1 Introduction

Let \(T \in (0,\infty )\) be a given deterministic time horizon and \(d \in \mathbb {N}\), let \(\Omega := C([0,T],\mathbb {R}^d)\) be the canonical space of continuous paths. We denote by B the canonical process and by \(\mathbb {P}\) the Wiener measure. Equip \(\Omega \) with \((\mathcal {F}_t)_{t\in [0,T]}\), the \(\mathbb {P}\)-completion of the canonical filtration of B. Given a d-dimensional vector \(\sigma \) and a function \(b: [0,T]\times \mathbb {R}\times \mathbb {R}^m\rightarrow \mathbb {R}\), we consider a controlled diffusion of the form

$$\begin{aligned} \,\textrm{d}X^\alpha (t) = b(t,X^\alpha (t),\alpha (t))\,\textrm{d}t +\sigma \,\textrm{d}B(t) ,\quad t \in [ 0,T] ,\quad X^\alpha (0) = x_0 \end{aligned}$$
(1)

and the control problem

$$\begin{aligned} V(x_0) := \sup _{\alpha \in \mathcal {A}}J(\alpha ). \end{aligned}$$
(2)

Hereby, the performance functional J is given by

$$\begin{aligned} J(\alpha ):= \mathbb {E}\Big [ \int _0^T f(s,X^\alpha (s),\alpha (s))\,\textrm{d}s + g(X^\alpha (T))\Big ], \end{aligned}$$

where, f and g may be seen as profit and bequest functions, respectively. The set \(\mathcal {A}\) is the set of admissible controls and is defined as the set of progressively measurable processes \(\alpha \) valued in a closed convex set \(\mathbb {A}\subseteq \mathbb {R}^m\) such that (1) admits a unique strong solution. The goal of the present article is to derive the maximum principle for the above control problem when the drift b is merely measurable in the state variable x.

The stochastic maximum principle is arguably one of the most prominent ways to tackle stochastic control problems as (2) by fully probabilistic methods. It is the direct generalization to the stochastic framework of the maximum principle of Pontryagin [47] in deterministic control. It gives a necessary condition of optimality in the form of a two-point boundary value problem and a maximum condition on the Hamiltonian. More precisely let the Hamiltonian H be defined as

$$\begin{aligned} H(t, x, y, a): = f(t,x,a) + b(t,x,a)y \end{aligned}$$

and assume just for a moment the functions bf and g to be continuously differentiable. Then, if \({\hat{\alpha }} \in \mathcal {A}\) is an optimal control, then according to the stochastic maximum principle, it holds \(H(t, X^{{\hat{\alpha }}}(t), Y(t), {\hat{\alpha }}(t)) \ge H(t, X^{{\hat{\alpha }}}(t), Y(t), a) \) \(P\otimes dt\)-a.s. for every \(a \in \mathbb {A}\) where (YZ) are adapted processes solving the so-called adjoint equation

$$\begin{aligned} \,\textrm{d}Y(t)= & {} - \partial _xf(t, X^{{\hat{\alpha }}}(t),{\hat{\alpha }}(t)) - \partial _xb(t, X^{{\hat{\alpha }}}(t),{\hat{\alpha }}(t))Y(t)\,\textrm{d}t + Z(t)\,\textrm{d}B(t),\\ Y(T)= & {} \partial _xg(X^{{\hat{\alpha }}}(T)). \end{aligned}$$

Under additional convexity conditions, this necessary condition is sufficient. The interest of the maximum principle is that it reduces the solvability of the control problem (2) to that of a (scalar) variational problem, and therefore allows to derive (sometimes explicit) characterizations of optimal controls. We refer for instance to [10, 52] for proofs and historical remarks. The maximum principle has far-reaching consequences and is widely used in the stochastic control and stochastic differential game literature [11, 12, 24, 30, 34, 45]. Its use also fueled by recent progress on the theory of forward backward SDEs. We refer the reader for instance to, [16, 32, 33, 35, 46, 53] and the references therein.

The maximum principle roughly presented above naturally requires differentiability of the coefficients of the control problem, which precludes the applicability of this method to control problems with non-smooth coefficients. The effort to extend the stochastic maximum principle to problems with non-smooth coefficients started with the work of Merzedi [41] who derived a necessary condition of optimality for a problem with a Lipschitz continuous drift, but not necessarily differentiable everywhere in the state and the control variable. His result was further extended, notably to degenerate diffusion cases and singular control problems in [1,2,3,4, 9, 31, 43]. See also [50] for the infinity horizon case. To the best of our knowledge, all existing results on the stochastic maximum principle assume some level of regularity, usually Lipschitz-continuous drifts.

The present work considers the case where b is Borel measurable in x and bounded, and we will derive both necessary and the sufficient conditions of optimality. At this point, an immediate natural question is: What form should the adjoint equation take in this case? The starting point of our argument is the following simple observation: When b is differentiable, the adjoint equation is explicitly solvable, with the solution given by

$$\begin{aligned} Y(t) = \mathbb {E}\Big [\Phi ^{{\hat{\alpha }}}(t,T)\partial _xg(X^{{\hat{\alpha }}}(T)) + \int _t^T\Phi ^{{\hat{\alpha }}}(t,s)\partial _xf(s, X^{{\hat{\alpha }}}(s),{\hat{\alpha }}(s))\,\textrm{d}s\mid \mathcal {F}_t \Big ], \end{aligned}$$

where the process

$$\begin{aligned} \Phi ^{{\hat{\alpha }}}(t,s) = e^{\int _t^s\partial _xb(u, X^{{\hat{\alpha }}}(u),{\hat{\alpha }}(u))\,\textrm{d}u}\quad 0\le t\le s\le T \end{aligned}$$
(3)

is the first variation process (in the Sobolev sense) of the dynamical system \(X^{{\hat{\alpha }},x}\) solving (1) with initial condition \(X^{{\hat{\alpha }},x}_0 = x\). This suggests the form of the adjoint process when b is not differentiable, since it is well-known that despite the roughness of the drift b, the dynamical system \(X^{{\hat{\alpha }},x}\) is still differentiable with respect to x (at least in the Sobolev sense), due to Brownian regularization [42] and therefore admits a Sobolev differentiable flow. The crux of our argument will be to make use of this Sobolev differential stochastic flow to define the adjoint process (rather than the adjoint equation) in the non-smooth case to prove necessary and sufficient conditions of optimality.

Throughout this work the functions f and g are assumed to be continuously differentiable with first derivatives of linear growth. In particular, we will assume

$$\begin{aligned} \sigma \in \mathbb {R}^d \text { satisfies } |\sigma |^2> & {} 0 \quad \text {and}\quad |f(t, x, a)| + |g(x)| \\\le & {} C(1 + |x|^2)\quad \hbox {for all}\, (t,x,a)\, \hbox {and some} C>0 \end{aligned}$$

and

$$\begin{aligned} |\partial _xf(t,x, a)| + |\partial _xg(x)| \le C(1 + |x|). \end{aligned}$$

The main results of this work are the following necessary and sufficient conditions in the Pontryagin stochastic maximum principle.

Theorem 1.1

Assume that b satisfies \(b(t,x,a):= b_1(t,x) + b_2(t,x,a)\) where \(b_1\) is a bounded, Borel measurable function and \(b_2\) is bounded measurable, and continuously differentiable in its second and third variables with bounded derivatives. Let \({\hat{\alpha }} \in \mathcal {A}\) be an optimal control and let \(X^{{\hat{\alpha }}}\) be the associated optimal trajectory. Then the flow \(\Phi ^{{\hat{\alpha }}}\) of \(X^{{\hat{\alpha }}}\) is well-defined and it holds

$$\begin{aligned} \partial _{a}H(t, X^{{\hat{\alpha }}}(t),Y^{{\hat{\alpha }}}(t),{\hat{\alpha }}(t) )\cdot (\beta - {\hat{\alpha }}(t)) \ge 0 \quad \mathbb {P}\otimes \,\textrm{d}t\text {-a.s. for all } \beta \in \mathcal {A}, \end{aligned}$$
(4)

where \(Y^{{\hat{\alpha }}}\) is the adjoint process given by

$$\begin{aligned} Y^{{\hat{\alpha }}}(t):= \mathbb {E}\Big [\Phi ^{{\hat{\alpha }}}(t, T) \partial _xg( X^{{\hat{\alpha }}}(T)) + \int _t^T\Phi ^{{\hat{\alpha }}}(t,s) \partial _xf(s, X^{{\hat{\alpha }}}(s), {\hat{\alpha }}(s))\textrm{d}s\mid \mathcal {F}_t \Big ]. \end{aligned}$$
(5)

Theorem 1.2

Let the conditions of Theorem 1.1 be satisfied, further assume that g and \((x,a)\mapsto H(t,x,y,a)\) are concave. Let \({\hat{\alpha \in }} \mathbb {A}\) satisfy

$$\begin{aligned} \partial _{a} H(t, X^{{\hat{\alpha }}}(t), Y^{{\hat{\alpha }}}(t), {\hat{\alpha }}(t))=0 \quad \mathbb {P}\otimes \,\textrm{d}t\text {-a.s.} \end{aligned}$$
(6)

with Y given by (5). Then, \({\hat{\alpha }}\) is an optimal control.

Theorems 1.1 and 1.2 constitute sharp improvements over existing results as far as regularity of the drift is concerned, since it assumes measurable drifts as opposed to Lipschitz-continuous in the literature. We will elaborate on the conditions imposed in the above theorems in Sect. 4. Let us at this point remark that the result remains true when assuming \(b_2\) Lipschitz-continuous (see Remark (2.2). We do not do so to ease the presentation and focus on the non-smoothness issue. However, the techniques of the proof presented here do not seem to extend to the random volatility case because the various applications of Girsanov’s theorem might fail in this case. It is conceivable that a technique based on Zvonkin’s transform allows to derive a maximum principle in the non-constant volatility case. However using this method, the roughness of the drift coefficient does not enable to show the differentiability with respect to the initial condition which is key in the proof of the maximum principle. In addition, observe that when b is smooth, our results correspond exactly to the classical version of the stochastic maximum principle. The only difference here being the fact that the process \(\Phi ^{{\hat{\alpha }}}\) seems abstract, as it is obtained from an existence result (of the first variation process). It turns out that when the drift is not smooth, the flow \(\Phi ^{{\hat{\alpha }}}\) still admits an explicit representation much similar to (3), but in terms of the local time of the controlled process. This representation will be investigated in the present case of controlled diffusions with rough drifts in the “Appendix” (see Theorem A.1) and will be used in the proof of the maximum principle. Further note that the explicit representation of the flow is a result of independent interest. We will not discuss it further to avoid losing focus from the paper’s subject.

1.1 Motivation: Optimal Consumption Under Wealth Tax Payment

This problem is motivated by the optimal consumption problem of a financial agent paying wealth tax. In fact, since the seminal work of Merton [39], this problem has attracted much attention and enormous progress has been made. Optimal investment and consumption in complete and incomplete markets is well understood and several methods have been developed. Let us refer for instance to [21, 25,26,27] for just a few milestones. The impetus for the present work is to extend this literature (especially the optimal consumption problem of Cuoco and Cvitanić [13, 14]) to the very practical situation of optimal consumption of an individual (e.g. a retiree) living off of their investment in the stock market while paying wealth taxes.

We consider the classical financial model of Cuoco [13] with m stocks \(S = (S^1,\dots ,S^m)\) with cumulative dividend processes \(D = (D^1,\dots , D^m)\) such that \(S+D\) is the Itô diffusion

$$\begin{aligned} S(t) + D(t) = S(0) + \int _0^tS(u)\mu \,\textrm{d}u + \int _0^tS(u)\sigma \,\textrm{d}B(u). \end{aligned}$$

In addition there is a bond with rate \(r=0\). The agent is endowed with an initial wealth \(x_0>0\) and a nonnegative stochastic income process y such that

$$\begin{aligned} \int _0^Ty(u )\,\textrm{d}u \le K_y \end{aligned}$$

for some \(K_y>0\). Further assume that the agent also consumes at rate c(t). If \(\theta \) represents the dollar amount invested at time t, the wealth of the agent evolves as

$$\begin{aligned} {\widetilde{X}}(t) = x(0) + \int _0^t\theta (u)\mu \,\textrm{d}u + \int _0^t\theta \sigma \,\textrm{d}B(u) - \int _0^tc(u) - y(u)\,\textrm{d}u. \end{aligned}$$

As in [13, 14], we assume that the agent fixes an investment strategy \(\theta \) (which we assume for simplicity to be constant) and looks for the optimal consumption plan \({\hat{c}}\). This problem is fully solved in [13] when general constraints are put on the admissibility of c.

In the present work, we will further assume that the agent pays wealth taxes. In most countries and states, tax categories are set with respect to the tax payer’s wealth. To simplify the exposition, we will assume that only two tax categories are given; low and high: The agent pays \(\ell \) if their wealth is below a given threshold e and h otherwise. Hence, the wealth process now takes the form

$$\begin{aligned} X(t)= & {} x(0) + \int _0^t\theta \mu \,\textrm{d}u + \int _0^t\theta \sigma \,\textrm{d}B(u) - \int _0^tc(u) - y(u)\,\textrm{d}u \\{} & {} - \int _0^t\ell 1_{\{X(s)\le e\}} + h1_{\{X(s)> e\}} \,\textrm{d}s. \end{aligned}$$

That is, the agent’s wealth is the sum of their initial endowment and their trading gains minus cumulative withdrawals and cumulative tax paid. The problem faced by the agent is thus

$$\begin{aligned} \begin{array}{l} \sup _{c \in \mathcal {A}}\mathbb {E}\Big [\int _0^TU(t,c(t))\,\textrm{d}t \Big ]\\ \,\textrm{d}X(t) = b_1(t,X(t))+ b_2(t, X(t), c(t))\,\textrm{d}t + \tilde{\sigma }\,\textrm{d}B(t),\quad X(0) = x(0) \end{array} \end{aligned}$$

with \(b_1(t,x) = \ell 1_{\{x\le e\}}+ h1_{\{x> e\}}\), \(b_2(t,x,c):= y_t + \theta \mu _t - c \), \({\tilde{\sigma }}: \theta \sigma \) and a utility function \(U:[0,T]\times \mathbb {R}\) that is assumed to be increasing, strictly concave, continuously differentiable in the second argument, and continuous in the first.

The problem here is the fact that, due to tax payment, the drift of the state process X is discontinuous. This stochastic control problem falls within the scope of our maximum principle discussed above, which then allows to provide a fully probabilistic characterization of the solution of this problem via the associated adjoint process.

To the best of our knowledge, optimal consumption problems under with wealth tax payment have not been considered so far. Note however the works of [6, 7] on optimal control under capital gain taxes using dynamic programming.

The remainder of the article is dedicated to the proofs of Theorems 1.1 and 1.2. The necessary condition is proved in the next section and the sufficient condition is proved in Sect. 3. In this section, we also present and example where, in addition to providing a characterization, our maximum principle allows to derive explicit solution to a control problem with non-smooth coefficients. In Sect. 4 we discuss the conditions imposed in the main theorems. The paper ends with an “Appendix” on explicit representations of the flow of SDEs with measurable and random drifts.

2 The Necessary Condition for Optimality

The goal of this section is to prove Theorem 1.1. Let us first precise the definition of the set of admissible controls. Let \(\mathbb {A}\subseteq \mathbb {R}^m\) be a closed convex subset of \(\mathbb {R}^m\). The set of admissible controls is defined as:

$$\begin{aligned} \mathcal {A}:= & {} \Big \{\alpha :[0,T]\times \Omega \rightarrow \mathbb {A}, \text { progressive}, (1)\text { has a unique strong solution and } \\{} & {} \mathbb {E}\big [\sup _{t\in [0,T]}|\alpha (t)|^{4}\big ]< M \Big \} \end{aligned}$$

for some \(M>0\). This set is clearly non-empty even when \(b_1\) is not trivial. In fact, \(\mathcal {A}\) already includes a large class of controls usually considered in the literature. Let us illustrate this with two examples:

Example 2.1

  • Markovian controls If one considers controls of the form \(\alpha _t =\varphi (t,X_t)\) for a measurable function \(\varphi \), then the SDE (1) admits a unique strong solution, see e.g. [42] or [22].

  • Open loop controls Consider the set \(\mathcal {A}'\) defined as: The set of progressively measurable processes \(\alpha :[0,T]\times \Omega \rightarrow \mathbb {A}\) which are Malliavin differentiable (with Malliavin derivative \(D_s\alpha (t)\)), with

    $$\begin{aligned} \mathbb {E}\Big [\int _0^T|\alpha (t)|^2\,\textrm{d}t \Big ] + \sup _{s\in [0,T]}\mathbb {E}\Big [\Big (\int _0^T|D_s\alpha (t)|^2\,\textrm{d}t\Big )^4 \Big ] <\infty \end{aligned}$$

    and such that there are constants \(C,\eta >0\) (possibly depending on \(\alpha \)) such that

    $$\begin{aligned} \mathbb {E}[|D_s \alpha (t) - D_{s'}\alpha (t)|^4] \le C|s-s'|^\eta . \end{aligned}$$

    It follows from [38, Theorem 1.2] that if the drift satisfies the conditions of Theorem 1.1, then the SDE (1) is uniquely solvable for every \(\alpha \in \mathcal {A}'\).

For later reference, note that for every \(\alpha \in \mathcal {A}\) it holds \(E[\sup _{t\in [0,T]}|X^\alpha (t)|^p]<\infty \) for every \(p\ge 1\).

In the rest of the article, we let \(b_n\) be a sequence of functions defined by

$$\begin{aligned} b_n:= b_{1,n} + b_{2} \end{aligned}$$
(7)

such that \(b_{1,n}: [0, T ] \times \mathbb {R} \rightarrow \mathbb {R}, n \ge 1\) are smooth functions with compact support and converging a.e. to \(b_1\). Since \(b_1\) is bounded, the sequence \(b_{1,n}\) can also be taken bounded. We denote by \(X^{\alpha }_n\) the solution of the SDE (1) with drift b replaced by \(b_n\). This process is clearly well-defined since \(b_n\) is a Lipschitz continuous function. Similarly, we denote respectively by \(J_n\) and \(V_n\) the performance and the value function of the problem when the drift b is replaced by \(b_n\). That is, we put

$$\begin{aligned} J_n(\alpha ):= \mathbb {E}\Big [\int _0^Tf(s, X^\alpha _n(s), \alpha (s))\,\textrm{d}s + g(X_n^\alpha (T)) \Big ],\quad V_n(x_0):= \sup _{\alpha \in \mathcal {A}}J_n(\alpha ) \end{aligned}$$

and

$$\begin{aligned} \,\textrm{d}X_n^\alpha (t) = b_n(t,X_n^\alpha (t),\alpha (t))\,\textrm{d}t +\sigma \,\textrm{d}B(t),\quad t \in [ 0,T],\quad X^\alpha (0) = x_0. \end{aligned}$$

Furthermore, we denote by \(\delta \) the distance

$$\begin{aligned} \delta (\alpha _1, \alpha _2): = \mathbb {E}\big [\sup _{t \in [0,T]}|\alpha _1(t) - \alpha _2(t)|^{4} \big ]^{1/4}. \end{aligned}$$

The general idea of the proof will be to start by showing that an optimal control for the problem (2) is also optimal for an appropriate perturbation of the approximating problem with value \(V_n(x_0)\). This is due to the celebrated variational principle of Ekeland. This maximum principle for control problems with smooth drifts will involve the state process \(X_n^{{\hat{\alpha _n}}}\) and its flow \(\Phi ^{{\hat{\alpha _n}}}_n\). The last and most demanding step is to pass to the limit and show some form of “stability” of the maximum principle.

Remark 2.2

When \(b_2\) is not continuously differentiably but Lipschitz-continuous, one also approximates it by smooth functions \(b_{2,n}\) which have uniformly bounded derivatives, i.e. such that \(\sup _n|\partial _xb_{2,n}|<\infty \). The rest of the proof is then the same.

We first address this limit step by a few intermediary technical lemmas that will be brought together to prove Theorem 1.2 at the end of this section.

Lemma 2.3

We have the following bounds:

  1. (i)

    For every sequence \((\alpha _n)_n\) in \(\mathcal {A}\), it holds \(\sup _n\mathbb {E}\big [\sup _{t \in [0,T]}|X^{\alpha _n}_n(t)|^2 \big ]<\infty \).

  2. (ii)

    For every \(\alpha _1,\alpha _2 \in \mathcal {A}\) it holds that

    $$\begin{aligned} \mathbb {E}\big [| X^{\alpha _1}_n(t) - X^{\alpha _2}(t)|^2 \big ]\le & {} C\Big ( \delta (\alpha _1,\alpha _2)^4 + \Big (\int _0^T\frac{1}{\sqrt{2\pi s}}e^{\frac{|x_0|^2}{2s}}\int _{\mathbb {R}^d}\\{} & {} \big |b_{1,n} (s,\sigma y)-b_{1} (s,\sigma y)\big |^4e^{-\frac{|y|^2}{4s}}\textrm{d}y\,\textrm{d}s\Big )^{1/2}\Big ). \end{aligned}$$
  3. (iii)

    Given \(k \in \mathbb {N}\), for every sequence \((\alpha _n)_{n\ge 1}\) in \(\mathcal {A}\) and \(\alpha \in \mathcal {A}\) such that \(\delta (\alpha _n,\alpha )\rightarrow 0\), it holds that

    $$\begin{aligned} \mathbb {E}\big [| X^{\alpha _n}_k(t) - X^{\alpha }_k(t)|^2 \big ] \rightarrow 0. \end{aligned}$$

Proof

The proof of (i) is standard, it follows from the linear growth property of \(b_n\) uniformly in n, i.e. \(|b_n(t,x,a)| \le C(1 + |x| + |a|)\) for some \(C>0\) and all \(n\ge 1\).

Let us turn to the proof of (ii). Adding and subtracting the same term and using the fundamental theorem of calculus, we arrive at

$$\begin{aligned} X_n^{\alpha _1}(t) {-} X^{\alpha _2}(t)&{=} \int _0^t\int _0^1\partial _xb_{1,n}(s, \Lambda _n(\lambda ,s)) {+} \partial _xb_2\big (s,\Lambda _n(\lambda ,s),\alpha _1(s)\big )\textrm{d}\lambda (X^{\alpha _1}_n(s) \\&\quad - X^{\alpha _2}(s))\textrm{d}s\\&\quad + \int _0^t b_{1,n}(s, X^{\alpha _2}(s)) - b_1(s, X^{\alpha _2}(s))\textrm{d}s \\&\quad + \int _0^tb_2(s, X^{\alpha _2}(s),\alpha _1(s)) - b_2(s, X^{\alpha _2}(s), \alpha _2(s))\,\textrm{d}s, \end{aligned}$$

where \(\Lambda _n(\lambda ,t)\) is the process given by \(\Lambda _n(\lambda ,t):= \lambda X^{\alpha _1}_n(t) + (1 - \lambda )X^{\alpha _2}(t)\). Therefore, we obtain that \(X^{\alpha _1}_n - X^{\alpha _2}\) admits the representation

$$\begin{aligned} X^{\alpha _1}_n(t) {-} X^{\alpha _2}(t)&{=} \!\!\int _0^t\exp \Big (\!\int _{s}^t\!\int _0^1\partial _xb_{1,n}(r, \Lambda _n(\lambda ,r)) {+} \partial _xb_2(r,\Lambda _n(\lambda ,r), \alpha _1(r))\textrm{d}\lambda \textrm{d}r \Big )\\&{\times } \Big (b_{1,n}(s, X^{\alpha _2}(s))\\&{-} b_1(s, X^{\alpha _2}(s)) {+}b_2(s, X^{\alpha _2}(s),\alpha _1(s)) {-} b_2(s, X^{\alpha _2}(s),{\alpha _2}(s))\Big )\textrm{d}s. \end{aligned}$$

Hence, taking the expectation on both sides above and then using twice the Cauchy–Schwarz inequality, we have that

$$\begin{aligned} \mathbb {E}\big [|X^{\alpha _1}_n(t) - X^{\alpha _2}(t)|^2\big ]&\le 4T^2\mathbb {E}\Big [\int _0^t \exp \Big (4\int _{s}^t\int _0^1\partial _xb_{1,n}(r, \Lambda _n(\lambda ,r)) \nonumber \\&\quad + \partial _xb_2(r,\Lambda _n(\lambda ,r),\alpha _1(r))\textrm{d}\lambda \textrm{d}r \Big )\textrm{d} s\Big ]^{1/2}\nonumber \\&\quad \times \mathbb {E}\Big [\int _0^{t}|b_1(s, X^{\alpha _2}(s)) - b_{1,n}(s, X^{\alpha _2}(s))|^4 \nonumber \\&\quad + |b_2(s, X^{\alpha _2}(s),\alpha _1(s)) \nonumber \\&\quad - b_2(s, X^{\alpha _2}(s),{\alpha _2}(s))|^4\,\textrm{d}s\Big ]^{1/2}. \end{aligned}$$
(8)

By the Lipschitz continuity of \(b_2\), the last term on the right hand side is estimated as

$$\begin{aligned}{} & {} \mathbb {E}\Big [\int _0^T|b_2(s, X^{\alpha _2}(s),\alpha _1(s)) - b_2(s, X^{\alpha _2}(s),{\alpha _2}(s))|^4\,\textrm{d}s \Big ] \nonumber \\{} & {} \quad \le C\mathbb {E}\Big [\int _0^T|\alpha _1(s) - \alpha _2(s)|^4\,\textrm{d}s \Big ]\le C(\delta (\alpha _1,\alpha _2))^4. \end{aligned}$$
(9)

Moreover, denoting

$$\begin{aligned} \mathcal {E}\Big (\int _0^Tq(s)\,\textrm{d}B(s) \Big ) = \exp \Big (\int _0^Tq(s)\,\textrm{d}B(s) - \frac{1}{2}\int _0^T|q(s)|^2\,\textrm{d}s \Big ), \end{aligned}$$

the second integral on the right side of (8) can be further estimated as follows:

$$\begin{aligned}&\mathbb {E}\Big [\int _0^{T}|b_1(s, X^{\alpha _2}(s)) - b_{1,n}(s, X^{\alpha _2}(s))|^4\textrm{d}s\Big ]\\&\quad = \mathbb {E}\Big [\mathcal {E}\Big (\frac{\sigma ^\top }{|\sigma |^2}\int _0^Tb(s, X^{\alpha _2}(s),{\alpha _2}(s))\textrm{d}B(s) \Big )^{1/2}\\&\qquad \times \mathcal {E}\Big (\int _0^T\frac{\sigma ^\top }{|\sigma |^2}b(s, X^{\alpha _2}(s),{\alpha _2}(s))\textrm{d}B(s) \Big )^{-1/2}\\&\qquad \times \int _0^{T}|b_1(s, X^{\alpha _2}(s)) - b_{1,n}(s, X^{\alpha _2}(s))|^4\textrm{d}s \Big ]\\&\quad \le C\mathbb {E}_{\mathbb {Q}}\Big [ \int _0^{T}|b_1(s, X^{\alpha _2}(s)) - b_{1,n}(s, X^{\alpha _2}(s))|^8\textrm{d}t \Big ]^{1/2} \end{aligned}$$

for some constant \(C>0\) and the probability measure \(\mathbb {Q}\) is the measure with density

$$\begin{aligned} \frac{\,\textrm{d}\mathbb {Q}}{\,\textrm{d}\mathbb {P}}:= \mathcal {E}\Big (\int _0^T\frac{\sigma ^\top }{|\sigma |^2}b(s, X^{\alpha _2}(s),{\alpha _2}(s))\textrm{d}B(s) \Big ). \end{aligned}$$
(10)

Note that we used the Cauchy–Schwarz inequality and then the fact that b is bounded to get \(\mathbb {E}[(\frac{\,\textrm{d}\mathbb {Q}}{\,\textrm{d}\mathbb {P}})^{-1}]\le C\). By the Girsanov’s theorem, under the measure \(\mathbb {Q}\), the process \((X^{\alpha _2}(t) - x_0)\sigma ^\top /|\sigma |^2 \) is a Brownian motion. Thus, it follows that

$$\begin{aligned}&\mathbb {E}_{\mathbb {Q}}\Big [\int _0^{T}|b_1(s, X^{\alpha _2}(s)) - b_{1,n}(s, X^{\alpha _2}(s))|^8 \textrm{d}s\Big ]^{1/2} \\&\quad \le C\mathbb {E}\Big [ \int _0^{T}|b_1(s, x_0+ \sigma B(s)) - b_{1,n}(s,x_0+ \sigma B(s))|^8\textrm{d}s \Big ]^{1/2} \end{aligned}$$

and using the density of Brownian motion, we have for every \(p\ge 1\)

$$\begin{aligned}&\mathbb {E}\Big [\Big | b_{1}(s,x_0+ \sigma B(s))- b_{1,n}(s,x_0+\sigma B(s))\Big |^p\Big ]\\&\quad = \frac{1}{\sqrt{2\pi s}}\int _{\mathbb {R}^d}\Big |b_{1,n} (s,x_0+\sigma y)-b_{1} (s,x_0+\sigma y)\Big |^pe^{-\frac{|y|^2}{2s}}\textrm{d}y\\&\quad =\frac{1}{\sqrt{2\pi s}}\int _{\mathbb {R}^d}\Big |b_{1,n} (s,\sigma y)-b_{1} (s,\sigma y)\Big |^pe^{-\frac{|y-x_0|^2}{2s}}\textrm{d}y\\&\quad =\frac{1}{\sqrt{2\pi s}}\int _{\mathbb {R}^d}\Big |b_{1,n} (s,\sigma y)-b_{1} (s,\sigma y)\Big |^pe^{-\frac{|y-2x_0|^2}{4s}}e^{-\frac{|y|^2}{4s}}e^{\frac{|x_0|^2}{2s}}\textrm{d}y\\&\quad \le \frac{1}{\sqrt{2\pi s}}e^{\frac{|x_0|^2}{2s}}\int _{\mathbb {R}^d}\big |b_{1,n} (s,\sigma y)-b_{1} (s,\sigma y)\big |^pe^{-\frac{|y|^2}{4s}}\textrm{d}y. \end{aligned}$$

By the Fubini’s theorem, this shows that

$$\begin{aligned}&\mathbb {E}\Big [\int _0^{T}|b_1(s, X^{\alpha _2}(s)) - b_{1,n}(s, X^{\alpha _2}(s))|^8\textrm{d}s\Big ]\nonumber \\&\quad \le C\Big (\int _0^T\frac{1}{\sqrt{2\pi s}}e^{\frac{|x_0|^2}{2s}}\int _{\mathbb {R}^d}\big |b_{1,n} (s,\sigma y)-b_{1} (s,\sigma y)\big |^8e^{-\frac{|y|^2}{4s}}\textrm{d}y\,\textrm{d}s\Big )^{1/2}. \end{aligned}$$
(11)

Let us now turn our attention to the first term in (8). Since \(\Lambda _{n}(\lambda ,t)\) takes the form

$$\begin{aligned} \Lambda _{n}(\lambda , t)&= x_0 {+} \int _0^t\Big \{\lambda b_{n}(s, X^{\alpha _1}_n(s),\alpha _1(s)) {+} (1{-}\lambda )b(s, X^{\alpha _2}(s),{\alpha _2}(s))\Big \}\,\textrm{d}s {+} \sigma B(t)\\&= x_0 {+}\int _0^tb^{\lambda ,\alpha _2}(s)\textrm{d}s {+} \sigma B(t), \end{aligned}$$

where \(b^{\lambda ,\alpha _2}(s)\) is the short hand notation for \(\lambda b_{n}(s, X^{\alpha _1}_n(s),\alpha _1(s)) + (1-\lambda )b(s, X^{\alpha _2}(s),{\alpha _2}(s))\). We use the Jensen’s inequality, the Girsanov’s theorem as above and Lipschitz continuity of \(b_2\) to get

$$\begin{aligned}&\mathbb {E}\Big [\exp \Big (4\int _{s}^t\int _0^1\partial _{x}b_{1,n}(r, \Lambda _n(\lambda ,r))+\partial _xb_2(r,\Lambda _n(\lambda ,r),\alpha _1(r))\,\textrm{d}\lambda \textrm{d}r\Big ) \Big ]\nonumber \\&\quad \le C\int _0^1 \mathbb {E}_{\mathbb {Q}^\lambda }\Big [\exp \Big (8\int _{s}^t\partial _xb_{1,n}(r, \Lambda _n(\lambda ,r))\textrm{d}r \Big ) \Big ]^{1/2}\,\textrm{d}\lambda \nonumber \\&\quad \le C\int _0^1\mathbb {E}\Big [\exp \Big (8\int _{s}^t\partial _xb_{1,n}(r, x_0+ \sigma B(r))\textrm{d}r \Big ) \Big ]^{1/2}\,\textrm{d}\lambda , \end{aligned}$$
(12)

with \(\,\textrm{d}\mathbb {Q}^\lambda = \mathcal {E}\big (\frac{\sigma ^\top }{|\sigma |^2}\int _0^Tb^{\lambda ,\alpha _2}(s)\textrm{d}B(s) \big )\,\textrm{d}\mathbb {P}\), and where we used the fact that \(b^{\lambda ,\alpha _2}\) is bounded. Since the sequence \((b_{1,n})_n\) is uniformly bounded, it follows from Lemma A.3 that

$$\begin{aligned} \sup _nE\Big [\exp \Big (4\int _{s}^t\partial _xb_{1,n}(r, x_0+ \sigma \cdot B(r))\textrm{d}r \Big ) \Big ] \le C. \end{aligned}$$
(13)

Therefore, putting together (8), (9), (11) and (13) concludes the proof.

Since \(b_k\) is Lipschitz continuous the convergence (ii) follows by classical arguments, the proof is omitted. \(\square \)

Lemma 2.4

Let \(\alpha \in \mathcal {A}\) and let \(\alpha _n\) be a sequence of admissible controls such that \(\delta (\alpha _n,\alpha )\rightarrow 0\). Then, it holds

  1. (i)

    \(| J_k(\alpha _n) - J_k(\alpha ) | \rightarrow 0\) as \(n\rightarrow \infty \) for every \(k \in \mathbb {N}\) fixed. In particular, the function \(J_k:(\mathcal {A},\delta ) \rightarrow \mathbb {R}\) is continuous.

  2. (ii)

    \(|J_n(\alpha ) - J(\alpha )| \le C\varepsilon _n\) for some \(C>0\) with \(\varepsilon _n\downarrow 0\).

Proof

  1. (i)

    The continuity of \(J_k\) easily follows by continuity of f and g. In fact, we have

    $$\begin{aligned} |J_k(\alpha _n) - J_k(\alpha )|&\le \mathbb {E}\Big [|g(X^{\alpha _n}_k(T)) - g(X^{\alpha }_k(T))| + \int _0^T|f(t, X^{\alpha _n}_k(t), \alpha _n(t)) \\&\quad - f(t, X^{\alpha }_k(t), \alpha (t)) |\,\textrm{d}t \Big ]\rightarrow 0, \end{aligned}$$

    where the convergence follows by the dominated convergence and Lemma 2.3.

  2. (ii)

    is also a direct consequence of Lemma 2.3. In fact, by the linear growth of \(\partial _xf\) and \(\partial _xg\) we have

    $$\begin{aligned}&| J_n(\alpha ) - J(\alpha ) |\le \mathbb {E}\Big [|g(X^{\alpha }_n(T)) - g(X^{\alpha }(T))| + \int _0^T|f(t, X^{\alpha }_n(t), \alpha (t)) \\&\qquad - f(t, X^{\alpha }(t), \alpha (t)) |\,\textrm{d}t \Big ]\\&\quad \le \mathbb {E}\Big [\int _0^1|\partial _xg(\lambda X^{\alpha }_n(T)) + (1-\lambda )X^{\alpha }(T))|\,\textrm{d}\lambda |X^{\alpha }_n(T) - X^{\alpha }(T)|\Big ]\\&\qquad + \mathbb {E}\Big [\int _0^T\int _0^1|\partial _{x}f\big (t, \lambda X^{\alpha }_n(t)+(1-\lambda )X^{\alpha }(t),\alpha (t) \big )|\,\textrm{d}\lambda |X^{\alpha }_n(t) - X^{\alpha }(t)|\,\textrm{d}t \Big ]\\&\quad \le C\mathbb {E}\Big [1+\sup _{t\in [0,T]}\big ( |X^{\alpha }_n(t)|^2 + |X^\alpha (t)|^2\big ) \Big ]^{1/2}\Big (\sup _{t \in [0,T]}\mathbb {E}\Big [|X^\alpha _n(t) - X^\alpha (t)|^2\Big ]\Big )^{1/2}, \end{aligned}$$

where we used Cauchy–Schwarz inequality and Fubini’s theorem. Therefore, by Lemma 2.3(i) we have

$$\begin{aligned} | J_n(\alpha ) - J(\alpha ) |&\le C\sup _{t\in [0,T]}\mathbb {E}[|X^{\alpha }_n(t) - X^{\alpha }(t)|^2]^{1/2} \le C\varepsilon _n, \end{aligned}$$

where the second inequality follows from Lemma 2.3. \(\square \)

The next lemma pertains to the stability of the adjoint process with respect to the drift and the control process. This result is based on similar stability properties for stochastic flows. Given \(x \in \mathbb {R}\) and the solution \(X^{\alpha ,x}\) of the SDE (1) with initial condition \(X^{\alpha ,x}_t = x\), the first variation process of \(X^{\alpha ,x}\) is the derivative \(\Phi ^\alpha (t,s)\) of the function \(x\mapsto X^{\alpha ,x}(s)\). Existence and properties of this Sobolev differentiable flow have been extensively studied by Kunita [29] for equations with sufficiently smooth coefficients. In particular, when the drift b is Lipschitz and continuously differentiable, the function \(\Phi ^\alpha (t,s)\) exists and, for almost every \(\omega \), is the (classical) derivative of \(x\mapsto X^{\alpha ,x}(s)\). The case of measurable (deterministic) drifts is studied by Mohammed et. al. [42] and extended to measurable and random drifts in [38]. These works show that, when b is measurable, then \(X^{\alpha ,\cdot }(s)\in L^2(\Omega , W^{1,p}(U))\) for every \(s \in [t,T]\) and \(p>1\), where \(W^{1,p}(U)\) is the usual Sobolev space and U an open and bounded subset of \(\mathbb {R}\). That is, \(\Phi ^\alpha (t,s)\) exists and is the weak derivative of \(X^{\alpha ,\cdot }\).

The proof of the stability result will make use of an explicit representation of the process \(\Phi ^\alpha \) with respect to the local time-space integral. Recall that for \(a\in \mathbb {R}\) and \(X=\{X(t),t\ge 0\}\) a continuous semimartingale, the local time \(L^{X}(t,a)\) of X at a is defined by the Tanaka–Meyer formula as

$$\begin{aligned} |X(t)-a|=|X(0)-a|+\int _0^t{{\,\textrm{sgn}\,}}(X(s)-a)\textrm{d}X(s) +L^{X}(t,a), \end{aligned}$$

where \({{\,\textrm{sgn}\,}}(x)=-1_{(-\infty ,0]}(x)+1_{(0,+\infty )}(x)\). The local time-space integral plays a crucial role in the representations of the Sobolev derivative of the flows of the solution to the SDE (1). It is defined for functions in the space \((\mathcal {H}_x, \Vert \cdot \Vert ^x)\) defined (see e.g. [17]) as the space of Borel measurable functions \(f:[0,T]\times \mathbb {R}\rightarrow \mathbb {R} \) with the norm

$$\begin{aligned} \left\| f\right\| _x&:=2\left( \int _0^T\int _{\mathbb {R}}f^2(s,z)\exp (-\frac{|z-x|^2}{2s})\frac{\textrm{d}s \,\textrm{d}z}{\sqrt{2\pi s}}\right) ^{\frac{1}{2}}\\&\quad +\int _0^T\int _{\mathbb {R}}|z-x| |f(s,x)|\exp (-\frac{|z-x|^2}{2s})\frac{\textrm{d}s \,\textrm{d}z}{s\sqrt{2\pi s}}. \end{aligned}$$

Since \(b_1\) is bounded, we obviously have \(b_1 \in \mathcal {H}^x\) for every x. Moreover, it follows from [19] (see also [5]) that for every continuous semimartingale X the local time-space integral of \(f\in \mathcal {H}^x\) with respect to \(L^{X}(t,z)\) is well defined and satisfies

$$\begin{aligned} \int _0^t\int _{\mathbb {R}}f(s,z) L^{X}(\textrm{d}s,\textrm{d}z) = - \int _0^t\partial _xf(s,X(s))\textrm{d}\langle X\rangle _ s \end{aligned}$$
(14)

for every continuous function (in space) \(f \in \mathcal {H}^x\) admitting a continuous derivative \(\partial _xf(s,\cdot )\) (see [19, Lemma 2.3]). This representation allows to derive the following:

Lemma 2.5

For every \(\alpha \in \mathcal {A}\) and \(c\ge 0\), it holds

$$\begin{aligned} \mathbb {E}\Big [e^{c\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ] <\infty . \end{aligned}$$

Proof

First observe that for every \(n \in \mathbb {N}\), it follows by the Cauchy–Schwarz inequality that

$$\begin{aligned} \mathbb {E}\Big [e^{c\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ]&= \mathbb {E}\Big [\mathcal {E}\Big (\frac{\sigma ^\top }{|\sigma |^2}\int _0^Tb(u, X^{\alpha }(u),{\alpha }(u))\textrm{d}B(u) \Big )^{1/2}\\&\quad \times \, \mathcal {E}\Big (\int _0^T\frac{\sigma ^\top }{|\sigma |^2}b(u, X^{\alpha }(u),{\alpha }(u))\textrm{d}B(u) \Big )^{-1/2}\\&\quad \times e^{c\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)} \Big ]\\&\le C\mathbb {E}_{\mathbb {Q}}\Big [ e^{2c\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)} \Big ]^{1/2} \end{aligned}$$

where \(\mathbb {Q}\) is the probability measure given as in (10) with \(\alpha _2\) therein replaced by \(\alpha \). Hence, since \((X^{\alpha ,x}-x_0)\sigma ^\top /|\sigma |^2\) is a Brownian motion under \(\mathbb {Q}\), it follows by (14) that

$$\begin{aligned} \mathbb {E}\Big [e^{c\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ]&\le C\mathbb {E}_{\mathbb {Q}}\Big [ e^{-2c\Vert \sigma \Vert ^2\int _s^{t}\partial _xb_{1,n}\left( u,X^{\alpha ,x}(u)\right) \textrm{d}u} \Big ]^{1/2}\\&= C\mathbb {E}\Big [ e^{-2c\Vert \sigma \Vert ^2\int _s^{t}\partial _xb_{1,n}\left( u,x_0 + \sigma B(u)\right) \textrm{d}u} \Big ]^{1/2}\le {\overline{C}} \end{aligned}$$

for some constant \({\overline{C}}>0\) which does not depend on n, where this latter inequality follows by Lemma A.3. Since \(b_1\) is bounded and \(b_{1,n}\) converges to \(b_1\) pointwise, it follows by [19, Theorem 2.2] that \(\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z) \rightarrow \int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z) \) as n goes to infinity. Thus, it follows by the continuity of the exponential function and dominated convergence that

$$\begin{aligned} \mathbb {E}\Big [e^{c\int _s^{t}\int _{\mathbb {R}}b_{1}\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ] = \lim _{n \rightarrow \infty }\mathbb {E}\Big [e^{c\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ]<\overline{C}. \end{aligned}$$

\(\square \)

We are now ready to prove stability of the flow and of the adjoint processes.

Lemma 2.6

Let \(\alpha \in \mathcal {A}\) and \(\alpha _n\) be a sequence of admissible controls such that \(\delta (\alpha _n,\alpha )\rightarrow 0\). Then, the processes \(X^{\alpha _n}_n\) and \(X^{\alpha }\) admit Sobolev differentiable flows denoted \(\Phi ^{\alpha _n}_n\) and \(\Phi ^{\alpha }\), respectively and for every \(0\le t\le s\le T\) it holds

  1. (i)

    \(\mathbb {E}\big [|\Phi ^{\alpha _n}_n(t,s) - \Phi ^\alpha (t,s) |^2 \big ] \rightarrow 0\) as \(n\rightarrow \infty \),

  2. (ii)

    \(\mathbb {E}\big [| Y^{\alpha _n}_n(t) - Y^\alpha (t)| \big ] \rightarrow 0\) as \(n\rightarrow \infty \),

where \(Y^\alpha \) is the adjoint process defined as

$$\begin{aligned} Y^{\alpha }(t):= \mathbb {E}\Big [\Phi ^{\alpha }(t,T) \partial _xg( X^{\alpha }(T)) + \int _t^T\Phi ^{\alpha }(t,s) \partial _xf(s, X^{\alpha }(s), \alpha (s))\textrm{d}s\mid \mathcal {F}_t \Big ], \end{aligned}$$

and \(Y^{\alpha _n}_n\) is defined similarly, with \((X^{\alpha },\alpha , \Phi ^\alpha )\) replaced by \((X^{\alpha _n}_n,\alpha _n, \Phi ^{\alpha _n}_n)\).

Proof

The existence of the process \(\Phi ^{\alpha _n}_n\) is standard, it follows for instance by [28]. The existence of the flow \(\Phi ^{\alpha }\) follows by [38, Theorem 1.3]. We start by proving the first convergence claim. As explained above, these processes admit explicit representations in terms of the space-time local time process. It fact, it follows from Theorem A.1 that \(\Phi ^\alpha \) admits the representation

$$\begin{aligned} \Phi ^\alpha (t,s) = e^{\int _t^{s}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}e^{\int _t^{s}\partial _xb_2\left( u,X^{\alpha ,x}(u),\alpha (u)\right) \textrm{d}u} \end{aligned}$$

and \(\Phi _n^{\alpha _n}\) admits the same representation with \((b_1, X^{\alpha ,x},\alpha )\) replaced by \((b_{1,n}, X^{\alpha _n,x}, \alpha _n)\). Using these explicit representations and the Hölder inequality, we have

$$\begin{aligned}&\mathbb {E}\Big [\Big |\Phi ^{\alpha }(t,s)- \Phi _n^{\alpha _n}(t,s)\Big |^2\Big ]\\&\quad \le 2\mathbb {E}\Big [\Big |e^{\int _t^{s}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big \{e^{\int _t^{s}\partial _xb_2\left( u,X^{\alpha ,x}(u),\alpha (u)\right) \textrm{d}u}\\&\qquad -e^{\int _t^{s}\partial _xb_{2}\left( u,X_n^{\alpha _n,x}(u),\alpha _n(u)\right) \textrm{d}u}\Big \}\Big |^2\Big ] +2\mathbb {E}\Big [\Big |e^{\int _t^{s}\partial _xb_{2}\left( u,X_n^{\alpha _n,x}(u),\alpha _n(u)\right) \textrm{d}u}\\&\quad \Big \{e^{\int _t^{s}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)} -e^{\int _t^{s}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X_n^{\alpha _n,x}}(\textrm{d}u,\textrm{d}z)}\Big \}\Big |^2\Big ]\\&\quad \le 2 \mathbb {E}\Big [e^{4\int _t^{s}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{2}}\mathbb {E}\Big [\Big \{e^{\int _t^{s}\partial _xb_2\left( u,X^{\alpha ,x}(u),\alpha (u)\right) \textrm{d}u}\\&\qquad -e^{\int _t^{s}\partial _xb_{2}\left( u,X_n^{\alpha _n,x}(u),\alpha _n(u)\right) \textrm{d}u}\Big \}^4\Big ]^{\frac{1}{2}}+2\mathbb {E}\Big [e^{4\int _t^{s}\partial _xb_{2}\left( u,X_n^{\alpha _n,x}(u),\alpha _n(u)\right) \textrm{d}u}\Big ]^{\frac{1}{2}}\\&\qquad \times \, \mathbb {E}\Big [\Big \{e^{\int _t^{s}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)} -e^{\int _t^{s}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X_n^{\alpha _n,x}}(\textrm{d}u,\textrm{d}z)}\Big \}^4\Big ]^{\frac{1}{2}}. \end{aligned}$$

Splitting up the terms in power 4, then applying the Hölder and Young inequalities we continue the estimations as

$$\begin{aligned}&\mathbb {E}\Big [\Big |\Phi ^{\alpha }(t,s)- \Phi _n^{\alpha _n}(t,s)\Big |^2\Big ]\nonumber \\&\quad \le 2^7\mathbb {E}\Big [e^{4\int _t^{s}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{2}} \mathbb {E}\Big [\Big \{e^{6\int _t^{s}\partial _xb_2\left( u,X^{\alpha ,x}(u),\alpha (u)\right) \textrm{d}u}\nonumber \\&\qquad +e^{6\int _t^{s}\partial _xb_{2}\left( u,X_n^{\alpha _n,x}(u),\alpha _n(u)\right) \textrm{d}u}\Big \}\Big ]^{\frac{1}{4}} \nonumber \\&\qquad \times \mathbb {E}\Big [\Big \{e^{\int _t^{s}\partial _xb_2\left( u,X^{\alpha ,x}(u),\alpha (u)\right) \textrm{d}u} -e^{\int _t^{s}\partial _xb_{2}\left( u,X_n^{\alpha _n,x}(u),\alpha _n(u)\right) \textrm{d}u}\Big \}^2\Big ]^{\frac{1}{4}}\nonumber \\&\qquad + 2^7 \mathbb {E}\Big [\Big |e^{4\int _t^{s}\partial _xb_{2}\left( u,X_n^{\alpha _n,x}(u),\alpha _n(u)\right) \textrm{d}u}\Big ]^{\frac{1}{2}}\nonumber \\&\qquad \times \mathbb {E}\Big [\Big \{e^{6\int _t^{s}\int _{\mathbb {R}}b_1 \left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)} +e^{6\int _t^{s}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X_n^{\alpha _n,x}} (\textrm{d}u,\textrm{d}z)}\Big \}\Big ]^{\frac{1}{4}} \nonumber \\&\qquad \times \mathbb {E}\Big [\Big \{e^{\int _t^{s}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)} -e^{\int _t^{s}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X_n^{\alpha _n,x}}(\textrm{d}u,\textrm{d}z)}\Big \}^2\Big ]^{\frac{1}{4}}\nonumber \\&\quad =CI_1^{\frac{1}{2}}\times I^{\frac{1}{2}}_{2,n}\times I^{\frac{1}{4}}_{3,n} +CI^{\frac{1}{2}}_{4,n}\times I^{\frac{1}{4}}_{5,n}\times I^{\frac{1}{4}}_{6,n}. \end{aligned}$$
(15)

It follows from Lemma 2.5 that \(I_1\) and \(I_{5,n}\) are bounded. Since \(\partial _xb_2\) is bounded, it follows that \(I_{2,n}\) and \(I_{4,n}\) are also bounded with bounds independent on n. Let us now show that \(I_{3,n}\) and \(I_{6,n} \) converge to zero. We show only the convergence of \(I_{6,n}\) since that of \(I_{3,n}\) will follow (at least for a subsequence) from Lemma 2.3 and dominated convergence since \(\partial _xb_{2}\) is continuous and bounded.

To that end, further define the processes \(A_n^{\alpha _n}\) and \(A^{\alpha }\) by

$$\begin{aligned} A_n^{\alpha _n}(t,s):=e^{\int _t^{s}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X_n^{\alpha _n,x}}(\textrm{d}u,\textrm{d}z)}\quad \text {and} \quad A^{\alpha }(t,s):=e^{\int _t^{s}\int _{\mathbb {R}}b_1\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}. \end{aligned}$$

In order to show that \(A_n^{\alpha _n}\) converges to \(A^{\alpha }\) in \(L^2\), we will show that \(A_n^{\alpha _n}\) converges weakly to \(A^{\alpha }\) in \(L^2\) and that \(E[|A_n^{\alpha _n}|^2]\) converges to \(E[|A^{\alpha }|^2]\) in \(\mathbb {R}\). We first prove the weak convergence. Since the set

$$\begin{aligned} \Big \{\mathcal {E}\Big (\int _0^1{\dot{\varphi }}(s)\textrm{d}B(s)\Big ):\varphi \in C^{1}_b([0,T],\mathbb {R}^d)\Big \} \end{aligned}$$

spans a dense subspace in \(L^2(\Omega )\), in order to the show weak convergence, it is enough to show that

$$\begin{aligned}{} & {} E\Big [A_n^{\alpha _n}(t,s) \mathcal {E}\Big (\int _0^1{\dot{\varphi }}(s)\textrm{d}B(s)\Big )\Big ]\rightarrow E\Big [A^{\alpha }(t,s) \mathcal {E}\Big (\int _0^1{\dot{\varphi }}(s)\textrm{d}B(s)\Big )\Big ]\quad \text {for every}\quad \\{} & {} \varphi \in C^{1}_b([0,T],\mathbb {R}^d). \end{aligned}$$

Denote by \({\tilde{X}}_n^{\alpha _n, x}\) and \({\tilde{X}}^{\alpha ,x}\) the processes given by

$$\begin{aligned} \textrm{d}{\tilde{X}}^{{\tilde{\alpha }}_n,x}_n(t)= \Big (b_{1,n}(t,\tilde{X}^{{\tilde{\alpha }}_n,x}_n(t))+ b_{2}(t,{\tilde{X}}^{\tilde{\alpha }_n,x}_n(t),{\tilde{\alpha }}_n)+\sigma {{\dot{\varphi }}}(t) \Big )\textrm{d}t +\sigma \textrm{d}B(t), \end{aligned}$$
(16)

and

$$\begin{aligned} \textrm{d}{\tilde{X}}^{{\tilde{\alpha }},x}(t) = \Big (b_{1}(t,\tilde{X}^{{\tilde{\alpha }},x}(t))+ b_{2}(t,{\tilde{X}}^{\tilde{\alpha },x}(t),{\tilde{\alpha }}_n)+\sigma {{\dot{\varphi }}}(t)\Big )\textrm{d}t +\sigma \textrm{d}B(t). \end{aligned}$$
(17)

Observe that these processes are well-defined, since we have \(\tilde{X}^{{\tilde{\alpha }},x}(t,\omega ) = X^{\alpha ,x}(t,\omega + \varphi )\) and \({\tilde{X}}_n^{{\tilde{\alpha }}_n,x}(t,\omega ) = X_n^{\alpha _n,x}(t,\omega + \varphi )\). Using the Cameron–Martin–Girsanov theorem as in the proof of Lemma 2.5, we have

$$\begin{aligned}&\Big |\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T{\dot{\varphi }}(s)\textrm{d}B(s)\Big )\Big \{A_n^{\alpha _n}(t,s)-A^{\alpha }(t,s)\Big \}\Big ]\Big |\\&\quad =\Big |\mathbb {E}\Big [e^{\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{{\tilde{X}}_n^{{\tilde{\alpha }}_n,x}}(\textrm{d}u,\textrm{d}z)}-e^{\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{{\tilde{X}}^{{\tilde{\alpha }},x}}(\textrm{d}u,\textrm{d}z)}\Big ]\Big |\\&\quad =\Big |\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\Big \{{\tilde{u}}_n(s,x+\sigma \cdot B(s),\alpha _n (s))+\sigma \cdot {\dot{\varphi }}(s)\Big \}\textrm{d}B(s)\Big )e^{\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{|\sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\\&\qquad -\mathcal {E}\Big (\int _0^T\Big \{{\tilde{u}}(s,x+\sigma \cdot B(s),\alpha (s))+\sigma \cdot {\dot{\varphi }}(s)\Big \}\textrm{d}B(s)\Big )e^{\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{|\sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]\Big |\\&\quad =\Big |\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\Big \{{\tilde{u}}_n(s,x+\sigma \cdot B(s),\alpha _n (s))+\sigma \cdot {\dot{\varphi }}(s)\Big \}\textrm{d}B(s)\Big )e^{\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{|\sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\\&\qquad -\mathcal {E}\Big (\int _0^T\Big \{{\tilde{u}}(s,x+\sigma \cdot B(s),\alpha (s))+\sigma \cdot {\dot{\varphi }}(s)\Big \}\textrm{d}B(s)\Big )e^{\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{|\sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]\Big | , \end{aligned}$$

where \({\tilde{u}}(s, x,\alpha (\omega )): = u(s, x,\alpha (\omega +\varphi ))\). Let us set

$$\begin{aligned} u(s, x,\alpha (\omega )): = \Big (\frac{\sigma ^1b}{|\sigma |^2}, \dots , \frac{\sigma ^db}{|\sigma |^2} \Big )(t,x,\alpha (\omega ))\quad \text {and}\quad B^x_\sigma :=x+\sum _{i=1}^d\frac{\sigma _i}{\Vert \sigma \Vert }B^i. \end{aligned}$$

Next, add and subtract the term \(\mathcal {E}\Big (\int _0^T\Big \{{\tilde{u}}_n(s,x+\sigma \cdot B(s),\alpha _n (s))+\sigma \cdot {\dot{\varphi }}(s)\Big \}\textrm{d}B(s)\Big )e^{\int _s^{t}\int _{\mathbb {R}}b_{1}\left( u,z\right) L^{|\sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\) and then use the inequality \(|e^x-e^y|\le |x-y||e^x+e^y|\) to obtain

$$\begin{aligned}&\Big |\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T{\dot{\varphi }}(s)\textrm{d}B(s)\Big )\Big \{A_n^{\alpha _n}(t,s)-A^{\alpha }(t,s)\Big \}\Big ]\Big |\\&\quad \le \Big |\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\{u_n(s,x+\sigma \cdot B(s),\alpha (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\\&\qquad \times \Big |\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)-\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)\Big |\\&\qquad \times \Big (e^{\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}+e^{\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big )\Big ]\Big |\\&\qquad +\Big |\mathbb {E}\Big [e^{\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big \{\mathcal {E}\Big (\int _0^T\{u_n(s,x+\sigma \cdot B(s),\alpha _n (s,\omega +\varphi ))\\&\qquad +\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\\&\qquad -\mathcal {E}\Big (\int _0^T\{u(s,x+\sigma \cdot B(s),\alpha (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\Big \}\Big ]\Big |. \end{aligned}$$

Therefore, an application of the Hölder’s inequality yields the estimate

$$\begin{aligned}&\Big |\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T{\dot{\varphi }}(s)\textrm{d}B(s)\Big )\Big \{A_n^{\alpha _n}(t,s)-A^{\alpha }(t,s)\Big \}\Big ]\Big |\nonumber \\&\quad \le 4\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\{u_n(s,x+\sigma \cdot B(s),\alpha _n (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )^4\Big ]^{\frac{1}{4}}\nonumber \\&\qquad \times \, \mathbb {E}\Big [\Big |\int _s^{t}\int _{\mathbb {R}}\Big (b_{1,n}\left( u,z\right) -b_1\left( u,z\right) \Big )L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)\Big |^2\Big ]^{\frac{1}{2}}\nonumber \\&\qquad \times \mathbb {E}\Big [e^{4\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}+e^{4\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{4}}\nonumber \\&\qquad +\mathbb {E}\Big [e^{2\int _s^{t}\int _{\mathbb {R}}b_1\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{2}}\mathbb {E}\Big [\Big \{\mathcal {E}\Big (\int _0^T\{u_n(s,x+\sigma \cdot B(s),\alpha _n (s,\omega +\varphi ))\nonumber \\&\qquad +{\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\nonumber \\&\qquad -\mathcal {E}\Big (\int _0^T\{u(s,x+\sigma \cdot B(s),\alpha (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\Big \}^2\Big ]^{\frac{1}{2}}\nonumber \\&\quad =J_{1,n}^{\frac{1}{4}}\times J_{2,n}^{\frac{1}{2}}\times J_{3,n}^{\frac{1}{4}}+J_{4}^{\frac{1}{2}}\times J_{5,n}^{\frac{1}{2}}. \end{aligned}$$
(18)

Using Lemma A.2, it follows that \(J_{2,n}\) converges to zero, and by the dominated convergence, the definition of u, the inequality \(|e^x-e^y|\le |x-y||e^x+e^y|\) and once more Lemma A.2, \(J_{5,n}\) also converges to zero. Thanks to Lemma A.3 and boundedness of \(b_{1,n}\) (respectively \(b_1\)), the term \(J_{3,n}\) (respectively \(J_{4,n}\)) is bounded. The bound of \(J_{1,n}\) follows by the uniform boundedness of \(u_n\).

It remains to show that \(\mathbb {E}[|A_n^{\alpha _n}(t)|^2]\) converges to \(\mathbb {E}[|A^{\alpha }(t)|^2]\) in \(\mathbb {R}\). Using the Girsanov transform as in the proof of Lemma 2.5, we have

$$\begin{aligned}&\mathbb {E}[|A_n^{\alpha _n}(t)|^2]=\mathbb {E}\Big [e^{2\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{X_n^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ]\nonumber \\&\quad = \mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\{u_n(s,x+\sigma \cdot B(s),\alpha _n (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\nonumber \\&\quad e^{2\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ] \end{aligned}$$
(19)

and

$$\begin{aligned}&\mathbb {E}[|A^{\alpha }(t)|^2]=\mathbb {E}\Big [e^{2\int _s^{t}\int _{\mathbb {R}}b_{1}\left( u,z\right) L^{X^{\alpha ,x}}(\textrm{d}u,\textrm{d}z)}\Big ]\nonumber \\&\quad = \mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\{u(s,x+\sigma \cdot B(s),\alpha (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\nonumber \\&\qquad e^{2\int _s^{t}\int _{\mathbb {R}}b_{1}\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]. \end{aligned}$$
(20)

Therefore using once more \(|e^x-e^y|\le |x-y||e^x+e^y|\) and the Cauchy–Schwarz inequality

$$\begin{aligned}&|\mathbb {E}[|A_n^{\alpha _n}(t)|^2]-\mathbb {E}[|A^{\alpha }(t)|^2]|\\&\quad =\Big |\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\{u_n(s,x+\sigma \cdot B(s),\alpha _n (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\nonumber \\&\qquad e^{2\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{\Vert \sigma \Vert B^{x}_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]\\&\qquad -\mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\{u(s,x+\sigma \cdot B(s),\alpha (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\nonumber \\&\qquad e^{2\int _s^{t}\int _{\mathbb {R}}b_{1}\left( u,z\right) L^{\Vert \sigma \Vert B^x_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]\Big |\\&\quad \le \Big |\mathbb {E}\Big [e^{4\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{\Vert \sigma \Vert B^{x}_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{2}}\mathbb {E}\Big [ \Big \{\mathcal {E}\Big (\int _0^T\{u_n(s,x+\sigma \cdot B(s),\alpha _n (s,\omega +\varphi ))\\&\qquad +\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )-\mathcal {E}\Big (\int _0^T\{u(s,x+\sigma \cdot B(s),\alpha (s,\omega +\varphi ))\\&\qquad +\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\Big \}^2\Big ]^{\frac{1}{2}}\Big |\\&\qquad +C\Big |\mathbb {E}\Big [\Big (\int _s^{t}\int _{\mathbb {R}}\{b_{1,n}\left( u,z\right) -b_{1}\left( u,z\right) \}L^{\Vert \sigma \Vert B^{x}}(\textrm{d}u,\textrm{d}z)\Big )^2\Big ]^{\frac{1}{2}}\\&\qquad \times \Big (\mathbb {E}\Big [e^{8\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{\Vert \sigma \Vert B^{x}_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{4}}+\mathbb {E}\Big [e^{8\int _s^{t}\int _{\mathbb {R}}b_{1}\left( u,z\right) L^{\Vert \sigma \Vert B^{x}_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{4}}\Big )\\&\qquad \times \mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\{u(s,x+\sigma \cdot B(s),\alpha (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )^4\Big ]^{\frac{1}{4}}\Big |. \end{aligned}$$

Now, introducing the random variables

$$\begin{aligned} V_n&:{=} \int _0^T\Big (u_n(s,x{+}\sigma \cdot B(s),\alpha _n (s,\omega {+}\varphi )){-}u(s,x{+}\sigma \cdot B(s),\alpha (s,\omega {+}\varphi ))\Big )\textrm{d}B(s)\\&\qquad {-}\frac{1}{2}\int _0^T\Big (|u_n(s,x{+}\sigma \cdot B(s),\alpha _n (s,\omega {+}\varphi )){+}\sigma \cdot {\dot{\varphi }}(s)|^2\\&\qquad {-}|u(s,x{+}\sigma \cdot B(s),\alpha (s,\omega +\varphi )){+}\sigma \cdot {\dot{\varphi }}(s)|^2\Big )\textrm{d} s \end{aligned}$$

and

$$\begin{aligned} F_{1,n} := \int _s^{t}\int _{\mathbb {R}}\{b_{1,n}\left( u,z\right) -b_{1}\left( u,z\right) \}L^{\Vert \sigma \Vert B^{x}}(\textrm{d}u,\textrm{d}z) \end{aligned}$$

we continue the above estimations as

$$\begin{aligned}&|\mathbb {E}[|A_n^{\alpha _n}(t)|^2] - \mathbb {E}[|A^{\alpha }(t)|^2]|\nonumber \\&\quad {\le } C\mathbb {E}\Big [V_n^2\Big \{\mathcal {E}\Big (\int _0^T\{u_n(s,x{+}\sigma \cdot B(s),\alpha _n (s,\omega {+}\varphi )){+}\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )\nonumber \\&\qquad {+} \mathcal {E}\left( \int _0^T\{u(s,x{+}\sigma \cdot B(s),\alpha (s,\omega +\varphi )){+}\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\right) \Big \}^2\Big ]\nonumber \\&\qquad {+}C\Big |\mathbb {E}\Big [|F_{1,n}|^2\Big ]^{\frac{1}{2}} \Big (E\Big [e^{8\int _s^{t}\int _{\mathbb {R}}b_{1,n}\left( u,z\right) L^{\Vert \sigma \Vert B^{x}_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{4}}{+} \mathbb {E}\Big [e^{8\int _s^{t}\int _{\mathbb {R}}b_{1}\left( u,z\right) L^{\Vert \sigma \Vert B^{x}_\sigma }(\textrm{d}u,\textrm{d}z)}\Big ]^{\frac{1}{4}}\Big )\nonumber \\&\qquad \times \mathbb {E}\Big [\mathcal {E}\Big (\int _0^T\{u(s,x{+}\sigma \cdot B(s),\alpha (s,\omega +\varphi ))+\sigma \cdot {\dot{\varphi }}(s)\}\textrm{d}B(s)\Big )^4\Big ]^{\frac{1}{4}}\Big |. \end{aligned}$$
(21)

By Lemma A.2, \(F_{1,n}\) converges to zero in \(L^2(\Omega )\). Using similar arguments as in [5, Lemma A.3], one can show that \(V_n\) converges to zero in \(L^2(\Omega )\) by the boundedness of \(u_n\) and the definition of the distance \(\delta \). Observe however that in this case, \(u_n\) depends on \(\alpha _n\) and not on \(\alpha \) as in [5, Lemma A.3]. Nevertheless using the fact that \(b_{1,n}\), \(b_1\) and \(b_2\) are bounded and Lipschitz in the second variable, one can show by the dominated convergence theorem and similar reasoning as in (9) that the overall term converges to zero. It is also worth mentioning that the other terms are uniformly bounded by application of either the Girsanov theorem and/or Lemma A.3 to the uniformly bounded sequences \((u_n)_{n\ge 1},(b_{1,n})_{n\ge 1}\) and the bounded functions \(u, b_{1}\).

Let us now turn our attention to the proof of (ii). Compute the difference \(Y_{n}^{\alpha _n}(t)-Y^\alpha (t)\), add and subtract the terms \(\Phi ^{\alpha }(t,T) \partial _xg(X_n^{\alpha _n}(T))\) and \(\int _t^T\Phi ^{\alpha }(t,u) \partial _xf(u,X_n^{\alpha _n}(u), \alpha _n(u))\,\textrm{d}u\) and then apply Hölder’s inequality to obtain

$$\begin{aligned}&\mathbb {E}[|Y_{n}^{\alpha _n}(t)-Y^\alpha (t)|]\nonumber \\&\quad \le C_T\Big \{\mathbb {E}\Big [\Big |\Phi ^{\alpha }(t,T)\Big |^2\Big ]^{\frac{1}{2}}\mathbb {E}\Big [|\partial _xg( X_n^{\alpha _n}(T)) - \partial _xg( X^{\alpha }(T))|^2\Big ]^{\frac{1}{2}}\nonumber \\&\qquad +\mathbb {E}\Big [|\partial _xg( X_n^{\alpha _n}(T))|^2\Big ]^{\frac{1}{2}}\mathbb {E}\Big [\Big |\Phi _n^{\alpha _n}(t,T) - \Phi ^{\alpha }(t,T)\Big |^2\Big ]^{\frac{1}{2}}\nonumber \\&\qquad +\mathbb {E}\Big [\int _t^T|\Phi ^{\alpha }(t,u)|^2\,\textrm{d}u\Big ]^{\frac{1}{2}} \mathbb {E}\Big [\int _0^T|\partial _xf(u, X^{\alpha }(u), \alpha (u))-\partial _xf(u, X_n^{\alpha _n}(u), \alpha _n(u))|^2\,\textrm{d}u\Big ]^{\frac{1}{2}}\nonumber \\&\qquad +\mathbb {E}\Big [\int _0^T|\partial _xf(u, X_n^{\alpha _n}(u), \alpha _n(u))|^2\,\textrm{d}u\Big ]^{\frac{1}{2}}\mathbb {E}\Big [\int _0^T|\Phi _n^{\alpha _n}(u)-\Phi ^{\alpha }(u)|^2\,\textrm{d}u\Big ]^{\frac{1}{2}}\Big \} \end{aligned}$$
(22)

for some constant \(C_T\) depending only on T. Since the process \(\Phi ^{\alpha }\) is square integrable, (see [38, Theorem 1.3]) it follows by boundedness and continuity of \(\partial _xg,\partial _xf\) as well as Lemma 2.3 that the first and third terms converge to zero as n goes to infinity. Moreover, by

linear growth of \(\partial _xf\) and \(\partial _xg\), Lemma 2.3(i) and the \(L^2\) convergence of \(\Phi _n^{\alpha _n}(t,u)\) to \(\Phi ^{\alpha }(t,u)\) given in part (i), we conclude that the second and last terms in (22) converge to zero, which shows (ii). \(\square \)

Proof

(of Theorem 1.1) Let \({\hat{\alpha }}\) be an optimal control and \(n\ge 1\) fixed. Observe that by the linear growth assumption on fg the function \(J_n\) is bounded from above. By Lemma 2.4 the function \(J_n\) is also continuous on \((\mathcal {A},\delta )\) and there exists \(\varepsilon _n\) such that

$$\begin{aligned} J({\hat{\alpha }}) - J_n({\hat{\alpha }})\le \varepsilon _n \text { and } J_n(\alpha ) - J(\alpha ) \le \varepsilon _n\quad \text {for all } \alpha \in \mathcal {A}. \end{aligned}$$

That is, \(J_n({\hat{\alpha }}) \le \inf _{\alpha \in \mathcal {A}}J_n(\alpha ) + 2\varepsilon _n\). Thus, by Ekeland’s variational principle, see e.g. [20], there is a control \({\hat{\alpha _n}} \in \mathcal {A}\) such that \(\delta ({\hat{\alpha }}, {\hat{\alpha _n}})\le (2\varepsilon _n)^{1/2}\) and

$$\begin{aligned} J_n({\hat{\alpha _n}}) \le J_n(\alpha ) + (2\varepsilon _n)^{1/2}\delta ({\hat{\alpha _n}},\alpha )\quad \text {for all}\quad \alpha \in \mathcal {A}. \end{aligned}$$

In other words, putting \(J^\varepsilon _n(\alpha ):= J_n(\alpha ) + (2\varepsilon _n)^{1/2}\delta ({\hat{\alpha _n}},\alpha )\), the control process \({\hat{\alpha _n}}\) is optimal for the problem with cost function \(J^\varepsilon _n\).

Now, let \(\beta \in \mathcal {A}\) be an arbitrary control and \(\varepsilon >0\) a fixed constant. By convexity of \(\mathbb {A}\), it follows that \({\hat{\alpha _n}} + \varepsilon \eta \in \mathcal {A}\), with \(\eta := \beta - {\hat{\alpha _n}}\). Thus, since \(b_n\) is sufficiently smooth, it is standard that the functional \(J_n\) is Gâteau differentiable (see [10, Lemma 4.8]) and its Gâteau derivative in the direction \(\eta \) is given by

$$\begin{aligned} \frac{\textrm{d}}{\textrm{d}\varepsilon }J_n(\alpha + \varepsilon \eta )_{|_{\varepsilon = 0}}&= \mathbb {E}\Big [\int _0^T\partial _xf(t, X_n^{{\hat{\alpha _n}}}(t), {\hat{\alpha _n}}(t))V_n(t) \\&\qquad + \partial _{\alpha }f(t, X_n^{{\hat{\alpha _n}}}(t), {\hat{\alpha _n}}(t))\eta (t)\textrm{d}t + \partial _xg(X_n^{{\hat{\alpha _n}}}(T))V_n(T) \Big ] , \end{aligned}$$

where \(V_n\) is the stochastic process solving the linear equation

$$\begin{aligned} \textrm{d}V_n(t) = \partial _xb_n(t, X_n^\alpha (t),\alpha (t))V_n(t)\textrm{d}t + \partial _\alpha b_n(t, X_n^\alpha (t),\alpha (t))\eta (t)\textrm{d}t,\quad V_n(0) = 0. \end{aligned}$$

On the other hand using triangular inequality, we have

$$\begin{aligned} \lim _{\varepsilon \downarrow 0}\frac{1}{\varepsilon }\big (\delta ({\hat{\alpha _n}}, \alpha + \varepsilon \eta ) - \delta ({\hat{\alpha _n}}, \alpha ) \big ) \le {\varepsilon }\mathbb {E}\big [ \sup _{t \in [0,T]}|\eta (t)|^{4} \big ]^{1/{4}}. \end{aligned}$$

Therefore, \(J^\varepsilon _n\) is also Gâteau differentiable and since \({\hat{\alpha _n}}\) is optimal for \(J^\varepsilon _n\), we have

$$\begin{aligned} 0\le \frac{\textrm{d}}{\textrm{d}\varepsilon }J^\varepsilon _n({\hat{\alpha _n}} + \varepsilon \eta )_{|_{\varepsilon = 0}}&= \frac{\textrm{d}}{\textrm{d}\varepsilon }J_n({\hat{\alpha _n}} + \varepsilon \eta )_{|_{\varepsilon = 0}} + \lim _{\varepsilon \downarrow 0} (2\varepsilon _n)^{1/2}\frac{1}{\varepsilon }\delta ({\hat{\alpha _n}},{\hat{\alpha _n}} + \varepsilon \eta ) \\&{\le }\mathbb {E}\Big [\int _0^T\partial _xf\big (t, X_n^{{\hat{\alpha _n}}}(t), {\hat{\alpha _n}}(t) \big )V_n(t) + \partial _{\alpha }f\big ( t, X_n^{{\hat{\alpha _n}}}(t), {\hat{\alpha _n}}(t) \big )\eta (t)\textrm{d}t\\&\qquad + \partial _xg(X_n^{{\hat{\alpha _n}}}(T))V_n(T) \Big ] + \big (2\varepsilon _n)^{1/2}(E[\sup _t|\eta (t)|^{4}] \big )^{1/{4}}\\&\le \mathbb {E}\Big [\int _0^T\partial _\alpha H_n\big (t, X_n^{{\hat{\alpha }}}, Y_n^{{\hat{\alpha _n}}}(t), {\hat{\alpha _n}}(t) \big )\eta (t)\textrm{d}t \Big ] + C_M\varepsilon _n^{1/2}, \end{aligned}$$

for a constant \(C_M>0\) depending on the constant M (introduced in the definition of \(\mathcal {A}\)). The inequality follows since \({\hat{\alpha _n\in }} \mathcal {A}\), and \(H_n\) is the Hamiltonian of the problem with drift \(b_n\) given by

$$\begin{aligned} H_n(t,x,y,a):= f(t, x,a) + b_n(t,x,a)y \end{aligned}$$

and \((Y^{{\hat{\alpha _n}}}_n, Z^{{\hat{\alpha _n}}}_n)\) the adjoint processes satisfying

$$\begin{aligned} \textrm{d}Y^{{\hat{\alpha _n}}}_n(t) = -\partial _xH_n(t, X_n^{{\hat{\alpha }}}, Y^{{\hat{\alpha _n}}}_n(t), {\hat{\alpha _n}}(t))\textrm{d}t + Z^{{\hat{\alpha _n}}}_n(t)\textrm{d}B(t). \end{aligned}$$

By standard arguments, we can thus conclude that

$$\begin{aligned} C_M\varepsilon _n^{1/2} +\partial _\alpha H_n(t, X_n^{{\hat{\alpha _n}}}(t), Y^{{\hat{\alpha _n}}}_n(t), {\hat{\alpha _n}}(t))\cdot (\beta - {\hat{\alpha _n}}(t)) \ge 0 \quad \mathbb {P}\otimes \textrm{d}t \mathrm {-a.s}. \end{aligned}$$

Recalling that \(b_{1,n}\) does not depend on \(\alpha \), this amounts to (recall definition of \(b_n\) in (7))

$$\begin{aligned}{} & {} C_M\varepsilon _n^{1/2} + \Big \{ \partial _{\alpha }f(t, X_n^{{\hat{\alpha _n}}}(t), {\hat{\alpha _n}}(t)) + \partial _{\alpha }b_{2}\big (t, X_n^{{\hat{\alpha _n}}}(t), {\hat{\alpha _n}}(t) \big )Y^{{\hat{\alpha _n}}}_n(t) \Big \}\cdot \\{} & {} (\beta - {\hat{\alpha _n}}(t)) \ge 0 \quad \mathbb {P}\otimes \textrm{d}t\text {-a.s.} \end{aligned}$$

We will now take the limit on both sides above as n goes to infinity. It follows by Lemma 2.3 and Lemma 2.6 respectively that \(X_n^{{\hat{\alpha _n}}}(t) \rightarrow X^{{\hat{\alpha }}}(t)\) and \(Y^{{\hat{\alpha _n}}}_n(t) \rightarrow Y^{{\hat{\alpha }}}(t)\) \(\mathbb {P}\)-a.s. for every \(t\in [0,T]\). Since \({\hat{\alpha _n\rightarrow }} \alpha \), we therefore conclude that

$$\begin{aligned} \Big \{ \partial _{\alpha }f(t, X^{{\hat{\alpha }}}(t), {\hat{\alpha }}(t)) + \partial _{\alpha }b_{2}\big (t, X^{{\hat{\alpha }}}(t), {\hat{\alpha }}(t) \big )Y^{{\hat{\alpha }}}(t) \Big \}\cdot (\beta - {\hat{\alpha }}(t)) \ge 0 \quad \mathbb {P}\otimes \textrm{d}t\text {-a.s.} \end{aligned}$$

This shows (4), which concludes the proof. \(\square \)

3 The Sufficient Condition for Optimality

3.1 Proof of Theorem 1.2

Let us now turn to the proof of the sufficient condition of optimality. Since we will need to preserve the concavity of H assumed in Theorem 1.2 after approximation, we specifically assume that the function \(b_n\) is defined by standard mollification. Therefore, \(H_n(t,x,y,a):= f(t,x,a)+ b_n(t,x,a)y\) is a mollification of H and thus remains concave.

Proof

(of Theorem 1.2) Let \({\hat{a}} \in \mathcal {A}\) satisfy (6) and \(\alpha '\) an arbitrary element of \(\mathcal {A}\). We would like to show that \(J({\hat{a}}) \ge J(\alpha ')\). Let \(n \in \mathbb {N}\) be arbitrarily chosen. By definition, we have

$$\begin{aligned}&J_n({\hat{a}}) - J_n(\alpha ')\\&\quad = \mathbb {E}\Big [g(X^{{\hat{a}}}_n(T)) {-} g(X^{\alpha '}_n(T)) {+} \int _0^Tf(u, X_n^{{\hat{a}}}(u), {\hat{a}}(u)) {-} f(u, X_n^{\alpha '}(u), \alpha '(u))\,\textrm{d}u \Big ] \\&\quad \ge \mathbb {E}\Big [\partial _xg(X^{{\hat{a}}}_n(T))\big \{X^{{\hat{a}}}(T) -X^{\alpha '}_n(T)\big \} + \int _0^T\big \{ b_n(u, X_n^{\alpha '}(u), \alpha '(u)) \\&\qquad - b_n(u, X_n^{{\hat{a}}}(u),{\hat{a}}(u))\big \} Y_n^{{\hat{a}}}(u)\,\textrm{d}u\\&\qquad + \int _0^T H_n(u, X_n^{{\hat{a}}}(u), Y_n^{{\hat{a}}}(u), {\hat{a}}(u)) - H_n(u, X_n^{\alpha '}(u),Y_n^{{\hat{a}}}(u), \alpha '(u))\,\textrm{d}u \Big ], \end{aligned}$$

where we used the definition of \(H_n\) and the fact that g is concave. Since \(Y_n^{{\hat{a}}}\) satisfies

$$\begin{aligned} Y^{{\hat{a}}}_n(t) = \mathbb {E}\Big [\Phi _n^{{\hat{a}}}(t,T) \partial _xg( X^{{\hat{a}}}_n(T)) + \int _t^T\Phi _n^{{\hat{a}}}(t,u) \partial _xf(u, X_n^{{\hat{a}}}(u), {\hat{a}}(u))\textrm{d}u\mid \mathcal {F}_t \Big ], \end{aligned}$$

it follows by the martingale representation and the Itô’s formula that there is a square integrable progressive process \((Y^{{\hat{a}}}_n,Z^{{\hat{a}}}_n)\) such that \(Y_n^{{\hat{a}}}\) satisfies the (linear) equation

$$\begin{aligned} Y^{{\hat{a}}}_n(t) = \partial _xg(X^{{\hat{a}}}_n) + \int _t^T\partial _xH_n(u, X^{{\hat{a}}}_n(u), Y_n^{{\hat{a}}}(u),{\hat{a}}(u))\,\textrm{d}u - \int _t^TZ_n^{{\hat{a}}}(u)\,\textrm{d}W(u). \end{aligned}$$

Recall that since \(b_n\) is smooth, so is \(H_n\). Therefore, by the Itô’s formula once again we have

$$\begin{aligned}&Y^{{\hat{a}}}_n(T)\big \{X_n^{{\hat{a}}}(T) {-} X_n^{\alpha '}(T)\big \} \!\!{=}\!\! \int _0^TY^{{\hat{a}}}_n(u)\big \{b_n(u, X^{{\hat{a}}}_n(u),{\hat{a}}(u)) {-} b_n(u, X^{\alpha '}_n(u),\alpha '(u)) \big \}\,\textrm{d}u\\&\quad {-} \int _0^T\big \{X^{{\hat{a}}}_n(u) {-} X^{\alpha '}_n(u) \big \}\partial _xH_n(u, X^{{\hat{a}}}_n(u), Y_n^{{\hat{a}}}(u),{\hat{a}}(u))\,\textrm{d}u \\&\quad {+} \int _0^T\big \{X^{{\hat{a}}}_n(u) {-} X^{\alpha '}_n(u) \big \} Z^{{\hat{a}}}_n(u)\,\textrm{d}W(u). \end{aligned}$$

Since the stochastic integral above is a local martingale, a standard localization argument allows to take expectation on both sides to get that

$$\begin{aligned} J_n({\hat{a}}) {-} J_n(\alpha ')&\ge \mathbb {E}\Big [{-} \int _0^T\big \{X^{{\hat{a}}}_n(u) {-} X^{\alpha '}_n(u) \big \}\partial _xH_n(u, X^{{\hat{a}}}_n(u), Y_n^{{\hat{a}}}(u),{\hat{a}}(u))\,\textrm{d}u \\&\quad {+} \!\!\int _0^T H_n(u, X_n^{{\hat{a}}}(u), Y_n^{{\hat{a}}}(u), {\hat{a}}(u)) {-} H_n(u, X_n^{\alpha '}(u),Y_n^{{\hat{a}}}(u), \alpha '(u))\,\textrm{d}u \Big ]\\&{\ge } \mathbb {E}\Big [\int _0^T \partial _\alpha H_n(u, X_n^{{\hat{a}}}(u), Y_n^{{\hat{a}}}(u), {\hat{a}}(u))\cdot ({\hat{a}}(u) {-} \alpha '(u))\,\textrm{d}u \Big ], \end{aligned}$$

where the latter inequality follows by concavity of \(H_n\).

Coming back to the expression of interest \(J({\hat{a}}) - J(\alpha ')\), we have

$$\begin{aligned} J({\hat{a}}) - J(\alpha ')&= J({\hat{a}}) - J_n({\hat{a}}) + J_n({\hat{a}}) - J_n(\alpha ') + J_n(\alpha ') - J(\alpha ')\\&\ge J({\hat{a}}) - J_n({\hat{a}}) + \mathbb {E}\Big [\int _0^T \partial _\alpha H_n(u, X_n^{{\hat{a}}}(u), Y_n^{{\hat{a}}}(u), {\hat{a}}(u))\cdot ({\hat{a}}(u) \\&\quad - \alpha '(u))\,\textrm{d}u \Big ] + J_n(\alpha ') - J(\alpha '). \end{aligned}$$

Since \(b_{1,n}\) does not depend on \(\alpha \), we have that \(\partial _\alpha H_n(u, X_n^{{\hat{a}}}(u), Y_n^{{\hat{a}}}(u), {\hat{a}}(u)) = \partial _\alpha b_{2}(u, X^{{\hat{a}}}_n(u),{\hat{a}}(u))Y^{{\hat{a}}}_n(u) + \partial _\alpha f(u, X^{{\hat{a}}}_n(u),{\hat{a}}(u))\). Therefore, taking the limit as n goes to infinity, it follows by Lemmas 2.3, 2.4 and 2.6 that it holds

$$\begin{aligned} J({\hat{a}}) - J(\alpha ') \ge \mathbb {E}\left[ \int _0^T \partial _\alpha H(u, X^{{\hat{a}}}(u), Y^{{\hat{a}}}(u), {\hat{a}}(u))\cdot ({\hat{a}}(u) - \alpha '(u))\,\textrm{d}u \right] . \end{aligned}$$

Since \({\hat{a}}\) satisfies (6), we therefore conclude that \(J({\hat{a}}) \ge J(\alpha ')\). \(\square \)

3.2 Example: Stochastic Predicted Miss Problem

Let us consider the following Stochastic predicted miss problem which first appeared in the works of Davis [15] and Gelder [51] as a conjecture of optimal laws for problems with finite fuel constraints. It was first rigorously solved by Beneš [8]. More precisely, let us consider the following optimal control problem, with the cost function given by

$$\begin{aligned} J(\alpha ):= \mathbb {E}\Big [ g(X^\alpha (T))\Big ], \end{aligned}$$

where the state process is the controlled SDE

$$\begin{aligned} \,\textrm{d}X^\alpha (t) {=} \Big ( b_1(X^\alpha (t)){+}b_2(t,X^\alpha (t))\alpha (t)\Big )\,\textrm{d}t {+}\sigma \,\textrm{d}B(t) ,\quad t \in [ 0,T] ,\quad X^\alpha (0) {=} x_0 \end{aligned}$$
(23)

and the control variable \(\alpha (t)\) takes values in \(\mathbb {A}=[-1,1]\). We are interested in the control problem

$$\begin{aligned} V(x_0) := \sup _{\alpha \in \mathcal {A}}J(\alpha ). \end{aligned}$$
(24)

As pointed earlier, this problem was first rigorously solved in [8] and then in [23] in the linear case. Further it was also considered in [4] in the case of Lipschitz coefficients. Our goal here is to provide an explicit (feedback) solution to this control problem when the function \(b_1\) is only measurable. Hence, we considered the following conditions:

  1. (A)

    The function \(g:\mathbb {R}\rightarrow \mathbb {R}\) is even, continuously differentiable and increasing on \(x > 0\) and concave. Moreover, it holds \(|g(x)|\le K(1+|x|^p)\) for all \(x \in \mathbb {R}\) and some \(p\ge 1\).

  2. (B)

    The function \(b_1:\mathbb {R}\rightarrow \mathbb {R}\) is odd and bounded; and the function \(b_2:[0,T]\times \mathbb {R}\rightarrow \mathbb {R}\) is odd, bounded, differentiable and Lipschitz-continuous in its second variable, uniform in the first one.

Proposition 3.1

If the conditions (A) and (B) are satisfied, then the control problem (24) is given in feedback form via

$$\begin{aligned} {\hat{\alpha }}(t) = \textrm{sgn}\Big (X^{{\hat{\alpha }}}(t)b_2(t, X^{{\hat{\alpha }}}(t))\Big ). \end{aligned}$$

Proof

Since the reward function g is even, maximizing it is similar to maximizing |X(T)|. Thus, we should take \(\alpha \) such that \(xb_2\alpha >0\). This can be seen applying Itô–Tanaka formula to |X(t)|. Thus, we make the Ansatz

$$\begin{aligned} \hat{\alpha }(t)={{\,\textrm{sgn}\,}}\Big (\hat{X}(t)b_2(t,\hat{X}(t))\Big ) \end{aligned}$$
(25)

with the notation \({\hat{X}}:= X^{{\hat{\alpha }}}\). We will use our maximum principle to show that this yields an optimal control.

In fact, by Theorem 1.2, it suffices to show that \({\hat{\alpha }}(t)\) maximizes the Hamiltonian \(H(t,{\hat{X}},Y,\alpha ):= \big (b_1({\hat{X}}(t)) + b_2(t, {\hat{X}}(t))\alpha \big )\cdot Y_t\), where Y denotes the adjoint process. Since the maximizer of this function is given by

$$\begin{aligned} {{\,\textrm{sgn}\,}}\Big (b_2(t,\hat{X}(t))Y(t)\Big ), \end{aligned}$$
(26)

it thus remains to show that for \({\hat{\alpha }}\) given by (25) it holds \({{\,\textrm{sgn}\,}}(Y(t)) = {{\,\textrm{sgn}\,}}({\hat{X}}(t))\).

Recall that the adjoint process takes the form

$$\begin{aligned} Y(t) = \mathbb {E}\Big [\Phi (t,T)\partial _xg({\hat{X}}(T)) \mid \mathcal {F}_t \Big ]. \end{aligned}$$
(27)

Since \(b_1\) is time-independent, the process \(\Phi \) takes the form

$$\begin{aligned} \Phi (t,T)=\exp \Big (-\int _{\mathbb {R}}b_1\left( z\right) L_T^{{\hat{X}}}(\textrm{d}z)+\int _t^{T}b'_2\left( u,{\hat{X}}(u)\right) {\hat{\alpha }}(u) \textrm{d}u\Big ), \end{aligned}$$

where \(L^{{\hat{X}}}_T\) denotes the local time of the process X at time T.

This follows by Theorem A.1 where, due to the Bouleau-Yor formula for semimartingales (see [48, Theorem 77, page 227] or [49, Exercise 1.28 page 236]) we can replace \(\int _0^T\int _{\mathbb {R}}b_1(z)L^{{\hat{X}}}(\,\textrm{d}u, \,\textrm{d}z)\) by \(\int _{\mathbb {R}}b_1(z)L^{{\hat{X}}}_T(\,\textrm{d}z)\). Next, let us define the function \({\widetilde{b}}_1\) by

$$\begin{aligned} {\widetilde{b}}_1(x):=\int _{-\infty }^xb_1(y)\textrm{d}y. \end{aligned}$$

Then \({\widetilde{b}}_1\) admits a bounded derivative \(b_1\) almost everywhere. Using again the Bouleau-Yor formula for continuous semimartingales we have

$$\begin{aligned} {\widetilde{b}}_1({\hat{X}}(T))&= {\widetilde{b}}_1({\hat{X}}(t))+\int _t^Tb_1({\hat{X}}(s))\textrm{d}{\hat{X}}(s) - \frac{1}{2} \int _{\mathbb {R}}b_1(z)L_T^{{\hat{X}}}(\textrm{d}z). \end{aligned}$$
(28)

Substituting the above into (27), yields

$$\begin{aligned} Y(t)&= \mathbb {E}\Big [\exp \Big \{2\Big ({\widetilde{b}}_1({\hat{X}}(T))-{\widetilde{b}}_1({\hat{X}}(t)) -\int _t^Tb_1({\hat{X}}(s))\sigma \textrm{d}B(s)\nonumber \\&\quad -\int _t^Tb^2_1({\hat{X}}(s))\textrm{d}s-\int _t^Tb_1({\hat{X}}(s))b_2(s,{\hat{X}}(s)){\hat{\alpha }}(s)\textrm{d}s\Big )\Big \}\nonumber \\&\quad \times \exp \Big \{\int _t^T\partial _xb_2(s, {\hat{X}}^{{\hat{\alpha }}}(s))\hat{\alpha }(s)\,\textrm{d}s\Big \}\partial _xg({\hat{X}}^{{\hat{\alpha }}}(T))\Big |\mathcal {F}_t \Big ]. \end{aligned}$$
(29)

Next we show that \({{\,\textrm{sgn}\,}}(Y(t)) = {{\,\textrm{sgn}\,}}({\hat{X}}(t))\). By (25) it holds that

$$\begin{aligned} \,\textrm{d}{\hat{X}}^{{\hat{\alpha }}} (t) {=} b_1({\hat{X}}(t)){+}|b_2(t,{\hat{X}}(t))|{{\,\textrm{sgn}\,}}({\hat{X}}(t))\,\textrm{d}t {+}\sigma \,\textrm{d}B(t) ,\quad t \in [ 0,T] ,\quad {\hat{X}}(0) {=} x_0. \end{aligned}$$
(30)

Now observe that the function \( x\mapsto b_1(x)+|b_2(t,x)|{{\,\textrm{sgn}\,}}(x)\) is odd and \({\tilde{B}}(t)=-B(t)\) is again a Brownian motion with the same law as B. Thus,

$$\begin{aligned} \,\textrm{d}( -{\hat{X}} (t) )&= - b_1({\hat{X}}(t))-|b_2(t,{\hat{X}}(t))|{{\,\textrm{sgn}\,}}({\hat{X}}(t))\,\textrm{d}t -\sigma \textrm{d}B(t) \nonumber \\&= b_1(-{\hat{X}}(t))+|b_2(t,-{\hat{X}}(t))|{{\,\textrm{sgn}\,}}(-{\hat{X}}(t))\,\textrm{d}t +\sigma \,\textrm{d}{\tilde{B}}(t), \end{aligned}$$
(31)

showing that \(-{\hat{X}}\) is a weak solution of the controlled SDE. By the weak uniqueness, it follows that \({\hat{X}}(s)\) and \(-{\hat{X}}(s)\) have the same distribution given \(s\ge \tau \) with

$$\begin{aligned} \tau :=\inf \{s\ge t,\,{\hat{X}}(s)=0\}. \end{aligned}$$

Now, we claim that

$$\begin{aligned} I_1&:=\mathbb {E}\Big [1_{\{\tau \le T\}} \exp \Big \{\int _t^T\partial _xb_2(u, {\hat{X}}(u))\{{{\,\textrm{sgn}\,}}(\hat{X}(u)b_2(u,\hat{X}(u)))\}\,\textrm{d}u\Big \}\partial _xg({\hat{X}}(T)) \\&\quad \times \exp \Big \{2\Big ({\tilde{b}}_1({\hat{X}}(T))-\tilde{b}_1(X(t))-\int _t^Tb_1({\hat{X}}(s))\textrm{d}B(s)-\int _t^Tb^2_1({\hat{X}}(s))\textrm{d}s\\&\quad -\int _t^Tb_1({\hat{X}}(s))b_2(s,{\hat{X}}(s))\alpha (s)\textrm{d}s\Big )\Big \}\Big |\mathcal {F}_t \Big ]=0. \end{aligned}$$

Indeed, by the weak uniqueness, we know that \(({\hat{X}},W)\) and \((-{\hat{X}},{\tilde{W}})\) have the same distribution.

Then using the facts that \(\partial _xg, b_2\) are odd and \({\widetilde{b}}_1\) is even, we obtain

$$\begin{aligned} I_1&=\mathbb {E}\Big [1_{\{\tau {\le } T\}} \exp \Big \{\!\!\int _t^T\!\!\partial _xb_2(u, {-}{\hat{X}}(u))\{{{\,\textrm{sgn}\,}}(-\hat{X}(u)b_2(u,{-}\hat{X}(u)))\}\,\textrm{d}u\Big \}\partial _xg({-}{\hat{X}}(T)) \\&\qquad \times \exp \Big \{2\Big ({\tilde{b}}_1(-{\hat{X}}(T))-{\tilde{b}}_1(-{\hat{X}}(t))-\int _t^Tb_1(-{\hat{X}}(s))\textrm{d}B(s)\\&\qquad -\int _t^Tb^2_1(-{\hat{X}}(s))\textrm{d}s-\int _t^Tb_1(-{\hat{X}}(s))b_2(s,-{\hat{X}}(s)){{\,\textrm{sgn}\,}}\\&\qquad \times \Big (-\hat{X}(s)b_2(s,-\hat{X}(s))\Big )\textrm{d}s\Big )\Big \}\Big |\mathcal {F}_t \Big ]\\&=\mathbb {E}\Big [1_{\{\tau \le T\}} \exp \Big \{\int _t^T\partial _xb_2(u, {\hat{X}}^{{\hat{\alpha }}}(u)){{\,\textrm{sgn}\,}}\Big (\hat{X}(u)b_2(u,\hat{X}(u))\Big )\,\textrm{d}u\Big \}({-}\partial _xg({\hat{X}}(T)) )\\&\qquad \times \exp \Big \{2\Big ({\tilde{b}}_1({\hat{X}}(T))-{\tilde{b}}_1({\hat{X}}(t))-\int _t^Tb_1({\hat{X}}(s))\textrm{d}{\tilde{B}}(s)\\&\qquad -\int _t^Tb^2_1({\hat{X}}(s))\textrm{d}s{-}\!\!\int _t^Tb_1({\hat{X}}(s))b_2(s,{\hat{X}}(s)){{\,\textrm{sgn}\,}}\Big (\hat{X}(s)b_2(s,\hat{X}(s))\Big )\textrm{d}s\Big )\Big \}\Big |\mathcal {F}_t \Big ] \\&=-\mathbb {E}\Big [1_{\{\tau \le T\}} \exp \Big \{\int _t^T\partial _xb_2(u, {\hat{X}}(u)){{\,\textrm{sgn}\,}}\Big (\hat{X}(u)b_2(u,\hat{X}(u))\Big )\,\textrm{d}u\Big \}\partial _xg({\hat{X}}(T)) \\&\qquad \times \exp \Big \{2\Big ({\tilde{b}}_1({\hat{X}}(T))-{\tilde{b}}_1(X(t))-\int _t^Tb_1({\hat{X}}(s))\textrm{d}{\tilde{B}}(s)\\&\qquad -\int _t^Tb^2_1({\hat{X}}(s))\textrm{d}s-\int _t^Tb_1({\hat{X}}(s))b_2(s,{\hat{X}}(s))\{{{\,\textrm{sgn}\,}}(\hat{X}(s)\\&\quad \times b_2(s,\hat{X}(s)))\}\textrm{d}s\Big )\Big \}\Big |\mathcal {F}_t \Big ]\\&= -I_1, \end{aligned}$$

where the latter equality follows from the fact that \(\tilde{B}(t)=-B(t)\) has the same law as B(t) as a process. Thus, we have \(2I_1=0\) and the claim follows.

Coming back to the adjoint process, we have

$$\begin{aligned} Y(t)&=\mathbb {E}\Big [1_{\{\tau > T\}} \exp \Big \{\int _t^T\partial _xb_2(u, {\hat{X}}(u)){{\,\textrm{sgn}\,}}\Big (\hat{X}(u)b_2(u,\hat{X}(u))\Big )\,\textrm{d}u\Big \}\partial _xg({\hat{X}}(T)) \\&\quad \times \exp \Big \{2\Big ({\tilde{b}}_1({\hat{X}}(T))-{\tilde{b}}_1({\hat{X}}(t))-\int _t^Tb_1({\hat{X}}(s))\textrm{d}B(s)-\int _t^Tb^2_1({\hat{X}}(s))\textrm{d}s\\&\quad -\int _t^Tb_1({\hat{X}}(s))b_2(s,{\hat{X}}(s))\alpha (s)\textrm{d}s\Big )\Big \}\Big |\mathcal {F}_t \Big ]. \end{aligned}$$

For \(T<\tau \), we have that \({{\,\textrm{sgn}\,}}({\hat{X}}(s))={{\,\textrm{sgn}\,}}({\hat{X}}(t))\) for all \(t\le s\le T\). If follows that the term inside the expectation is zero or has the same sign as \({\hat{X}}(t)\) (by properties of \(\partial _xg\)). Thus, it holds \({{\,\textrm{sgn}\,}}(Y(t))={{\,\textrm{sgn}\,}}({\hat{X}}(t))\), which yields the result. \(\square \)

4 Concluding Remarks

Let us conclude the paper by briefly discussing our assumptions. The condition \(b=b_1+b_2\) seems essential to derive existence and uniqueness results of the controlled system. For instance, the crucial bound (13) derived in [5, 37] is unknown when \(b_1\) depends on \(\alpha \). This condition is also vital in obtaining the explicit representation of the Sobolev derivative of the flows of the solution to the SDE in terms of its local time. This representation cannot be expected in multidimensions due to the non-commutativity of matrices and the local time. Therefore, much stronger (regularity) conditions are needed to derive the maximum principle in this case (see for example [1, 2, 4]). Note in addition that the boundedness assumption on b is made mostly to simplify the presentation. The results should also hold with b of linear growth in the spatial variable, albeit with more involved computations and with T small enough, since the flow in this case is expected to exist in small time.

Given the drift b, some known conditions on the control \(\alpha \) that guaranty existence and uniqueness of the strong solution to the SDE (1) satisfied by the controlled process are given in Example 2.1. These conditions involve the Malliavin derivative of \(\alpha \). Let us remark that the Malliavin differentiability of the control is not an uncommon assumption. This condition appears implicitly in the works [36, 40, 44] on the stochastic maximum principle where the coefficients are required to be at least two times differentiable with bounded derivatives.