (12) that
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 , . Assume that is a progressively measurable process such that a.s. For , , and , , it then holds for all that
(49) |
Proof.
Integration by parts implies that
(50) |
Furthermore, it holds by integration by parts, (8), (3) and (4) that for all
(51) |
Also by integration by parts, and using (9), (3) and (44), we obtain for all that
(52) |
It follows from (51) and (52) for all that
(53) |
We then plug (51), (52) and (53) into (50), which yields (49). ∎
Proof of Lemma˜1.6.
We denote , , and , . It then holds that , . We use Lemma˜5.1 and substitute in (49) to obtain for all that
This proves the dynamics in (18).
In particular, satisfies an SDE that is linear in and . Furthermore, boundedness of implies that the coefficients of the SDE are bounded. Since moreover by assumption (A1) and (cf. (17)) is square integrable, we have that (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 that
Due to assumption (5) on and , we have that . This, assumption (A1), and the Cauchy–Schwarz inequality imply that also . Since is bounded, we conclude that
(54) |
where all conditional expectations are well-defined and finite. Moreover, (17) implies that , and thus . Inserting this and (54) into (16), we obtain (19). ∎
Lemma 5.2.
Let and . Then, (20) defines a metric on (identifying any processes that are equal -a.e.).
Proof.
Note first that it holds for all that , and that is finite due to (A1). Symmetry of is obvious. The triangle inequality follows from the Cauchy–Schwarz inequality.
Let with associated deviation processes .
If -a.e., then -a.e., and thus .
For the other direction, suppose that . This implies that -a.e. By definition of and it further follows from a multiplication by that Observe that and consider the stochastic integral equation
(55) |
Define by ,
It then follows from (51) that (55) can be written as , . This has the unique solution . We therefore conclude that -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 , , and with associated deviation and scaled hidden deviation . Suppose in addition that is a sequence in such that . for the associated deviation processes , . It then holds for the associated scaled hidden deviation processes , , that
Proof.
Define , , and let for , ,
In view of (18) it then holds for all that
Linearity of , , , and boundedness of imply that there exists such that for all and all it holds -a.e. that
By boundedness of and Jensen’s inequality, we have some such that for all ,
E.g., [41, Theorem 3.2.2] (see also [41, Theorem 3.4.2]) now implies that there exists such that for all
The claim follows from the assumption that . ∎
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 defined by
(56) |
Observe that by Itô’s lemma, solves the SDE
(57) |
Lemma 5.4.
Let and let . Then there exists a sequence of bounded càdlàg finite variation processes such that
In particular, for the sequence of processes defined by , , it holds for all that is a càdlàg semimartingale and for any (in particular, ), and that
Proof.
Define by , . Moreover, let be defined by , . We verify the assumptions of Lemma 2.7 in Section 3.2 of [33]. The process is continuous, adapted and nondecreasing. Note that boundedness of , and implies that the coefficients of (57) are bounded. It follows for any that (see, e.g., [41, Theorem 3.4.3]), and hence . Since , we have that is progressively measurable and satisfies . Thus, Lemma 2.7 in Section 3.2 of [33] applies and yields that there exists a sequence of (càglàd) simple processes , , such that . Define , , , , and , . Then, is a sequence of bounded càdlàg finite variation processes such that . Note that for each , defined by , , is càdlàg. Since is bounded for all and for any , we have that is finite for all and any . It furthermore holds that as . ∎
For the part in ˜1.7 on completeness of we show how to construct an execution strategy based on a square integrable process and a process that satisfies SDE (18) (with instead of ). This result is also crucial for Lemma˜2.2.
Lemma 5.5.
Let and . Suppose that , and let be given by ,
(58) |
Define by , , , . Then, , and for the associated deviation process it holds .
Proof.
First, is progressively measurable and has initial value and terminal value . Furthermore, it holds that
since and have a.s. continuous paths and . We are therefore able to define by (14). Moreover, denote , , and , . It follows from Lemma˜5.1 and , , that for all
We combine this with
to obtain for all that
(59) |
Note that . We thus conclude that is the unique solution of (59), and hence , . This implies that , i.e., , , and . The fact that then immediately yields that (A1) holds. This proves that . ∎
We finally are able to prove ˜1.7.
Proof of ˜1.7.
(i) Denote by , , , the deviation processes associated to , , , and let and , , be the scaled hidden deviation processes. By Lemma˜1.6 it holds for all that
Boundedness of and implies (recall also (12)) that there exists some such that for all it holds that
(60) |
We treat the terminal costs first. It holds for all that
From
(61) |
(cf. (20)) and Lemma˜5.3 we have that
(62) |
Since furthermore , we obtain that The second term in (60) converges to using (61). For the third term in (60) we have for all that
(63) |
By Lemma˜1.6 and (5) it holds that . Moreover, due to (61), we have that is uniformly bounded in . It thus follows from (61), (62) and (63) that the third term in (60) converges to as . The last term in (60) converges to using (5) and (62). This proves claim (i).
(ii) Suppose that . Let be defined by , , where denotes the deviation associated to . Then, is a progressively measurable process, and due to assumption (A1) it holds that . By Lemma˜5.4 there exists a sequence of bounded càdlàg finite variation processes such that , where is defined in (56). Set , . This is a sequence of càdlàg semimartingales in that satisfies . Moreover, it holds for all and any that . For each , , let be the solution of (58). We then define a sequence of càdlàg semimartingales , , by , , , . By Lemma˜5.5 we have for all that and that for the associated deviation process . It follows for all that , . Therefore, it holds for all that
Due to , we thus have that . We next show that for all , has finite variation. To this end, we observe that for all and it holds by integration by parts that
(64) |
Again by integration by parts, and using (57), we have for all and that
This and (58) yield for all and that
(65) |
Moreover, it follows from (44) for all and that
(66) |
We combine (64), (65), and (66) to obtain for all and that
Since has finite variation for all , this representation shows that also has finite variation for all . Note that for all , by ˜1.3, the process (6) associated to the càdlàg finite variation process is nothing but . Since is bounded, there exists such that for all
This implies (A2). Similarly, by boundedness of , we obtain (A3). We thus conclude that for all .
(iii) Let be a Cauchy sequence in . For we denote by the deviation process associated to . It then holds that is a Cauchy sequence in . Since is complete (see, e.g., Lemma 2.2 in Section 3.2 of [33]), there exists such that . Define by , , , , where is given by (58). By Lemma˜5.5 it holds that . We furthermore obtain from Lemma˜5.5 that, for the associated deviation, . By definition of , this yields , . It follows that
and hence . ∎
Proof of Lemma˜2.1.
By definition of we have that is progressively measurable and, due to assumption (A1), satisfies ; hence, .
Proof of Lemma˜2.2.
It follows from Lemma˜5.5 that . Moreover, we have from Lemma˜5.5 that the associated deviation satisfies , i.e., , , and is the scaled hidden deviation of . It thus holds that is given by (19). In the definition (23) of , we may replace under the integrals with respect to the Lebesgue measure by . This shows that . ∎
Proof of Lemma˜2.5.
(i) We have that is progressively measurable. Furthermore, the facts that , , , and (24) imply that . Hence, . Substituting , , in (22) leads to (25). For the cost functional, observe that
(67) |
(ii) Note that (25) is an SDE that is linear in , , and . Furthermore, boundedness of and (24) imply that the coefficients of the SDE are bounded. Since moreover and is square integrable, we know that (see, e.g., [41, Theorem 3.2.2 and Theorem 3.3.1]). We can thus argue similar to (i) that . A substitution of 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 / 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.