1. Introduction
In many biological and economical control problems, the decision maker is faced with the situation where the information of the evolving system is not available all the time. Instead, the decision maker might observe the state of the system only at discrete times, for example daily or weekly. Thus, in the following we model the times when the controller receives the information of the evolving system as jump times of a Poisson process with a parameter $\lambda$. It is assumed that the decision maker can only exert control at these exogenously given times, in other words, they cannot act in the dark. Also, we restrict ourselves to controls of impulse type. Whenever control is applied, the decision maker has to pay a cost which is directly proportional to the size of the impulse. Otherwise, when there are no interventions, we assume that the system evolves according to one-dimensional linear diffusion X that is independent of the Poisson process. In the literature these types of restriction processes on the controllability of X are often referred to as constraints or signals, see [Reference Lempa24, Reference Lempa26, Reference Menaldi and Robin30, Reference Menaldi and Robin31, Reference Menaldi and Robin32].
In the classical case, the decision maker has continuous and complete information, and hence control is allowed whenever the decision maker wishes. The objective criterion to be minimized is often either a discounted cost or an ergodic cost (average cost per unit time). Both discounted cost and ergodic problems have been studied in the literature, but the ergodic problems have received less attention. This is because they are often mathematically more involved. However, from the point of view of many applications this is a bit surprising, as the discounting factor is often very hard or impossible to estimate. Also, outside of financial applications the discounting factor might not have a very clear interpretation.
The simplest case in the classical setting is where control is assumed to be costless. As a result, the optimal policy is often a local time of X at the corresponding boundaries; see [Reference Alvarez1, Reference Matomäki29] for discounted problems and [Reference Alvarez3, Reference Alvarez and Hening4, Reference Jack and Zervos19] for ergodic problems. One drawback of this model is that the optimal strategies are often singular with respect to Lebesgue measure, which makes them unappealing for applications. One way to make the model more realistic is to add a fixed transaction cost on the control. Then the optimal policy is often a sequential impulse control where the decision maker chooses a sequence of stopping times $\{ \tau_1, \tau_2, \ldots \}$ to exert control, and corresponding impulse sizes $\{ \zeta_1, \zeta_2, \ldots \}$; see [Reference Alvarez2, Reference Alvarez and Lempa5, Reference Harrison, Sellke and Taylor18]. In addition, it is possible that the flow of information is continuous but imperfect. This type of problem, often referred to as filtering problems, is also widely studied; see [Reference Bain and Crisan8, Reference Fleming14, Reference Picard33] for a textbook treatment and further references. In this case, the disturbance in the information flow is assumed to be such that the decision maker sees the underlying process all the time, but only observes a noisy version of it.
As in the model at hand, another possibility is to allow the decision maker to control only at certain discrete exogenously given times. These times can be, for example, multiples of integers, as in [Reference Kushner and Dupuis23, Reference Rogers35], or given by a signal process. Often, as in our model, the times between arrivals of the signal process are assumed to be exponentially distributed; see [Reference Lempa26, Reference Rogers and Zane36, Reference Wang39]. In [Reference Rogers and Zane36], this framework was used as a simple model for liquidity effects in a classical investment optimization problem. Reference [Reference Wang39] investigates both discounted cost and an ergodic cost criterion while tracking a Brownian motion under quadratic cost, and [Reference Lempa26] generalizes the discounted problem to a more general payoff and underlying structure. Related studies in optimal stopping are [Reference Dupuis and Wang12, Reference Lempa24]. In [Reference Dupuis and Wang12], the authors consider a perpetual American call with underlying geometric Brownian motion, and in [Reference Lempa24] the results are generalized to a larger class of underlying processes. Studies related to more general signal processes are found in [Reference Menaldi and Robin30, Reference Menaldi and Robin31, Reference Menaldi and Robin32]. In these, the signal process can be a general, not necessarily independent, renewal process, and the underlying process is a general Markov–Feller process. There are also multiple studies that are less directly related, where an underlying Poisson process brings a different friction to the model by either affecting the structure of the underlying diffusion [Reference Guo and Zhang17, Reference Jiang and Pistorius20] or the payoff structure [Reference Alvarez and Stenbacka6, Reference Lempa25, Reference Lempa27].
The main contribution of this paper is that we allow the underlying stochastic process X to follow a general one-dimensional diffusion process, and also allow a rather general cost function. This is a substantial generalization of [Reference Wang39], where the case of Brownian motion with quadratic cost is considered. We emphasize this in the illustrations in Section 5 by explicitly solving multiple examples with different underlying dynamics and cost functions. These generalizations have not, to the best of our knowledge, been studied before in the literature. Furthermore, we are able to connect the problem to a related problem in optimal ergodic singular control [Reference Alvarez3].
The rest of the paper is organized as follows. In Section 2 we define the control problem and prove auxiliary results. In Section 3, we first investigate the necessary conditions of optimality by forming the associated free boundary problem, followed by the verification. We connect the problem to a similar problem of singular control in Section 4, and then illustrate the results by explicitly solving a few examples in Section 5. Finally, Section 6 concludes our study.
2. The control problem
2.1. The underlying dynamics
Let $(\Omega, \mathcal{F}, \{\mathcal{F}_t\}_{t \geq 0}, \mathbb{P})$ be a filtered probability space which satisfies the usual conditions. We consider an uncontrolled process X defined on $(\Omega, \mathcal{F}, \{\mathcal{F}_t\}_{t \geq 0}, \mathbb{P})$ that evolves in $\mathbb{R}_+$ and is modelled as a solution to regular linear Itô diffusion
where $W_t$ is the Wiener process and the functions $\mu,\sigma\,{:}\, (0,\infty) \to \mathbb{R}$ are continuous and satisfy the condition $\int_{x-\varepsilon}^{x+\varepsilon} \frac{1+|\mu(y)|}{\sigma^2(y)}{\rm d} y < \infty$. These assumptions guarantee that the diffusion has a unique weak solution (see [Reference Karatzas and Shreve22, Section 5.5]). Although we consider the case where the process evolves in $\mathbb{R}_+$, we remark that this is done only for notational convenience, and the results would remain the same, with obvious changes, if the state space were replaced with any interval of $\mathbb{R}$.
We define the second-order linear differential operator $\mathcal{A}$ that represents the infinitesimal generator of the diffusion X as
and for a given $\lambda > 0$ we respectively denote the increasing and decreasing solutions to the differential equation $(\mathcal{A}-\lambda)\,f=0$ by $\psi_{\lambda} > 0$ and $\varphi_{\lambda} > 0$.
The differential operator $\lambda-\mathcal{A}$ has an inverse operator called the resolvent $R_{\lambda}$ defined by
for all $x \in \mathbb{R}_+$ and functions $f \in \mathcal{L}_1^{\lambda}$, where $\mathcal{L}_1^{\lambda}$ is the set of functions f on $\mathbb{R}_+$ which satisfy the integrability condition $\mathbb{E}_x [ \int_0^{\tau}{\rm e}^{-\lambda s} |f(X_s)| {\rm d} s ] < \infty $. Here, $\tau$ is the first exit time from $\mathbb{R}_+$, i.e. $\tau = \inf \{ t \geq 0 \mid X_t \not \in \mathbb{R}_+ \}$. We also define the scale density of the diffusion by
which is the (non-constant) solution to the differential equation $\mathcal{A}\,f=0$, and the speed measure of the diffusion by
It is well known that the resolvent and the solutions $\psi_{\lambda}$ and $\varphi_{\lambda}$ are connected by the formula
where
denotes the Wronskian determinant (see [Reference Borodin and Salminen10, p. 19]). We remark that the value of $B_{\lambda}$ does not depend on the state variable x because an application of the harmonicity properties of $\psi_\lambda$ and $\varphi_\lambda$ yields
In calculations, it is sometimes also useful to use the identity
2.2. The control problem
We consider a control problem where the goal is to minimize the average cost per unit time, so that the controller is only allowed to control the underlying process at exogenously given times. These times are given as the arrival times of an independent Poisson process, called the signal process or constraint, and thus the interarrival times are exponentially distributed.
Assumption 1. The Poisson process $N_t$ and the controlled process $X_t$ are assumed to be independent, and the process $N_t$ is $\{\mathcal{F}_t\}_{t \geq 0}$-adapted.
More precisely, the set of admissible controls $\mathcal{Z}$ is given by those non-decreasing left-continuous processes $\zeta_{t \geq 0}$ that have the representation
where N is the signal process and the integrand $\eta$ is $\{\mathcal{F}_t\}_{t \geq 0}$-predictable. The controlled dynamics are then given by the Itô integral
where $\tau_0^\zeta$ is the first exit time of $X^{\zeta}_t$ from $\mathbb{R}_+$, i.e. $\tau_0^\zeta = \inf \{ t \geq 0 \mid X^{\zeta}_t \not \in \mathbb{R}_+ \}$.
Define the average cost per unit time or ergodic cost criterion as
where $\gamma$ is a given positive constant and $\pi\,{:}\,\mathbb{R}_+ \to \mathbb{R}$ is a function measuring the cost of continuing the process. Now define the value function
and denote by $\beta$ the minimum average cost. The objective of the control problem is to minimize $J(x, \zeta)$ over all the admissible controls $\zeta \in \mathcal{Z}$ and to find, if possible, the optimal control $\zeta^*$ such that $\beta = \inf_{\zeta \in \mathcal{Z}} J(x,\zeta) = J(x,\zeta^*)$.
We now define the auxiliary functions $\pi_{\gamma}\,{:}\,\mathbb{R}_+ \to \mathbb{R}$,
and $\pi_{\mu}\,{:}\,\mathbb{R}_+ \to \mathbb{R}$,
In order for our solution to be well behaved, we must pose the following assumptions:
Assumption 2. We assume that
(i) the lower boundary 0 and upper boundary $\infty$ are natural;
-
(ii) the cost $\pi$ is continuous, non-negative, and minimized at 0;
-
(iii) the function $\pi_{\mu}$ and $\mathrm{id}\,{:}\,x \mapsto x$ are in $\mathcal{L}_1^{\lambda}$;
-
(iv) there exists a unique state $x^* \in \mathbb{R}_+$ such that $\pi_{\mu}$ is decreasing on $(0, x^*)$ and increasing on $[x^*, \infty)$. Also, $\lim_{x \uparrow \infty} \pi_{\mu}(x) > 0$.
The boundaries of the state space are assumed to be natural, which means that, in the absence of interventions, the process cannot become infinitely large or infinitely close to zero in finite time. In biological applications these boundary conditions guarantee that the population does not explode or become extinct in the absence of harvesting. We refer to [Reference Borodin and Salminen10, pp. 18--20] for a more thorough discussion of the boundary behaviour of one-dimensional diffusions. Also, it is worth mentioning that no second-order properties of $\pi_{\mu}$ are assumed.
In addition, the following limiting and integrability conditions on the scale density and speed measure must be satisfied. These conditions assure the existence of a stationary distribution of the underlying diffusion.
Assumption 3. We assume that
(i) $m(0,y)=\int_{0}^{y} m'(z){\rm d} z <\infty$ and $\int_{0}^{y} \pi_{\mu}(z) m'(z){\rm d} z<\infty$ for all $y \in \mathbb{R}_+$;
-
(ii) $\lim_{x \downarrow 0} S'(x) = \infty$.
Remark 1. The conditions of Assumption 3 alone guarantee that the lower boundary 0 should be either natural or entrance, and hence unattainable. However, in the proof of Lemma 1 we must exclude the possibility of entrance to assure that L(x) (defined below) also attains negative values. If we wanted to include this possibility, we would also have to assume that $\lim_{x \to 0} \pi_{\mu}(x) = \infty$; see the proof of Lemma 1.
2.3. Auxiliary results
Define the auxiliary functions $L\,{:}\, \mathbb{R_+} \to \mathbb{R}$ and $H\,{:}\, \mathbb{R_+} \to \mathbb{R}$ as
These functions will offer a convenient representation of the optimality equation in Section 3, and thus their properties play a key role when determining the optimal control policy.
Lemma 1. Under Assumption 1, the functions L(x) and H(0,x) aer such that there exists a unique $\tilde{x} < x^*$ and a unique $\hat{x} > x^*$ such that
(i) $L(x) \, \substack{< \\ = \\ >} \, 0$ when $x \, \substack{< \\ = \\ >} \, \tilde{x}$,
-
(ii) $H(0,x) \, \substack{< \\ = \\ >} \, 0$ when $x \, \substack{> \\ = \\ <} \, \hat{x}$.
Proof. The proof of the claim on L is similar to that of [Reference Lempa26, Lemma 3.3]. However, to show that the results are in accordance with Remark 1, we need to adjust the argument on finding a point $x_1 < x^*$ such that $L(x_1)<0$. Thus, assume for a while that $\lim_{x \to 0} \pi_{\mu}(x) = \infty$, and that $x^* > y > x$. Then
which shows that $\lim_{x \to 0} L(x) = - \infty$.
To prove the second part, assume first that $y>x>x^*$. Since the function $\pi_\mu$ is increasing on $(x^*,\infty)$, we see that
proving that H is decreasing on $(x^*, \infty)$. It also follows from
that $\lim_{y \to \infty} H(0,y) < 0$. Next, assume that $x^* > y > x$. Because $\pi_\mu$ is decreasing on $(0, x^*)$, we find similarly that
implying that H is increasing on $(0, x^*)$. Furthermore, H is positive when $x < x^*$. Hence, by continuity, H has a unique root, which we denote by $\hat{x}$.□
Proposition 1. There exists a unique solution $\bar{x} \in (\tilde{x}, \hat{x})$ to the equation
Proof. Define the function
Assuming that $x_1 > \hat{x} > x^*$, we get, by Lemma 1, that
Similarly, when $x_2 < \tilde{x} < x^*$ we have that
By continuity, the function P(x) must have at least one root. We denote one of these roots by z.
To prove that the root z is unique, we first notice that the naturality of the upper boundary implies that [Reference Borodin and Salminen10, p. 19]
Hence,
Thus, we see that the equation $P(x)=0$ is equivalent to
Now, differentiating the left-hand side yields
where $I(x) = \int_{x}^{\infty} \varphi_{\lambda}(y)m'(y){\rm d} y$. Differentiating the right-hand side and evaluating it at z, we get, using the equation $P(z)=0$,
Because $L(y)>0$ in the region $(\tilde{x}, \hat{x})$, and all the other terms are positive everywhere, by comparing the derivatives we find that
Therefore, by continuity, the intersection between the curves
and
is unique. This unique point is denoted by $\bar{x}$. □
In the next lemma we make some further computations that are needed for the sufficient conditions of the control problem. Define the functions $J\,{:}\,\mathbb{R}_+ \to \mathbb{R}$ and $I\,{:}\,\mathbb{R}_+ \to \mathbb{R}$ as
Lemma 2. Under Assumption 2,
(i) $J'(x) \substack{> \\ = \\ <} 0$ when $x \substack{> \\ = \\ <} \tilde{x}$,
(ii) $I'(x) \substack{> \\ = \\ <} 0$ when $x \substack{> \\ = \\ <} \hat{x}$.
Here, $\tilde{x}$ and $\hat{x}$ are as in Lemma 1.
Proof. The first claim follows from the formula
which can be derived using representation (1) and straightforward differentiation (see [Reference Lempa26, Lemma 3] for details). The claim on I follows similarly, as differentiation yields
□
3. The solution
3.1. Necessary conditions
Denote the candidate solution for (3) as F(T, x). We use the heuristic that F(T, x) can be separated for large T as
In mathematical finance literature, the constant $\beta$ usually denotes the minimum average cost per unit time and W(x) is the potential cost function (see [Reference Fleming15, Reference Sethi, Zhang and Zhang37]). The fact that the leading term $\beta T$ is independent of x is, of course, dependent on the ergodic properties of the underlying process. We also note that this heuristic can be used as a separation of variables to solve a partial differential equation of parabolic type related to the expectation in (3) via the Feynman–Kac formula [Reference Fleming and McEneaey16].
We shall proceed as in [Reference Lempa26]. We assume that the optimal control policy exists and is given by the following. When the process is below some threshold $y^*$ (called the waiting region) we let the process run, but if the process is above the threshold value $y^*$ (called the action region) and the Poisson process jumps we exert the exact amount of control to push the process back to the boundary $y^*$ and start it anew.
In the waiting region $[0,y^*]$ we expect that the candidate solution satisfies Bellman’s principle,
where U is an exponentially distributed random variable with mean $1/\lambda$.
Using the heuristic (6) and noticing the connection between the random times U and the resolvent, we get, by independence and the strong Markov property, that
Hence, we arrive at the equation
We next choose $f(x) = W(x) - \lim_{r \to 0}(R_r \pi)(x)$ in [Reference Lempa24, Lemma 2.1], and notice that the lemma remains unchanged even if we add a constant $\beta / \lambda$. Thus, we expect, by our heuristic arguments, that the pair $(W, \beta)$ satisfies the differential equation
This type of equation often arises in ergodic control problems and there is lots of literature on sufficient conditions for the existence of a solution [Reference Arapostathis, Borkar and Ghosh7, Reference Fleming15, Reference Robin34]. We remark here that usually these conditions rely heavily on the solution of the corresponding discounted infinite-horizon control problem, and thus apply the so-called vanishing discount method. However, in our case we can proceed by explicit calculations.
Next, we shall determine the equation for the pair $(W, \beta)$ in the action region $[y^*, \infty]$. The Poisson process jumps in infinitesimal time with probability $\lambda {\rm d} t$, and in that case the agent has to pay a cost $\gamma (x- y^*) + F(T, y^*)$. On the other hand, the Poisson process does not jump with probability $1- \lambda {\rm d} t$, and in this case the agent has to pay $\pi(x) {\rm d} t + \mathbb{E}_x\left[ F(T-{\rm d} t,X_{{\rm d} t}) \right]$. Thus, the candidate function F should satisfy the condition
Now, again using the heuristic (6) and that, intuitively, ${\rm d} t^2=0$, we find that
By formally using Dynkin’s formula on the last term and simplifying we get
We conclude that in the action region, the pair $(W, \beta)$ should satisfy the differential equation
Now, we first observe that
which implies that W(x) satisfies the $C^1$-condition $W'(y^*)=\gamma$. We have thus arrived at the following free boundary problem: find a function W(x) and constants $y^*$ and $\beta$ such that
Remark 2. Another common approach to heuristically form the Hamilton–Jacobi–Bellman (HJB) equation of the problem is to use the value function $J_r(x)$ of the corresponding discounted problem (see [Reference Lempa26, p. 4]) and the vanishing discount limits $rJ_r(\bar{x}) \to \beta$ and $W_r(x) = J_r(x)-J_r(\bar{x}) \to W(x)$, where $\bar{x}$ is a fixed point in $\mathbb{R}_+$ (see [Reference Robin34, p. 284] and [Reference Fleming15, p. 427]). This argument yields the HJB equation
where $S = [y^*,\infty)$ is the control region.
To solve the free boundary problem, we consider (8) first. In this case we write the differential operator $\mathcal{A}$ as [Reference Revuz and Yor38, p. 285]
which allows us to find that
Therefore, integrating over the interval $(0, y^*)$ gives
Hence, by Assumption 3 and the $C^1$-condition $W'(y^*)=\gamma$, we get
Finally, using (2), we arrive at
Next, we consider (7). We immediately find that a particular solution is
Hence, we conjecture analogously to [Reference Lempa26, p. 113] that the solution to (7) is
To find the constants C and $\beta$, we first use the continuity of W at the boundary $y^*$, which allows us to substitute $x=y^*$ into (11). This yields
Then, by applying the condition $W'(y^*) = \gamma$ to (11), we find that
Combining this with (12) gives
To rewrite this expression, we first notice that a straightforward differentiation gives
Thus, by the fundamental theorem of calculus and the naturality of the upper boundary, we get
Next, using (1), we find that
Combining these observations with (5), the constant $\beta$ reads
Finally, by recalling the definition of $\pi_{\mu}(x)$, we have
Now, by equating the representations (10) and (14) of $\beta$, we find the optimality condition
which can be re-expressed, using the functions L(x) and H(0, x), as
We proved in Proposition 1 that there exists a unique solution $\bar{x}$ to (15), and thus we will assume in the following that $y^*=\bar{x}$.
Remark 3. As often in ergodic optimal control problems, the potential value function W(x) satisfies second-order differentiability across the boundary $\lim_{x \downarrow y^*} W''(x)= \lim_{x \uparrow y^*} W''(x)$. This can be verified as follows. When $x>y^*$, the differentiation and the harmonicity properties $(R_{\lambda} (\mathcal{A-\lambda})\pi_{\gamma})(x)+\pi_{\gamma}(x) = 0$ and $(\mathcal{A}-\lambda)\varphi_{\lambda}(x) = 0$ yield
which after cancellation and applying (13) and (1) equals
Therefore, by using (5) and (14) we find that
On the other hand, when $x < y^*$ we notice that
Hence, by differentiating the left-hand side and plugging in $x = y^*$, we find by the first-order condition $W'(y^*) = \gamma$ that
3.2. Sufficient conditions
We begin with an initial remark. When $x > y^*$, we get, by differentiating (11) and using Lemma 2, that
In the opposite case, when $x < y^*$, we have
Thus, by integrating over the interval (0, x), and using (10) and Lemma 2, we find that
These observations imply that, under the standing assumptions, the function $W(x)-\gamma x$ has a global minimum at $y^*$, which shows that W(x) satisfies the variational equality
Proposition 2. (Verification.) Under Assumptions 1, 2, and 3, the optimal policy is as follows. If the controlled process $X^{\zeta}$ is above the threshold $y^*$ at a jump time of N, i.e. $X^{\zeta}_{T_-} > y^*$, the decision maker should take the controlled process $X^{\zeta}$ to $y^*$. Further, the threshold $y^*$ is uniquely determined by (15), and the constant $\beta$ characterized by (10) and (14) gives the minimum average cost,
Proof. Define the function
and a family of almost surely finite stopping times $\tau(\rho)_{\rho > 0}$ as $\tau(\rho)\coloneqq \tau_0^\zeta \wedge \rho \wedge \tau^\zeta_{\rho}$, where $\tau_{\rho}^\zeta = \inf \{t \geq 0\,{:}\, X_t^\zeta \not \in (1/\rho, \rho) \}$. By applying the Doléans–Dade–Meyer change of variables formula to the process $W(X_t)$, we obtain
Because the control $\zeta$ jumps only if the Poisson process N jumps, we have that $W(X_s^{\zeta}) - W(X_{s-}^{\zeta}) + \gamma(\Delta \zeta_s) \geq \Phi(X_{s-}^{\zeta})$. By combining these two observations with (16), we get
where
Here, $\tilde{N}_t = (N_t-\lambda t)_{t \geq 0}$ is the compensated Poisson process. It follows from the calculation above that $Z_{t \wedge \tau(\rho)} + M_{t \wedge \tau(\rho)}$ is a submartingale and thus $\mathbb{E}_x [ Z_{t \wedge \tau(\rho)} + M_{t \wedge \tau(\rho)} ] \geq 0$. Taking expectation on both sides, dividing by ${t \wedge \tau(\rho)}$, and letting $t, \rho \to \infty$, we find that
Thus, if $\liminf_{T\to \infty} \frac{1}{T} \mathbb{E}_x [ W(X_T^{\zeta})] = 0$, it follows that $J(x,\zeta) \geq \beta$; we postpone the proof of this limiting property to the following lemma.
Next, we prove that ${J(x,\zeta^*) \leq \beta}$. We proceed as above and note that (17) holds as equality when $\zeta = \zeta^*$. Hence, the local martingale term $M_T+Z_T$ is now uniformly bounded from below by $-W(x)-\beta T$, and is therefore a supermartingale. Thus, we have
Finally, dividing by T and letting $T \to \infty$, we get $J(x,\zeta^*) \leq \beta$, which completes the proof. □
As is usual in ergodic control problems, we noticed in the proof that the verification theorem holds under the assumption that $\liminf_{T\to \infty} \frac{1}{T} \mathbb{E}_x [ W(X_T^{\zeta})] = 0$. Thus, in the following lemma we give the sufficient condition on $\pi(x)$ under which the limit equals zero.
Lemma 3. The limit
holds if $\pi(x) \geq C (x^{\alpha}-1)$, where $\alpha$ and C are positive constants.
Proof. Let $x > y^*$. Then W(x) reads
Because $\varphi_{\lambda}(x)$ is bounded in this region, we only need to deal with the resolvent term. By the Markov property and the substitution $k=s+T$, we find that
By l’Hôpital’s rule and the assumption that id and $\pi$ are elements of $\mathcal{L}^{\lambda}_1$, we find that
On the other hand, if
there exists $T_1$ such that
for all $s > T_1$. Together with the assumption $\pi(x) \geq C (x^{\alpha}-1)$, this leads to contradiction as
Similarly, we must have
and thus conclude that the limit (18) must vanish.
In the opposite case $x < y^*$, we find by integrating in (9) that
Multiplying by S $^{\prime}(x)$ and integrating over the interval $(x,y^*)$, we have
Since the second term is negative and $\pi(x)$ is positive everywhere, this has the upper bound
As the last integral is positive, we can by Assumption 3 expand the region of the inner integral to get
Thus,
Consequently, the upper bound is of the form $W(x) \leq C_0 S(x) + C_1$, where $C_0$ and $C_1$ are constants. Since $S(X_t)$ is a non-negative local martingale [Reference Bass9, p. 88] and hence a supermartingale, we have
Hence, also in this case the limit (18) must vanish. □
Remark 4. Another approach to see that the limit vanishes is to get a suitable upper bound for W(x). Indeed, if $W(x) \leq A_0 + A_1\pi(x)$ for some constant $A_0$ and $A_1$, then the result also holds by [Reference Wang39, Lemma 3.1].
4. Ergodic singular control problem: Connecting the problems
The singular control problem, where the agent is allowed to control the process $X_t$ without any constraints, is studied in the case of Brownian motion in [Reference Karatzas21] and in the case of a more general one-dimensional diffusion in [Reference Alvarez3, Reference Jack and Zervos19]. In this corresponding singular problem the optimal policy is a local time reflecting barrier policy. The threshold $y^*_s$ characterizing the optimal policy is the unique solution to the optimality condition [Reference Alvarez3, p. 17]
Heuristically, one would expect that in the limit $\lambda \to \infty$, this optimal boundary $y^*_s$ coincides with the optimal boundary $y^*$. This is because, in the limit, the decision maker has more frequent opportunities to exercise control. This is shown in the next proposition after an auxiliary lemma.
Lemma 4. Let $\varphi_{\lambda}(x)$ be the decreasing solution to the differential equation $(\mathcal{A}-\lambda)\,f=0$, and assume that $x < z$; then
Proof. Taking the limit $\lambda \to \infty$ in [Reference Borodin and Salminen10, p. 18]
where $\tau_z = \inf \{ t \geq 0 \mid X_t = z \}$ is the first hitting time to z, yields the result by monotone convergence. □
We now have the following result.
Proposition 3. Define a function $G\,{:}\,\mathbb{R}_+\to \mathbb{R}$ as
Let $y^*$ and $y^*_s$ be the unique solutions to $\hat{G}(x)=0$ and $H(0,x) = 0$, respectively. Then $\hat{G}(y^*_s) \to 0$ as $\lambda \to \infty$.
Proof. Because $H(0,y^*_s)=0$ and the upper boundary $\infty$ is natural, we have
Thus, taking the limit $\lambda \to \infty$ yields the result by Lemma 4. □
It is also reasonable to expect that when $\lambda$ increases, it is more likely that the decision maker postpones the exercise of control as they have more information about the underlying process available. Therefore, we expect that the optimal threshold $y^*$ is increasing as a function of $\lambda$. The next proposition shows that this is indeed the case.
Proposition 4. Assume that $\mu(x)>0$. Then the unique root $y^*_{\lambda}$ of the function
is increasing in $\lambda$.
Proof. Let $\hat{\lambda}>\lambda$. From the proof of Lemma 4, we find that, for every $x<z$,
which is equivalent to
Because $\hat{\lambda}>\lambda$, there exists $r >0$ such that $\hat{\lambda} = \lambda + r$. Thus, utilizing the fact that $(\mathcal{A}-\lambda)\varphi_{\lambda+r}= (\mathcal{A}-(\lambda+r))\varphi_{\lambda+r} + r\varphi_{\lambda+r} = r\varphi_{\lambda+r}$ with [Reference Alvarez2, Corollary 3.2], we have
Reorganizing the above, we get
Combining (20) and (21), we deduce that
Hence, the function $G_{\lambda}(x)$ satisfies
This implies that $y^*_{\lambda} \leq y^*_{\hat{\lambda}}$ as, by (4), $G_{\lambda}(x)$ is positive in the interval $(0,y^*_{\lambda})$ and has a unique root. □
Remark 5. The assumption that $\mu(x)>0$ is somewhat restricting, and is there to guarantee that $\pi_{\mu}(x)>0$. It would be enough that
It is often hard to show this exactly; however, in applications it can be verified numerically.
Remark 6. Denote by $\beta_s$ the average cost per unit time of the singular problem. Then $\beta \xrightarrow{\lambda \to \infty} \beta_s$ [Reference Lempa and Saarinen28, p. 12].
5. Illustrations
5.1. Verhulst–Pearl diffusion
We consider a standard Verhulst–Pearl diffusion
where $\mu >0$, $\sigma >0$, and $\beta > 0$. This diffusion is often used as a model for stochasticly fluctuating populations [Reference Alvarez and Hening4, Reference Evans, Hening and Schreiber13]. The scale density and speed measure in this case are
We assume that the cost $\pi(x) = x^2$ and $\gamma=1$. Hence, $\pi_{\mu}(x) = x^2 - x \mu (1-\beta x) $. In this setting, we note that if $\mu > \sigma^2/2$ then
The minimal excessive functions read [Reference Dayanik and Karatzas11, pp. 201–203]
where U and M are Kummer’s confluent hypergeometric functions of the second and first kind respectively, and
We see that our assumptions are satisfied, and thus the result applies. Unfortunately, (15) for the optimal threshold $y^*$, and the formula for the minimum average cost $\beta$ (14), are complicated and therefore left unstated. However, we can illustrate the results numerically. In Table 1 the optimal threshold $y^*$ is calculated with the values $\mu = 1$, $\sigma=1$, and $\beta = 0.01$ for a few different values of $\lambda$. We see from the table that as $\lambda$ increases the threshold $y^*$ gets closer to the corresponding threshold $y^*_s \approx 0.743$ of the singular control problem (19).
5.2. The standard Ornstein–Uhlenbeck process
As remarked in the introduction, the results also hold for $\mathbb{R}$ with straightforward changes. Indeed, we only have to adjust the assumptions slightly, by changing the lower boundary from 0 to $-\infty$ in Assumptions 2 and 3, and change all the formulas accordingly. With this change we can study a larger class of processes.
Consider dynamics that are characterized by the stochastic differential equation
where $\beta > 0$. This diffusion is often used to model continuous-time systems that have mean-reverting behaviour. To illustrate the results we choose the running cost $\pi(x) = |x|,$ and consequently $\pi_{\mu}(x) = |x|-\gamma \beta x$. The scale density and the density of speed measure in this case are
and the minimal excessive functions read [Reference Borodin and Salminen10, pp. 141]
where $D_{\nu}(x)$ is a parabolic cylinder function. Equation (15) for the optimal threshold again takes a rather complicated form and thus the results are only illustrated numerically in Table 2. In the singular control case (19) gives $y^*_s \approx 0.535$. Thus, as expected, the threshold value $y^*$ gets closer to $y^*_s$ when $\lambda$ increases.
6. Conclusions
We have considered ergodic singular control problems with the constraint of a regular one-dimensional diffusion. Relying on basic results from the classical theory of linear diffusions, we characterized the state after which the decision maker should apply an impulse control to the process. Our results are in agreement with the findings of [Reference Alvarez3], where the corresponding unconstrained singular control problem is studied. Indeed, no second-order or symmetry properties of the cost are needed. In addition, we proved that as the decision maker gets more frequent chances to exercise control, the value of the problem converges to that of the singular problem.
There are a few directions in which the constrained problems could be studied further. To the best of our knowledge, the finite-horizon problem with constraint remains open, even for the case of Brownian motion. Thus, it would be interesting if a similar analysis to [Reference Karatzas21] could be extended to also cover this case. In this case, we would expect similar connections between the finite-horizon time and the present problem as for those without any constraints [Reference Karatzas21, p. 241].
Moreover, the related two-sided problem, where the decision maker could control both downwards and upwards, but only at jump times of a Poisson process, could be studied. Unfortunately, these extension are outside the scope of the present study, and are therefore left for future research.
Acknowledgement
Two anonymous referees are acknowledged for helpful comments. We would like to gratefully acknowledge the emmy network foundation under the aegis of the Fondation de Luxembourg for its continued support.