June 18, 2024
In a continuous-time economy, this paper formulates the Epstein-Zin preference for discounted dividends received by an investor as an Epstein-Zin singular control utility. We introduce a backward stochastic differential equation with an aggregator integrated with respect to a singular control, prove its well-posedness, and show that it coincides with the Epstein-Zin singular control utility. We then establish that this formulation is equivalent to a robust dividend policy chosen by the firm’s executive under the Maenhout’s ambiguity-averse preference. In particular, the robust dividend policy takes the form of a threshold strategy on the firm’s surplus process, where the threshold level is characterized as the free boundary of a Hamilton–Jacobi–Bellman variational inequality. Therefore, dividend-caring investors can choose firms that match their preferences by examining stock’s dividend policies and financial statements, whereas executives can make use of dividend to signal their confidence, in the form of ambiguity aversion, on realizing the earnings implied by their financial statements.
Dividend (or cash payout) policy is an important topic in both corporate finance and asset pricing. The Miller-Modigliani [1] dividend irrelevance theory suggests that a firm’s dividend policy does not affect its value and stock prices in a perfect market. However, the underlying incentives for stockholders to receive dividends and for the firm’s executives to pay dividends are largely unknown in reality. For instance, a stockholder can liquidate some of his stocks for consumption instead of receiving cash dividend, which is subject to a higher tax rate in the USA. Hence, Black [2] proposes the dividend puzzle which has generated a great of attentions in the literature.
A school of thought is the dividend signaling theory (DST) (see Bhattacharya [3], John and Williams [4], Miller and Rock [5]) that a firm’s executive passes signals about the firm’s information to the public through dividends. In particular, Black [6] wrote,
“The idea that dividends convey information beyond that conveyed by the firm’s financial statements and public announcements stretches the imagination.… I think we must assume that investors care about dividends directly. We must put dividends into the utility functions.”
Although most empirical studies support DST, the survey in Brav et al. [7] shows that executives viewed the mechanism behind DST as broadly misguided. However, dividend changes do contain information about future earnings according to Ham et al. [8]. Additionally, Michaely et al. [9] shows that payout announcements signal changes in cash flow volatility (in the opposite direction) but not stock volatility, which the authors refer to as “signaling safety”. A firm’s cashflow is related to its earnings so that the cashflow volatility indicates the earnings’ volatility. By modeling the firm’s uncontrolled surplus process as diffusion models with a stochastic drift, there is a mathematical toolkit as in Reppen et al. [10] to compute the optimal dividend policy under a random profitability.
In this paper, we consider the suggestions by Black [6]. We assume that the firm’s financial reports contain information on the firm’s expected earnings and risk. Thus, we model the firm’s uncontrolled surplus by Itô process and put dividends into the utility functions for investors. Investors who care directly about dividends have a value function that considers the short-term cash inflow from discounted cash dividends and the certainty equivalence of long-term stock holdings subject to bankruptcy risk. This consideration requires us to separate risk aversion from elasticity of intertemporal substitutions (EIS) in the decision making process.
The preference introduced by Epstein and Zin (EZ) [11] is a natural candidate designed for this purpose in the economic literature. In our formulation, we increase the degree of risk aversion of intertemporal preference ordering via a positive parameter \(R\), without affecting “certainty preferences” of a dividend payment policy. We also follow the continuous-time dividend models that the discounted dividend stream is paid in a singular control manner (see Cadenillas et al. [12], Jeanblanc-Picqué and Shiryaev [13], Reppen et al. [10], Shreve et al. [14]). This problem formulation turns out to be a singular control problem with recursive utilities for which we have to develop the well-definedness that there exists a unique value function for the EZ investor within the recursion in singular control setting.
Alternatively, the firm’s executive aims to maximize the expected discounted dividend stream subject to the bankruptcy cost. Although the financial reports indicate the firm’s expected earnings and risk, the executive encounters uncertainty in the predicted earnings. We view this uncertainty as model ambiguity so that the executive’s value function is postulated as the Maenhout’s [15] robust preference. Specifically, the executive specifies ambiguity aversion on the expected earnings, such that the robust dividend decision is made within a maximin singular control problem subject to bankruptcy risk and Maenhout’s regularity.
This paper makes three theoretical contributions. The first contribution is to prove that the EZ singular control utility is well-defined. We establish the existence and uniqueness of the value function satisfying the recursion using backward stochastic differential equations (BSDEs) in Theorem 5. The corresponding proof is non-trivial due to the presence of a non-decreasing process (i.e., singular control) within a non-Lipschitz aggregator, combined with the bankruptcy stopping time. Compared to the existing literature on the EZ consumption model (see, e.g., Aurand and Huang [16], Xing [17]), this paper distinctively emphasizes the risk of ruin rather than solely focusing on cash flow utility. The consideration of bankruptcy time captures the ruin risk, which results in a nonconvex set of dividend strategies that remain constant after ruin. This nonconvexity contributes to the lack of concavity in our newly formulated EZ singular control utility, distinguishing it from typical EZ consumption models. Furthermore, the EZ preference model is based on a power function defined by a positive parameter \(R\), introducing distinct analytical challenges depending on whether \(R>1\) or \(R<1\). This paper focuses on the case where \(R<1\), where, unlike the \(R>1\) scenario, we cannot utilize a priori bounds for solutions of classical BSDEs (see, e.g., [16], [17], Popier [18]). Despite these challenges, this paper provides new insights into the feasibility of the EZ singular control utility.
As the second contribution, we establish the equivalence between the EZ singular control utility and Maenhout’s counterpart (Theorem 9). While the connection between EZ and Maenhout preferences is initially addressed in [15] and formally established in Skiadas [19] for consumption models, this paper is, to the best of our knowledge, the first to develop this equivalence within a singular control setting, even accounting for the bankruptcy time. Unlike the regular control situation discussed in [15], our BSDE framework reveals that the singular control problem with Maenhout preferences cannot be reduced to a classic utility maximization with singular control. Our last contribution develops a novel shooting method to show that under certain regularities on the reference parameters of the firm’s surplus process, the robust dividend policy is a threshold strategy. When the firm’s surplus hits the upside threshold, the dividend is paid. The threshold depends on the expected earnings, volatility and the ambiguity aversion of the executive.
When there is no bankruptcy, the shooting method has been used to a Hamilton-Jacobi-Bellman (HJB) variational inequality (VI) (see Cohen et al. [20]). Our Maenhout’s singular control problem with bankruptcy can be transformed into a HJB-VI. However, the bankruptcy time complicates the analysis because it is involved within the dividend decision. In particular, we have to prove that there is a single threshold on the firm’s surplus process that can shoot two targets on the bankruptcy boundary condition and the HJB-VI, simultaneously.
Our theoretical results lead to the following interpretations for the dividend signaling theory: While investors are informed with the firm’s expected earnings from the firm’s financial statements, executives set dividend threshold to showcase their confidence on realizing the expected earnings. The level of confidence is reflected by the ambiguity aversion parameter in our model. Therefore, our model predicts that dividend policy conveys information about earnings uncertainty or, equivalently, the executive’s confidence in realizing the expected earnings. We call this notion “signaling confidence”.
We now proceed to a discussion of some related literature. In the context of regular stochastic control, most of the fundamental and often technically demanding questions in recursive utility maximization problems are understood fairly well by now (see, e.g., Duffie and Epstein [21], Herdegen et al. [22]–[24], Kraft et al. [25], Matoussi and Xing [26], Melnyk et al. [27], Monoyios and Mostovyi [28], Schroder and Skiadas [29], Xing [17]). While the dominant mathematical framework therein is the BSDE approach, it does not involve a non-decreasing process in the aggregator. Although certain proof techniques in the present paper bear similarities to some in the above references, our choice of singular control with random terminal time makes certain BSDEs arguments intricate.
On the other hand, considerable efforts have been made to characterize classical (i.e., ambiguity-neutral) singular optimal control. Specifically, we refer to e.g. De Angelis et al. [30], Ferrari [31], Fleming and Soner [32], Shreve and Soner [33], Soner et al. [34] for corresponding free boundary problems, and to e.g. Bank [35], Bank and El Karoui [36], Bank and Kauppila [37], Bank and Riedel [38] for stochastic representation theorem approach. Recently, there has been an intensive interest in robust analogue of the above references (e.g., Chakraborty et al. [39], Cohen et al. [20], Ferrari et al. [40], [41], Park et al. [42]). We also refer to Bayraktar and Huang [43], Bayraktar and Yao [44], [45], Nutz and Zhang [46], Park and Wong [47], Park et al. [48], Riedel [49] for relations to optimal stopping time under ambiguity. Finally, for completeness, let us mention that the Maenhout preference, which can be viewed as a form of relative entropy deriving ambiguity-aversion (see, e.g., Anderson et al. [50], Ben-Tal [51], Csiszár [52], Laeven and Stadje [53]), has been utilized in robust optimization problems in finance and economics (see, e.g., Branger and Larsen [54], Jin et al. [55], Maenhout [56], Yi et al. [57]).
While this paper focuses on the EZ utility preference and its equivalence to the Maenhout’s preference, it is instructive to compare it with the forward utility preference (see, e.g., El Karoui and Mrad [58], Liang and Zariphopoulou [59], Musiela and Zariphopoulou [60]–[62]), an alternative important framework for modeling dynamic utility preferences. Both preferences extend beyond the classical expected utility framework to better capture investor behavior in dynamic settings, allowing for time-varying preferences and often leading to nonlinear HJB equations or BSDEs. EZ utilities are recursive and backward-looking, defined via BSDEs and commonly used in macroeconomic modeling. In contrast, forward utility processes evolve without a fixed terminal condition, making them well-suited for problems with uncertain or endogenous horizons.
The rest of the paper is organized as follows. Section 2 introduces the EZ singular control utility process for an admissible endogenous dividend policy, presents its BSDE characterization, and establishes well-posedness results. Section 3 formulates the Maenhout’s robust dividend-value utility process and establishes its equivalence with the EZ singular control utility. Section 4 is devoted to the characterization of the robust dividend policy, using HJB-VI. Section 5 studies the dependence of Maenhout’s robust dividend-value utility and optimal threshold on the ambiguity parameter. Remaining proofs for the main theorems can be found in Sections 6–10.
Following the suggestion of Black [2], we put dividends into an investor’s utility functions through recursive utilities in this section. We first lay out the setup of a stochastic differential utility (SDU) on singular dividend flows. Consider a probability space \((\Omega,\mathcal{F},\mathbb{P})\) that supports a one-dimensional Brownian motion \(W:=(W_t)_{t\ge0}\). Let \(\mathbb{F}:=(\mathcal{F}_{t})_{t\ge 0}\) be the augmented filtration generated by \(W\). We assume that \({\@fontswitch\relax\mathcal{F}}_0\) is trivial. We write \(\mathbb{E}[\cdot]\) for the expectation under \(\mathbb{P}\), with \(\mathbb{E}_t[\cdot]:=\mathbb{E}[\cdot|{\@fontswitch\relax\mathcal{F}}_t]\) denoting the conditional expectations given \({\@fontswitch\relax\mathcal{F}}_t\) for \(t\geq 0\). \(\mathcal{P}\) is the set of all real-valued \(\mathbb{F}\)-progressively measurable processes, and \({\@fontswitch\relax\mathcal{P}}_+\subset\mathcal{P}\) is the subset of all processes that take nonnegative values. Moreover, denote for every \(p\ge1\) and \(t\geq0\) by \(L^p(\mathcal{F}_t)\) the set of all \(\mathcal{F}_t\)-measurable random variable with norm \(\|X\|_{L^p}^p:=\mathbb{E}[|X|^p]<\infty\). The following notations are useful in the rest of the paper. \[\begin{align} \mathcal{T}:=&\{\tau : \tau \text{ is a \mathbb{P}-a.s.~finite \mathbb{F}-stopping time}\},\nonumber\\ \mathcal{A}:=&\{D:=(D_t)_{t\geq0}:D \text{ is \mathbb{F}-progressively measurable, continuous \mathbb{P}-a.s.,} \nonumber\\ &\qquad\qquad\quad\quad\quad \text{and non-decreasing with D_{0-}=0}\}, \end{align}\] where we use \(D_{0-} = 0\) to indicate that \(D_0>0\) can only be achieved by a jump of the process at time zero, followed by a continuous path for all \(t\ge0\).
Throughout this paper, the following quantities are taken as primitives:
Definition 1.
\(D\in \mathcal{A}\) denotes the accumulated dividend flow.
\(\rho > 0\) is the constant discount rate.
\(R \in [0,1)\) is the risk aversion parameter appearing in the Epstein-Zin (EZ) aggregator; see 3 .
\(\tau \in \mathcal{T}\) denotes the ruin time, typically defined as the first time the surplus process hits zero; see 19 .
\(\xi^0_{\tau} \in L^1(\mathcal{F}_\tau)\) represents a lump-sum payment at the ruin time in the classical dividend-value process; see 1 , typically modeled as \(\xi^0_{\tau} = e^{-\rho \tau} \xi_0\) with some positive constant \(\xi_0\) (see, e.g., Yin and Wen [63]).
\(\xi_{\tau} \in L^1(\mathcal{F}_\tau)\) denotes a payoff at the random terminal time (i.e., ruin time) in the recursive utility process; see 2 . In general, \(\xi_{\tau}\) differs from \(\xi^0_{\tau}\); for the EZ utility process, it is often modeled as \(\xi_{\tau} = (\xi^0_{\tau})^{1-R}\); see 4 .
In order to formulate the dividend-value and utility processes mentioned above, we introduce the stochastic integral with respect to (w.r.t.) \(D\) as follows: for any \(\mathbb{F}\)-progressively measurable and locally bounded process \((g_t)_{t\ge 0}\), \[\int_{0}^{\tau} g_{t} dD_t:= g_0D_0+\int_{(0,\tau]}g_{t}dD_t,\] which is well-defined in the Stieltjes sense (see Protter [64]). Here, the jump of \(D\) at time zero is accounted, so that \(\int_0^{t}dD_s = D_t\) for \(t\ge0\).
Then for the classical dividend optimization problem given in Alvarez and Virtanen [65], the dividend-value process \(K^D:=(K^D_t)_{t\ge0}\) associated with \(D\) is defined as \[\begin{align} \label{eq:std95singular95V} K^D_t := \mathbb{E}_{t}\bigg[ \int_{t\wedge\tau}^{\tau} e^{-\rho s} dD_s + \xi^{0}_{\tau} \bigg], \quad t\ge 0. \end{align}\tag{1}\] If we incorporate an intertemporal aggregator denoted as \(g\), then we are able to define a dividend-value recursive utility process \(V^D:=(V^D_t)_{t\ge 0}\) of the investor associated with \(D\) by \[\begin{align} \label{eq:recur95singular95V} V^D_t := \mathbb{E}_t \bigg[ \int_{t\wedge \tau}^{\tau} g(s, V^D_s)dD_s + \xi_{\tau}\bigg], \quad t\ge 0. \end{align}\tag{2}\]
Definition 2. We define \(\mathbb{I}(g,D)\) the set of all processes \(V\in {\@fontswitch\relax\mathcal{P}}\) such that the stochastic integral \(\int g(t,V_t) dD_t\) exists and \(V\) satisfies \(\mathbb{E}[\int_0^{\tau}|g(s,V_s)|dD_s]<\infty\).
\(V\in\mathbb{I}(g,D)\) is a utility process associated with \((g,D)\) if it satisfies 2 ;
Denote by \(\mathbb{UI}(g,D)\) the subset of all utility processes \(V\) that are uniformly integrable.
This formulation is inspired by the concept of SDU (see Duffie and Epstein [21], Epstein and Zin [11], Schroder and Skiadas [29]), implying the dividend-value utility at time \(t\) may depend in a nonlinear way on its value at future times.
To proceed further, we introduce the EZ aggregator (see [11], [21]) given by \[\begin{align} \label{dfn:g95EZ} g_{\text{EZ}}(s,v): = (1-R)e^{-\rho s}v^{\frac{-R}{1-R}}. \end{align}\tag{3}\] The EZ singular control utility process \(V^{\text{EZ},D} = (V^{\text{EZ},D}_t)_{t \geq 0}\) is defined as \[\begin{align} \label{eq:EZ95V} V^{\text{EZ},D}_t = \mathbb{E}_t \bigg[ \int_{t\wedge \tau}^{\tau} e^{-\rho s}(1-R) (V^{\text{EZ},D}_s)^{\frac{-R}{1-R}}dD_s + \xi_{\tau}\bigg], \quad t\ge 0. \end{align}\tag{4}\] Here we observe that when \(R=0\), \(V^{\text{EZ},D}\) reduces to \(K^D\) given in 1 . In this setting, the degree of risk aversion of intertemporal preference ordering is increased (i.e., \(R>0\)) without affecting the “certainty preferences”.
Remark 3. The utility specification in 4 is motivated by its connection to the conventional EZ aggregator in the infinite elasticity of intertemporal substitution limit. While \(g_{\text{EZ}}\) in 3 differs from the standard form by a scaling factor \((1-R)\) (rather than \((1-R)^{-\frac{R}{1-R}}\)), this does not affect preferences for any \(R \in (0,1)\) (see Herdegen et al. [23]). Our scaling choice in 4 clarifies the equivalence between EZ utility and robust Maenhout utility (see Theorem 9). For \(R \in (1,\infty)\), the sign and domain of the generator become more delicate, see e.g. Aurand and Huang [16], Herdegen et al. [23], [24], Xing [17] for further discussion.
In the remainder of this section, our goal is to establish the well-posedness (i.e., existence and uniqueness) of the EZ singular control utility given by 4 . Given a triplet \((\tau, D, \xi_{\tau})\) satisfying the appropriate conditions, the EZ utility is characterized as the unique solution \((Y, Z)\) to the following BSDE involving the dividend process \(D\) and the EZ aggregator: \[\begin{align} \label{dfn:BSDE1} Y_t = \xi_{\tau} + \int_{t \wedge \tau}^{\tau} g_{\text{EZ}}(s, Y_s) \, dD_s - \int_{t \wedge \tau}^{\tau} Z_s \, dW_s, \quad t \geq 0. \end{align}\tag{5}\] According to Briand et al. [66] and Popier [18], we introduce the following definition of a solution of BSDEs with random terminal times. First we denote for every \(p\geq 1\) by \(L_{\mathrm{loc}}^p(\mathbb{R}_+)\) the set of all locally \(p\)-integrable functions on \({\mathbb{R}_+}\).
Definition 4. A solution of the BSDE 5 with parameter \((\tau,\xi_{\tau},g_{\text{EZ}},D)\) is a pair \((Y_t,Z_t)_{t\ge0}\in {\@fontswitch\relax\mathcal{P}}_+\times {\@fontswitch\relax\mathcal{P}}\) satisfying the following conditions \(\mathbb{P}\)-a.s.:
\(Y_t = \xi_{\tau}\) and \(Z_t = 0\) on \(\{t \ge \tau\}\);
\(t\mapsto \mathbb{1}_{t\le\tau} g_{\text{EZ}}(t,Y_t)\) belongs to \(L^1_{\mathrm{loc}}(\mathbb{R}_+)\), and \(t\mapsto Z_t\) belongs to \(L_{\mathrm{loc}}^2(\mathbb{R}_+)\);
For every \(T\ge 0\), it holds that for \(t\in[0,T]\), \[\begin{align} Y_{t\wedge\tau} = Y_{T\wedge\tau} + \int_{t\wedge\tau}^{T\wedge\tau} g_{\text{EZ}}(s,Y_s) dD_s - \int_{t\wedge\tau}^{T\wedge\tau}Z_s dW_s. \end{align}\]
Furthermore, we call the solution \((Y,Z) = (Y_t,Z_t)_{t\ge0}\) of 5 an \(L^2\)-solution if \[\begin{align} \mathbb{E}\bigg[ \sup_{t\ge0} |Y_{t\wedge\tau}|^2 + \int_{0}^{\tau} |Z_t|^2 dt \bigg] <\infty \end{align}\] holds. In particular, the \(L^2\)-solution implies that \(Y\in\mathbb{UI}(g_{\text{EZ}},D)\).
Condition 1. The triplet \((\tau, D, \xi_{\tau})\) satisfies the following integrability and positivity conditions: \[\int_{0}^{\tau} e^{-\rho s} dD_s \in L^2(\mathcal{F}_\tau), \quad \xi_{\tau} \in L^{\frac{2}{1-R}}(\mathcal{F}_\tau), \quad \text{and} \quad \xi_{\tau} > 0 \quad\mathbb{P}\text{-a.s..}\]
Theorem 5. Let \((\tau, D, \xi_{\tau})\) satisfy Condition 1, and let \(g_{\mathrm{EZ}}(\cdot, \cdot)\) be defined as in 3 . Then the following statements hold:
There exists a unique utility process \(V^{\text{EZ},D} \in \mathbb{UI}(g_{\mathrm{EZ}}, D)\) with continuous paths, which is strictly positive and satisfies \(\mathbb{E}[\sup_{t \geq 0} (V^{\text{EZ},D}_t)^2] < \infty\). Moreover, there exists a process \(Z \in \mathcal{P}\) such that \(\int_0^{\tau} Z_t^2 \, dt < \infty\) \(\mathbb{P}\)-a.s., and the pair \((V^{\text{EZ},D}, Z)\) solves the BSDE 5 .
If, in addition, \(\xi_{\tau} > C\) \(\mathbb{P}\)-a.s.for some positive constant \(C\), then there exists a unique \(L^2\)-solution \((V^{\text{EZ},D}, Z) \in \mathcal{P}_+ \times \mathcal{P}\) to 5 . In particular, \(V^{\text{EZ},D}\) has continuous paths, satisfies \(V^{\text{EZ},D}_t \geq C\) \(\mathbb{P}\)-a.s.for every \(t \geq 0\), and \(\mathbb{E}[\sup_{t \geq 0} (V^{\text{EZ},D}_t)^{\frac{2}{1-R}}] < \infty\). Therefore, \(V^{\text{EZ},D}\) belongs to \(\mathbb{UI}(g_{\mathrm{EZ}}, D)\) and is the unique utility process associated with \((g_{\mathrm{EZ}}, D)\).
Theorem 5 i. relies on a restrictive condition (i.e., \(\xi_{\tau}>C\) \(\mathbb{P}\)-a.s.), while the corresponding results allow us to construct a sequence of \(L^2\)-solutions of BSDEs that converges to a utility process in \(\mathbb{UI}(g_{\text{EZ}},D)\) without requiring this condition–which in turn establishes the existence result in Theorem 5 i.. In what follows, we provide the proof of Theorem 5 i., while the proof of Theorem 5 ii. is deferred to Section 6, where all the steps for the proof are presented in detail and Theorem 34 consolidates them with proving part ii.
Proof of Theorem 5i.. Throughout the proof, we suppress the superscript of \(V^{\text{EZ},D}\) and simply write \(V\) for notational convenience. For each \(n\in\mathbb{N}\), set \(\xi_{\tau}^n := \frac{1}{n}\vee \xi_{\tau}\). By Theorem 34, there exists a unique \(L^2\)-solution \((Y^n,Z^n)\in {\@fontswitch\relax\mathcal{P}}_+\times {\@fontswitch\relax\mathcal{P}}\) of 5 (with \(\xi_\tau\) replaced by \(\xi_{\tau}^n\)) such that \[\begin{align} Y^n_t = \mathbb{E}_t \bigg[ \int_{t\wedge \tau}^{\tau} e^{-\rho s}(1-R) (Y^n_s)^{\frac{-R}{1-R}}dD_s + \xi^n_{\tau}\bigg],\quad t\ge0. \end{align}\] Note that \(Y_t^n\) is non-increasing in \(n \in \mathbb{N}\) for every \(t\geq0\) (see Proposition 29) and \(Y^n_t > 0\) \(\mathbb{P}\)-a.s. for every \(t\geq0\) and \(n\in \mathbb{N}\). Hence \(V_t:= \lim_{n\rightarrow \infty} Y_t^n\) for \(t\geq0\) is well-defined. Moreover, the monotone convergence theorem (MCT) ensures that for \(t\geq 0\), \[\begin{align} V_t =& \lim_{n\rightarrow\infty}\mathbb{E}_t \bigg[ \int_{t\wedge \tau}^{\tau} e^{-\rho s}(1-R) (Y^n_s)^{\frac{-R}{1-R}}dD_s + \xi^n_{\tau}\bigg] \\ =& \mathbb{E}_t \bigg[ \int_{t\wedge \tau}^{\tau} e^{-\rho s}(1-R) (V_s)^{\frac{-R}{1-R}}dD_s + \xi_{\tau}\bigg]. \end{align}\] Since \(\xi_{\tau}>0\) \(\mathbb{P}\)-a.s., it holds that \(V_t \ge \mathbb{E}_t[\xi_{\tau}]>0\) \(\mathbb{P}\)-a.s. for all \(t\geq0\).
We claim that \(V\in\mathbb{UI}(g_{\text{EZ}},D)\). Since \(Y_0^n\) is non-increasing in \(n \in \mathbb{N}\) and \(V_0=\lim_{n\rightarrow \infty}Y^n_0\), it holds that \(\mathbb{E}[ \int_{0}^{\tau} e^{-\rho s}(1-R) (V_s)^{\frac{-R}{1-R}}dD_s] < V_0 \le Y_0^n<\infty\) for every \(n\in\mathbb{N}\).
Similarly, we have \(\mathbb{E}[\sup_{t\ge 0} (V_t)^{\frac{2}{1-R}}] \le \mathbb{E}[\sup_{t\ge 0} (Y^n_t)^{\frac{2}{1-R}}]<\infty\). This ensures that \(\mathbb{E}[\sup_{t\ge 0} (V_t)^2]<\infty\) by Jensen’s inequality with exponent \(\frac{1}{1-R}>1\). Hence, \(V\) is uniformly integrable and therefore is of class \(\mathbb{UI}(g_{\text{EZ}},D)\).
It remains to show that \(V\) is the unique utility process. By the martingale representation theorem, there exists \(Z\in \mathcal{P}\) such that \(\int_0^{\tau} Z_t^2 dt <\infty\) \(\mathbb{P}\)-a.s. and \((V,Z)\) solves 5 . Let \(V'\) be another utility process of class \(\mathbb{UI}(g_{\text{EZ}},D)\) and \(Z'\in \mathcal{P}\) be such that \(\int_0^{\tau} (Z'_t)^2 dt <\infty\) \(\mathbb{P}\)-a.s., and \((V',Z')\) solves 5 . Here we note that \(V_t = V'_t = \xi_{\tau}\) on \(\{t\ge \tau\}\).
Set \((\Delta V, \Delta Z):= (V-V', Z-Z')\). Then (\(\Delta V, \Delta Z)\) solves for all \(t\geq0\) \[\begin{align} \Delta V_t = \int_{t\wedge\tau}^{\tau} \frac{g_{\text{EZ}}(s,V_s) - g_{\text{EZ}}(s,V'_s)}{\Delta V_s} \Delta V_s \mathbb{1}_{\Delta V_s\neq0} dD_s - \int_{t\wedge\tau}^{\tau} \Delta Z_s dW_s. \end{align}\] Let \(\alpha=(\alpha_s)_{s\ge0}\) be defined by \(\alpha_s: = \mathbb{1}_{\Delta V_s\neq0}(g_{\text{EZ}}(s,V_s) - g_{\text{EZ}}(s,V'_s))/\Delta V_s\). Since \(y\mapsto g_{\text{EZ}}(\cdot,y)\) is decreasing, we have \(\alpha\le0\). It then follows that \[\begin{align} e^{\int_{0}^t \alpha_u du}\Delta V_t = - \int_{t\wedge\tau}^{\tau} e^{\int_{0}^s \alpha_u du} \Delta Z_s dW_s\quad for all t\geq0. \end{align}\] From the fact that \(\alpha\le 0\) and \(V,V'\in\mathbb{UI}(g_{\text{EZ}},D)\), the local martingale \(\int_0^{t}e^{\int_{0}^s \alpha_u du} \Delta Z_s dW_s\) is a martingale. Hence, \(\Delta V_t= 0\) \(\mathbb{P}\)-a.s. for all \(t\geq0\). ◻
Remark 6. If we let \(\widetilde{\@fontswitch\relax\mathcal{A}} \subseteq {\@fontswitch\relax\mathcal{A}}\) denote a subset of dividend processes \(D\) that satisfy Condition 1 and any additional admissibility requirements, then the dividend optimization problem under EZ preferences is \[\begin{align} \label{eq:opt95EZ} V^{\text{EZ},*} := \sup_{D \in \widetilde{\mathcal{A}}} (V^{\text{EZ},D}_0)^{\frac{1}{1-R}}. \end{align}\tag{6}\] We specify \(\widetilde{\@fontswitch\relax\mathcal{A}}\) and solve this problem in Section 4, after establishing its equivalence with the robust criterion under Maenhout’s preference in Section 3.
In this section, we take the perspective of the firm’s executive, who aims to maximize expected discounted dividends while considering bankruptcy costs. At the same time, she/he suffers from ambiguity regarding the firm’s earnings or the drift of the firm’s surplus. In the presence of ambiguity, the executive prefers a family of unspecified alternative models that are in close proximity to the reference model. In line with this, we start with defining a collection of processes that serves as the Girsanov kernels. Then we aim to link the investor’s EZ singular control utility in Section 2 with the executive’s robust dividend policy by establishing an equilibrium where both parties are aligned.
Throughout this section, let \({\@fontswitch\relax\mathcal{R}}\in [0,1)\) denote the ambiguity aversion parameter in the robust singular control criterion. Moreover, we consider the primitive quantities \((\tau, D, \xi_{\tau})\) as in Definition 1, assuming that they satisfy Condition 1 under setting \(R \equiv {\@fontswitch\relax\mathcal{R}}\).
Definition 7. A Girsanov kernel process \((\theta_t)_{t\ge 0}\in \mathcal{P}\) is said to be admissible if the stochastic exponential \((\eta^{\theta}_t)_{t\ge0}\) defined as \[\begin{align} \label{eq:eta} \eta_t^{\theta} = \exp\bigg(\int_{0}^t \theta_s dW_s - \frac{1}{2}\int_0^t \theta_s^2 ds\bigg), \quad t\ge0, \end{align}\tag{7}\] is a martingale. We denote by \(\Theta\) the admissible set of all Girsanov kernels \(\theta\) .
For each \(\theta \in {\@fontswitch\relax\mathcal{P}}\), we define a probability measure \(\mathbb{Q}^{\theta}\) on \((\Omega,\mathcal{F},\mathbb{F})\) by \[\begin{align} \label{eq:measureQ} \frac{d\mathbb{Q}^{\theta}}{d\mathbb{P}}\bigg\vert_{\mathcal{F}_T} = \eta^{\theta}_T\quad for each T>0. \end{align}\tag{8}\] Then, the process \(W_t^{\theta}: = W_t - \int_0^t \theta_s ds\), \(t\in[0,T]\), is a \(\mathbb{Q}^{\theta}\)-Brownian motion. We denote by \(\mathbb{E}^\theta_t[\cdot]\) the conditional expectation under \(\mathbb{Q}^\theta\) given \(\mathcal{F}_t\). It follows from Skiadas [19] that for any \(\tau \in {\@fontswitch\relax\mathcal{T}}\), the discounted relative entropy process, given as \(\frac{1}{2}\mathbb{E}^{\theta}_t[ \int_{t\wedge\tau}^{\tau} e^{-\rho (s-t)} \theta_s^2 ds]\), \(t>0\), is well-defined and takes values in \([0,\infty]\) \(\mathbb{P}\)-a.s.
Following the ambiguity-averse preference framework of Maenhout [15], for any \({\@fontswitch\relax\mathcal{R}}\in(0,1)\), we define the robust singular control criterion for the executive, denoted by \(V^{\text{rob}, D}:=(V_t^{\text{rob}, D})_{t \geq 0}\), as \[\begin{align} \label{dfn:rbs95V} V^{\text{rob},D}_t := \mathop{\mathrm{ess\,inf}}\{V_{t}^{\theta,D}:\theta\in \Theta^D \}, \quad t\ge 0, \end{align}\tag{9}\] where \[\begin{align} \tag{10} V_{t}^{\theta,D} &:= \mathbb{E}^{\theta}_t\bigg[ \int_{t\wedge\tau}^{ \tau} e^{-\rho s} dD_s +(\xi_{\tau})^{\frac{1}{1-{\@fontswitch\relax\mathcal{R}}}}\bigg] + \frac{1}{2{\@fontswitch\relax\mathcal{R}}} \mathbb{E}^{\theta}_t\bigg[\int_{t\wedge\tau}^{\tau} V^{\theta,D}_s \theta_s^2 ds\bigg],\\ \Theta^D&:= \bigg\{\theta \in \Theta : \theta satisfies that \mathbb{E}\bigg[\bigg(\int_{0}^{\tau} \eta_{t}^{\theta}e^{\int_{0}^{t} \frac{\theta^2_s}{2{\@fontswitch\relax\mathcal{R}}}ds -\rho t} dD_t\bigg)^2\bigg]<\infty, \nonumber \\ &\qquad \qquad \qquad and that\mathbb{E}\big[(\eta_{\tau}^{\theta})^2e^{\int_{0}^{\tau} \frac{\theta^2_u}{{\@fontswitch\relax\mathcal{R}}}du } \big((\xi_{\tau})^{\frac{2}{1-{\@fontswitch\relax\mathcal{R}}}}\vee 1\big) \big]<\infty \bigg\}, \tag{11} \end{align}\] with \(\eta^{\theta}\) defined in 7 . The objective function \(V^{\theta,D}\) consists of two expected terms: i. present value of the future dividend payments under alternative probability measure \(\mathbb{Q}^{\theta}\); ii. a penalty term reflecting the agent’s ambiguity aversion, scaled w.r.t. the objective function. The ambiguity aversion of the executive is thus increasing in \({\@fontswitch\relax\mathcal{R}}\in(0,1)\). For the definition of \(V^{\text{rob},D}\) at \({\@fontswitch\relax\mathcal{R}}=0\), see Remark 22 in Section 5, where its continuity w.r.t. \({\@fontswitch\relax\mathcal{R}}\in[0,1)\) is established. For the remainder of this section, we fix \({\@fontswitch\relax\mathcal{R}} \in (0,1)\).
Proposition 8. For any \(\theta\in \Theta^D\), there exists a utility process \(V^{\theta,D}\) satisfying 10 , and it is unique in the class \(\{Y\in\mathcal{P}:\mathbb{E}[\sup_{t\ge 0} Y_{t\wedge\tau}^2]<\infty\}\). Hence, \(V^{\text{rob},D}\) in 9 is well-defined. Furthermore, \(V^{\theta,D}\) admits the representation \[\begin{align} \label{eq:thm:rob95rec:0} V_t^{\theta,D} = \mathbb{E}^{\theta}_t\bigg[ \int_{t\wedge\tau}^{ \tau} e^{\int_{t}^{s} \frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du -\rho s} dD_s + e^{\int_{t\wedge\tau}^{\tau} \frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du }(\xi_{\tau})^{\frac{1}{1-{\@fontswitch\relax\mathcal{R}}}}\bigg],\quad t\ge0. \end{align}\qquad{(1)}\]
Proof. We start with constructing \(V^{\theta,D}\) given in ?? and then show that it satisfies 10 . As \(\theta\in \Theta^D\), by Hölder’s inequality, \[\mathbb{E}[\eta_{\tau}^{\theta}e^{\int_{0}^{\tau} \frac{\theta_u^2}{2{\@fontswitch\relax\mathcal{R}}}du }(\xi_{\tau})^{\frac{1}{1-{\@fontswitch\relax\mathcal{R}}}} ]\le \mathbb{E}[(\eta_{\tau}^{\theta})^2e^{\int_{0}^{\tau} \frac{\theta^2_u}{{\@fontswitch\relax\mathcal{R}}}du } ]^{\frac{1}{2}} \mathbb{E}[(\xi_{\tau})^{\frac{2}{1-{\@fontswitch\relax\mathcal{R}}}}]^{\frac{1}{2}}<\infty.\] Furthermore, Jensen’s inequality (with exponent 2) ensures that \[\mathbb{E}\bigg[ \int_{0}^{ \tau} \eta_s^{\theta} e^{\int_{0}^{s} \frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du -\rho s} dD_s \bigg]\leq \mathbb{E}\bigg[\bigg( \int_{0}^{ \tau} \eta_s^{\theta} e^{\int_{0}^{s} \frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du -\rho s} dD_s\bigg)^2 \bigg] <\infty.\] Hence, the representation in ?? is a well-defined utility process.
By Jensen’s inequality, we have \(\mathbb{E}[\sup_{t\ge 0} (V^{\theta,D}_{t\wedge\tau})^2]<\infty\). It further implies that there exists \({\iota}\in\mathcal{P}\) such that \(\int_0^{\infty} \iota^2_t dt<\infty\), \(\mathbb{Q}^\theta\)-a.s., and \(({V}^{\theta,D},{\iota})\) solves \[\begin{align} \label{eq:thm:rob95rec:1} {V}^{\theta,D}_t = (\xi_{\tau})^{\frac{1}{1-{\@fontswitch\relax\mathcal{R}}}} + \int_{t\wedge\tau}^{\tau} e^{-\rho s} dD_s + \int_{t\wedge\tau}^{\tau} \frac{1}{2{\@fontswitch\relax\mathcal{R}}}{V}^{\theta,D}_s\theta^2_s ds - \int_{t\wedge\tau}^{\tau} {\iota}_s dW_s^{\theta}. \end{align}\tag{12}\] Since \((M_t)_{t\ge0} = (\int_{0}^{t} {\iota}_s dW_s^{\theta})_{t\ge0}\) is a local martingale under \(\mathbb{Q}^{\theta}\), we shall find an increasing sequence of stopping times \(\{T_n\}_{n\in\mathbb{N}}\) such that \(T_n\ge t\), \(\mathbb{Q}^{\theta}(\lim_{n\rightarrow\infty} T_n = \infty) = 1\) and the stopped process \((M_{t\wedge T_n})_{t\ge 0}\) is a martingale under \(\mathbb{Q}^{\theta}\). Applying \(\mathbb{E}_t^{\theta}\) to both sides of 12 gives \[\begin{align} V_{t}^{\theta,D} = \mathbb{E}^{\theta}_t\bigg[ \int_{t\wedge\tau}^{T_n\wedge\tau} e^{-\rho s} dD_s\bigg] + \frac{1}{2{\@fontswitch\relax\mathcal{R}}} \mathbb{E}^{\theta}_t\bigg[\int_{t\wedge\tau}^{T_n\wedge\tau} V^{\theta,D}_s \theta_s^2 ds\bigg]+ \mathbb{E}^{\theta}_t[V_{T_n\wedge\tau}^{\theta,D}]. \end{align}\] Then the first two terms on the right-hand side converge as \(n\to \infty\) by MCT, while the last term converges by the dominated convergence theorem (DCT). We conclude that \(V_t^{\theta}\) defined via ?? satisfies 10 . The uniqueness follows from standard localization arguments, and we omit it here. ◻
The main result of this section extends the core equivalence between EZ utility and Maenhout’s robustness criterion to settings with singular control and random time horizons (such as ruin time) — a case not addressed in the existing literature. The general roadmap for this equivalence was developed by Skiadas [19], with a recent alternative perspective provided by Maenhout et al. [67]. While some arguments in our proof build on techniques from [19], our contribution is to rigorously develop the technical details necessary for this extension, particularly in establishing the well-posedness of the EZ utility process under these more general conditions (see Theorem 5 and Section 6).
Theorem 9. Recall that \(g_{\text{EZ}}\) is given in 3 with setting \(R\equiv{\@fontswitch\relax\mathcal{R}}\). Consider \(\theta\in\Theta^D\), let \(V^{\text{rob}, D}\) and \(V^{\theta,D}\) be defined as in 9 and 10 respectively.
Recall that \(V^{\text{EZ},D}\) is the unique utility process associated with \((g_{\text{EZ}},D)\) (see Theorem 5 i.). It holds that \(V^{\text{EZ},D} = (V^{\text{rob},D})^{1-R}\).
If \(\xi_{\tau}>C\) \(\mathbb{P}\)-a.s. with \(C>0\), let \((V^{\text{EZ},D},Z)\) be the unique \(L^2\)-solution to 5 with \((\tau,\xi_{\tau},g_{\text{EZ}},D)\) (see Theorem 5 ii.), and set \(\mathbb{Y} := (V^{\text{EZ},D})^{\frac{1}{1-R}}\) and \({\mathbb{Z}} := (\mathbb{Y})^{R}{Z}/(1-R)\), then it holds \[\begin{align} \label{eq:thm:rob95rec95link:2} V^{\theta,D}_t = \mathbb{Y}_t + \mathbb{E}_t^{\theta}\bigg[\int_{t\wedge \tau}^{\tau} \frac{R}{2{\mathbb{Y}}_s} \Big({\mathbb{Z}}_s + \frac{{\mathbb{Y}}_s}{R}\theta_s\Big)^2 e^{\frac{1}{2R}\int_t^{s}\theta^2_u du } d s \bigg]. \end{align}\qquad{(2)}\] Consequently, \(V^{\text{rob},D} \ge \mathbb{Y}\). In addition, \(\theta^*:= (-R\mathbb{Z}_t/\mathbb{Y}_t)_{t\ge0} \in \Theta^D\), which implies that \(V^{\text{rob},D} = V^{\theta^*,D} = \mathbb{Y} = (V^{\text{EZ},D})^{\frac{1}{1-R}}\).
Proof. We start by proving ii., which is stated under the condition that \(\xi_{\tau}>C\) \(\mathbb{P}\)-a.s.. By Theorem 5 ii. and Itô’s lemma, \((\mathbb{Y},\mathbb{Z})\) solves \[\begin{align} \mathbb{Y}_t =& (\xi_{\tau})^{\frac{1}{1-R}} + \int_{t\wedge\tau}^{\tau} e^{-\rho s} dD_s - \frac{R}{2} \int_{t\wedge\tau}^{\tau} \frac{(\mathbb{Z}_s)^2}{\mathbb{Y}_s}ds - \int_{t\wedge\tau}^{\tau} \mathbb{Z}_s dW_s \\ =& (\xi_{\tau})^{\frac{1}{1-R}}+ \int_{t\wedge\tau}^{\tau} e^{-\rho s} dD_s- \frac{R}{2} \int_{t\wedge\tau}^{\tau} \frac{(\mathbb{Z}_s)^2}{\mathbb{Y}_s}ds- \int_{t\wedge\tau}^{\tau}\mathbb{Z}_s\theta_s ds - \int_{t\wedge\tau}^{\tau} \mathbb{Z}_s dW_s^{\theta}. \end{align}\] For any \(\theta\in \Theta^D\), denote by \(({V}^{\theta,D},{\iota}^{\theta})\) a solution of 12 . Then \[\begin{align} {V}^{\theta,D}_t -\mathbb{Y}_t &= \int_{t\wedge\tau}^{\tau}\bigg( \frac{R}{2\mathbb{Y}_s} \bigg(\mathbb{Z}_s + \frac{\mathbb{Y}_s}{R}\theta_s\bigg)^2 + \frac{\theta_s^2({V}^{\theta,D}_s -\mathbb{Y}_s )}{2R}\bigg) ds \\ &\quad - \int_{t\wedge\tau}^{\tau}({\iota}^{\theta}_s - \mathbb{Z}_s) dW_s^{\theta}. \end{align}\] Set \(\beta_{t}: = \exp(\frac{1}{{2R}}\int_0^{t}{\theta^2_s} ds)\) for \(t\geq 0\). Then using Itô’s lemma, \[\begin{align} \label{eq:pf:thm:rob95rec951} \beta_{t}( {V}^{\theta,D}_t -\mathbb{Y}_t) = \int_{t\wedge\tau}^{\tau}\beta_{s} \frac{R}{2\mathbb{Y}_s} \Big(\mathbb{Z}_s + \frac{\mathbb{Y}_s}{R}\theta_s\Big)^2ds - (M_{\tau} - M_{t\wedge \tau}), \end{align}\tag{13}\] where \(M_t := \int_0^t \beta_{s}({\iota}^{\theta}_s - \mathbb{Z}_s)dW^{\theta}_s\). Since \((M_s)_{s\geq0}\) is a \(\mathbb{Q}^{\theta}\)-local martingale, we shall find an increasing sequence of stopping times \(\{T_n\}_{n\in\mathbb{N}}\) such that \(T_n\ge t\), \(\mathbb{Q}^{\theta}(\lim_{n\rightarrow\infty} T_n = \infty) = 1\) and the stopped process \((M_{s\wedge T_n})_{s\geq t}\) is a \(\mathbb{Q}^{\theta}\)-martingale. Applying \(\mathbb{E}_t^{\theta}\) to both sides of 13 , we have for \(t\ge 0\) \[\begin{align} {V}^{\theta,D}_t -\mathbb{Y}_t &= \mathbb{E}_t^{\theta}\bigg[ \int_{t\wedge \tau}^{T_n\wedge \tau}\frac{R}{2\mathbb{Y}_s} \Big(\mathbb{Z}_s + \frac{\mathbb{Y}_s}{R}\theta_s\Big)^2 e^{\int_{t\wedge \tau}^{s}\frac{\theta^2_u}{2R} du} d s\bigg] \nonumber\\ &\quad + \mathbb{E}_t^{\theta} [e^{\int_{t\wedge \tau}^{T_n\wedge \tau}\frac{\theta^2_u}{2R} du} {V}^{\theta,D}_{T_n\wedge \tau}]-\mathbb{E}_t^{\theta} [e^{\int_{t\wedge \tau}^{T_n\wedge \tau}\frac{\theta^2_u}{2R} du}\mathbb{Y}_{T_n\wedge \tau}]\\ &=: \operatorname{I}_t^n+\operatorname{II}_t^n - \operatorname{III}_t^n. \end{align}\] We first note that \(\operatorname{I}_t^n\) converges to the second term given in ?? as \(n\to \infty\), by the MCT (as the integrand is nonnegative). Next, by using the representation given in ?? , we have that as \(n\to \infty\) \[\begin{align} \operatorname{II}_t^n = \mathbb{E}_t^{\theta}\bigg[ \int_{T_n\wedge\tau}^{\tau} e^{\int_{t\wedge \tau}^{s}\frac{\theta^2_u}{2R} du-\rho s} dD_s + e^{\int_{t\wedge \tau}^{\tau}\frac{\theta^2_u}{2R} du} (\xi_{\tau})^{\frac{1}{1-R}}\bigg] \end{align}\] goes to \(\mathbb{E}_t^{\theta}[e^{\int_{t\wedge \tau}^{\tau}\frac{\theta^2_u}{2R} du} (\xi_{\tau})^{\frac{1}{1-R}}]\), where the convergence holds by the MCT (as the integrand is nonnegative).
Lastly, we claim that the term \(\operatorname{III}_t^n\) converges to the same limit as \(\operatorname{II}_t^n\). Indeed, by the conditions that \(\theta\in\Theta^D\) and \(\xi_{\tau}\ge C\) \(\mathbb{P}\)-a.s., we have \[\begin{align} \mathbb{E}^{\theta}_t\bigg[e^{\int_{0}^{\tau}\frac{\theta^2_u}{2R} du} \sup_{s\ge t}\mathbb{Y}_{s\wedge\tau}\bigg] &\le \mathbb{E}_t\bigg[\frac{\eta^{\theta}_{\tau}}{\eta^{\theta}_{t}}e^{\int_{0}^{\tau}\frac{\theta^2_u}{2R} du} \xi_{\tau}^{\frac{1}{1-R}} \sup_{t\ge0}\Big(\mathbb{Y}_{t\wedge\tau}\xi_{\tau}^{-\frac{1}{1-R}}\Big) \bigg]\\ &\le \frac{C^{-\frac{1}{1-R}}}{\eta^{\theta}_{t}}\mathbb{E}_t [(\eta^{\theta}_{\tau})^2e^{\int_{0}^{\tau}\frac{\theta^2_u}{R} du} \xi_{\tau}^{\frac{2}{1-R}}]^{\frac{1}{2}} \mathbb{E}_t\bigg[\sup_{s\ge t} \mathbb{Y}_{s\wedge\tau}^2\bigg]^{\frac{1}{2}}<\infty, \end{align}\] where the second inequality holds by Hölder’s inequality.
Therefore, \(\mathbb{E}_t^{\theta}[e^{\int_{t\wedge \tau}^{T_n\wedge \tau}\frac{\theta^2_u}{2R} du} \mathbb{Y}_{T_n\wedge \tau}]\) is dominated by \(\mathbb{E}_t^{\theta}[e^{\int_{0}^{\tau}\frac{\theta^2_u}{2R} du} \sup_{s\ge t}\mathbb{Y}_{s\wedge\tau}]\) which is finite from the above analysis. Hence we can apply the MCT to ensure the claim to hold. Therefore, the equality ?? holds.
Moreover, ?? implies that \(V^{\theta,D}_t \ge \mathbb{Y}_t = (V^{\text{EZ},D}_t)^{\frac{1}{1-R}}\) and equality holds for \(\theta^* = -R\mathbb{Z}/\mathbb{Y}\) if \(\theta^*\in \Theta^D\). It remains to prove that \(\theta^* = -R\mathbb{Z}/\mathbb{Y}\in \Theta^D\). We first show that \((\eta^{\theta^*}_t)_{t\geq 0}\) is a martingale for \(t\ge0\). By Itô’s lemma, \[\begin{align} \label{eq:pf95thm346295logY} \ln\Big(\frac{\mathbb{Y}_{t\wedge\tau}}{\mathbb{Y}_0}\Big) = \int_{0}^{t\wedge\tau} \frac{R-1}{2} \Big(\frac{\mathbb{Z}_s}{\mathbb{Y}_s}\Big)^2 ds + \int_{0}^{t\wedge\tau} \frac{\mathbb{Z}_s}{\mathbb{Y}_s} dW_s -\int_{0}^{t\wedge\tau} \frac{e^{-\rho s}}{\mathbb{Y}_s} dD_s. \end{align}\tag{14}\] Since \((\mathbb{Y}_t)_{t\ge0}\) is continuous and never \(0\), \(1/\mathbb{Y}\) is locally bounded. Moreover, by Definition 4 i., it holds that \(\theta^* = 0\) on \(\{t\ge\tau\}\), Hence we have for\(t\ge0\), \[\begin{align} &\int_{0}^{t} \theta^*_s dW_s -\int_0^t\frac{(\theta^*_s)^2}{2}ds\\ &\quad = -R\ln\Big(\frac{\mathbb{Y}_{t\wedge\tau}}{\mathbb{Y}_0}\Big) - R\int_{0}^{t\wedge\tau}\bigg( \frac{e^{-\rho s}}{\mathbb{Y}_s} d D_s + \frac{1}{2}\bigg( \frac{\mathbb{Z}_s}{\mathbb{Y}_s}\bigg)^2ds\bigg). \end{align}\] By definition of \(\eta^{\theta^*}\) in 7 , we have that for \(t\ge0\), \(\mathbb{P}\text{-a.s.}\), \[\begin{align} \label{eq:pf95thm346295eta} \eta^{\theta^*}_t =\Big(\frac{\mathbb{Y}_0}{\mathbb{Y}_{t\wedge\tau}}\Big)^R e^{ - \int_{0}^{t\wedge\tau} \frac{R}{\mathbb{Y}_s}e^{-\rho s} d D_s -\frac{R}{2} \int_{0}^{t\wedge\tau}( \frac{\mathbb{Z}_s}{\mathbb{Y}_s})^2ds}\le \mathbb{Y}_0^R C^{-\frac{R}{1-R}}. \end{align}\tag{15}\] Here we have used the condition that \(\xi_{\tau}> C\) \(\mathbb{P}\)-a.s.. Hence \(V_t^{\text{EZ},D}\ge C\) \(\mathbb{P}\)-a.s., for \(t\geq 0\) (see Theorem 5 ii.). We know that \((\eta^{\theta^*}_t)_{t\geq 0}\) is a \(\mathbb{P}\)-local martingale. The uniform bound 15 further implies that it is a \(\mathbb{P}\)-martingale.
Next we verify \(\theta^*\in \Theta^D\) (see 11 ). For the first condition, by 14 \[\begin{align} \mathbb{E}[(\eta^{\theta^*}_{\tau})^2e^{\int_0^{\tau} \frac{(\theta^*_s)^2}{R} ds} (\xi_{\tau})^{\frac{2}{1-R}}] = \mathbb{E}[e^{ - 2R\int_{0}^{\tau} \frac{1}{\mathbb{Y}_s}e^{-\rho s} dD_s} \mathbb{Y}_{\tau}^{2(1-R)}Y_0^{2R}]<\infty, \end{align}\] where we have used the fact that \({\@fontswitch\relax\mathcal{F}}_0\) is trivial (see Section 2) in the inequality.
For the last condition, it holds that \[\begin{align} &\mathbb{E}\bigg[\bigg(\int_{0}^{\tau} e^{\int_{0}^{t} \frac{(\theta^*_s)^2}{2R}ds -\rho t}\eta_t^{\theta^*} dD_t\bigg)^2\bigg] = \mathbb{E}\bigg[\bigg(\int_{0}^{\tau} e^{ R(\ln(\frac{\mathbb{Y}_0}{\mathbb{Y}_{t}}) - \int_{0}^{t} \frac{1}{\mathbb{Y}_s}e^{-\rho s} dD_s) -\rho t} dD_t\bigg)^2\bigg]\\ &\quad \le\mathbb{Y}_0^{2R} \mathbb{E}\bigg[\bigg( \sup_{t\ge 0} \mathbb{Y}_{t\wedge\tau}^{1-R} \int_{0}^{\tau} e^{ - R \int_{0}^{t} \frac{1}{\mathbb{Y}_s}e^{-\rho s} dD_s}\frac{1}{\mathbb{Y}_t} e^{-\rho t} dD_t\bigg)^2\bigg]\\ &\quad\le \frac{\mathbb{Y}_0^{2R}}{R} \mathbb{E}\bigg[\sup_{t\ge 0} (V^{\text{EZ},D}_{t\wedge\tau})^2\bigg] <\infty, \end{align}\] as claimed. Hence, this completes the proof of ii..
Part i. follows by considering a non-increasing sequence \(\xi_{\tau}^n:= \frac{1}{n}\vee\xi_{\tau}\), \(n\in \mathbb{N}\), and then using the same arguments presented for Theorem 5i. ◻
Remark 10. As a counterpart to the dividend optimization \(V^{\text{EZ},*}\) under the EZ preference in 6 , the robust dividend optimization problem under the Maenhout’s preference is defined, for the set \(\widetilde{\@fontswitch\relax\mathcal{A}}\subseteq {\@fontswitch\relax\mathcal{A}}\) appearing in 6 , as \[\begin{align} \label{eq:opt95Vrob} V^{\text{rob},*}_0= \sup_{D \in \widetilde{\mathcal{A}}} V^{\text{rob},D}_0 = \sup_{D \in \widetilde{\mathcal{A}}} \inf_{\theta\in\Theta^D} V^{\theta,D}_0. \end{align}\tag{16}\] A direct consequence of Theorem 9 is that the two optimization problems \(V^{\text{EZ},*}\) and \(V^{\text{rob},*}_0\) are equivalent in the sense that \(V^{\text{EZ},*}=(V^{\text{rob},*})^{1-R}\).
Now, we consider the determination of an optimal robust dividend strategy under ambiguity in earning ability (take \({\@fontswitch\relax\mathcal{R}}\in(0,1)\)). We state a set of weak conditions under which the optimal robust dividend policy can be characterized across a wide range of surplus dynamics.
We start with the state space \(\mathcal{X}:= (\ell,\infty)\), \(-\infty \le \ell < 0\), and model the uncontrolled surplus process as \(\mathcal{X}\)-valued diffusion \(X^x=(X^x_t)_{t\ge 0}\) denote the solution of the stochastic differential equation (SDE) \[\begin{align} \label{eq:uncontrolX} X^x_t = x + \int_{0}^{t} \mu(X_s) ds + \int_{0}^{t}\sigma(X_s) dW_t, \quad t\ge 0, ~X_0 = x \in \mathcal{X}, \end{align}\tag{17}\] where functions \(\mu:\mathcal{X}\rightarrow \mathbb{R}\), \(\sigma:\mathcal{X}\rightarrow \mathbb{R}\backslash\{0\}\) are Borel functions. The value \(\mu(X_t)/X_t\) is the expected per surplus growth rate and \(\sigma(X_t)/X_t\) stands for infinitesimal volatility of fluctuations in the per surplus growth rate.
For each \(x\in \mathbb{R}_+\subset {\@fontswitch\relax\mathcal{X}}\) and \(D\in\mathcal{A}\), the dynamics of the controlled surplus process \(X^{x,D}:=(X_t^{x,D})_{t\ge 0}\) is governed by \[\begin{align} \label{eq:controlX} X_t^{x,D} = x + \int_{0}^{t} \mu(X_{s}^{x,D}) ds + \int_{0}^{t}\sigma(X_{s}^{x,D}) dW_t - D_t, \quad X_{0-}^{x,D} =x. \end{align}\tag{18}\] Throughout this section, for each \(x\in \mathbb{R}_+\subset {\@fontswitch\relax\mathcal{X}}\) and \(D\in\mathcal{A}\), we specify the primitive quantities \((\tau,\xi_{\tau})\) as \((\tau^{x,D},\xi_{\tau^{x,D}})\) which is defined as follows:
Definition 11.
Let \(\tau^{x,D}\in{\@fontswitch\relax\mathcal{T}}\) be the first hitting time of \(X^{x,D}\) at zero, i.e., \[\begin{align} \label{def:bankrupcytime} \tau^{x,D} := \inf\{t\ge0: X_{t}^{x,D} \le 0 \}. \end{align}\tag{19}\]
Let \(\xi_{\tau^{x,D}}\equiv (\xi_{\tau^{x,D}}^0)^{1-{\@fontswitch\relax\mathcal{R}}}\), with \(\xi^0_{\tau^{x,D}} = e^{-\rho \tau^{x,D}} \xi_0\) and some constant \(\xi_0>0\).
For each \(x \in \mathbb{R}_+\), we define \(\mathcal{A}^x \subseteq \mathcal{A}\) as the set of admissible dividend strategies: \[\begin{align} \mathcal{A}^x := \bigg\{ D \in \mathcal{A} : \int_{0}^{\tau^{x,D}} e^{-\rho s} dD_s \in L^2(\mathcal{F}_{\tau^{x,D}}),~D_t - D_{t-} \le X_{t-}^{x,D} \text{ for } t \ge 0 \bigg\}. \end{align}\] For each \(x\in \mathbb{R}_+\), \(D\in{\@fontswitch\relax\mathcal{A}}^x\), and \(\theta\in \Theta^D\), let us define \[\begin{align} J^{\theta}(x;D):=& \mathbb{E}^{\theta}\bigg[\int_{0}^{\tau^{x,D}}e^{-\rho t} \bigg(dD_t + \frac{J^{\theta}(X_t^{x,D};D) \theta_t^2 }{2{\@fontswitch\relax\mathcal{R}}}dt\bigg) + e^{-\rho \tau^{x,D}}\xi_0 \bigg]\nonumber\\ =& \mathbb{E}^{\theta}\bigg[\int_{0}^{\tau^{x,D}}e^{\frac{1}{2{\@fontswitch\relax\mathcal{R}}}\int_0^{t}\theta_s^2 ds-\rho t} dD_t + e^{\frac{1}{2{\@fontswitch\relax\mathcal{R}}}\int_0^{\tau^{x,D}}\theta_s^2 ds-\rho \tau^{x,D}}\xi_0 \bigg],\label{eq:object95J} \end{align}\tag{20}\] so that \(J^{\theta}(x; D)=V^{\theta,D}_0\) (see 10 ; with taking \(\mathbb{E}^\theta[\cdot]\) in \(V^{\theta,D}_0\) with \(X_0^{x,D}=x\)), where the last equality in 20 follows from Proposition 8.
Recall that \(\widetilde{\@fontswitch\relax\mathcal{A}} \subseteq {\@fontswitch\relax\mathcal{A}}\) appears in 6 and 16 . For any \(x \in \mathbb{R}_+\) and \(D \in {\@fontswitch\relax\mathcal{A}}^x\), \((\tau^{x,D}, D, \xi_{\tau^{x,D}})\) satisfies Condition 1. Thus, the robust dividend optimization problem here is a special case of 16 with \(\widetilde{\@fontswitch\relax\mathcal{A}} = {\@fontswitch\relax\mathcal{A}}^x\), i.e., for each \(x \in \mathbb{R}_+\), \[\begin{align} J^{*}(x) := \sup_{D\in \mathcal{A}^x} \inf_{\theta\in\Theta^D} J^{\theta}(x;D) = \sup_{D \in {\mathcal{A}^x}} V^{\text{rob},D}_0 = V^{\text{rob},*}_0. \label{eq:obj95maxmin} \end{align}\tag{21}\] The value functions are defined on \(\mathbb{R}_+\) (rather than \({\@fontswitch\relax\mathcal{X}}\)) due to bankruptcy: the value is trivially zero when \(x<0\).
We also assume the following standing assumption:
Condition 2. The SDE 17 admits a unique in law weak solution, which is a regular diffusion with unattainable boundaries \(\{\ell, \infty\}\). Moreover, there is a constant \(\overline{\mu}\) such that \(\sup_{x\ge 0}( \mu(x) -\rho x ) \le \overline{\mu}\).
Condition 2 asserts that the net appreciation rate of the surplus is uniformly bounded. This also ensures that the process \(X^x\) does not explode at infinity, and is regular at zero, so bankruptcy can occur as intended in our model. This further guarantees continuity of the optimal value function \(J^*\) on \(\mathbb{R}_+\), particularly at \(x=0\), which is crucial for the VI formulation (see 23 ) and the uniqueness of its solution.
Remark 12. In dividend optimization problems under model ambiguity, the regularity of the corresponding value function w.r.t. the state process is typically established only under specific models, such as constant-coefficient Brownian motion (Chakraborty et al. [39]). Related works without ambiguity, such as Choulli et al. [68] and Alvarez and Virtanen [65], assume boundedness of the coefficients of the Itô process to guarantee such regularity. Condition 2 aligns with such conditions, and it enables to extend the regularity results to general processes even under model ambiguity.
The regularity of \(J^*\) as well as its growth property is established in the following proposition. The proof is provided in Section 7.
Proposition 13. Under Condition 2, it holds that for \(x>0\), \[x+\xi_0 \le J^*(x) \le x+ \xi_0 + {\overline{\mu}}/{\rho},\] where \(\overline{\mu}\) is the constant that dominates function \(\mu(x)-\rho x\) in Condition 2. In addition, \(J^{*}\) is continuous and non-decreasing subject to \[\begin{align} &J^{*}(0+) = \xi_0. \label{eq:boundary95con95B} \end{align}\qquad{(3)}\]
In the rest of this section, we focus on candidate controls known as threshold dividend strategies. Following Cohen et al. [20] and Chakraborty et al. [39], we first provide a notion of a Skorokhod map on an interval. Denote by \(\mathcal{D}(\mathbb{R}_+)\) the the space of right-continuous functions with left limits mapping \(\mathbb{R}_+\) into \(\mathbb{R}\). Fix \(\beta>0\). Given \(\psi \in \mathcal{D}(\mathbb{R}_+)\) there exists a unique pair of functions \((\phi, \eta) \in \mathcal{D}(\mathbb{R}_+)^2\) that satisfy the following:
For every \(t \in\mathbb{R}_+, \phi(t)=\psi(t)+\eta(t)\);
\(\eta(0-)=0\), \(\eta\) is nondecreasing, and \(\int_0^{\infty} \mathbb{1}_{\{\phi(s)<\beta\}} d \eta(s)=0.\)
We define the map \(\Gamma_{\beta}[\psi]= (\Gamma_{\beta}^1,\Gamma_{\beta}^2)[\psi]:=(\phi, \eta)\). We refer to Kruk et al. [69] for the existence and uniqueness of \(\Gamma_\beta\) as well as its continuity.
Definition 14. Fix \(x,b\in\mathbb{R}_+\). We call \(D = D^{(b)}\) a \(b\)-threshold dividend strategy if for every continuous \(\psi\in\mathcal{D}(\mathbb{R}_+)\), one has \((X^{x,D},D)(\psi) =\Gamma_{b}[\psi]\).
One can deduce that any \(b\)-threshold control is admissible (i.e., \(D^{(b)}\in \mathcal{A}^x\)).
Since the optimal value function is continuous (see Proposition 13), this suggests that the threshold level \(b\) can be characterized by a dynamic programming principle. Motivated by the robust structure of objective function, together with the dynamic given in 18 , let \({\@fontswitch\relax\mathcal{L}}\) be the nonlinear operator with Hamiltonian over \(\theta\in \mathbb{R}\) which acts on \(v\in C^2({\mathbb{R}_+})\) as \[\begin{align} {\@fontswitch\relax\mathcal{L}}v(x):=&\inf_{\theta \in \mathbb{R}}\Big\{\frac{\sigma^2(x)}{2}v''(x)+\left(\mu(x)+\sigma(x) \theta\right)v'(x)+ \frac{\theta^2}{2{\@fontswitch\relax\mathcal{R}}} v(x) -\rho v(x) \Big\}\nonumber\\ =&\frac{\sigma^2(x)}{2}v''(x) + \mu(x)v'(x)-{\@fontswitch\relax\mathcal{R}}\frac{\sigma^2(x)}{2} \frac{(v'(x))^2}{v(x)}-\rho v(x),\label{eq:nonlinear95hamilton} \end{align}\tag{22}\] where the last equality is valid provided that \(v> 0\) on \({\mathbb{R}_+}\).
Then \(J^*\) in 21 is expected to be a solution of the following Hamilton-Jacobi-Bellman variational inequality (HJB-VI): \[\begin{align} \label{eq:HJB95VI} \left\{ \begin{aligned} &\max\{{\@fontswitch\relax\mathcal{L}}v(x),1-v'(x)\}=0\qquad on\;\;{\mathbb{R}_+}; \\ &v(0)=\xi_0. \end{aligned} \right. \end{align}\tag{23}\] We note that the boundary condition in 23 should be \(v(0+) = \xi_0\) as ?? shows, rather than \(v(0)= \xi_0\). Here and in what follows, we adopt a convention that the solution of the HJB-VI equation is extended to 0 by continuity and the value of \(v\) at 0 is its limit from the right.
Motivated by our educated guess that the optimal control is a threshold control, we consider an explicit formulation in terms of a free-boundary. Specifically, we aim to find \(b>0\) and \(v\in C^2({\mathbb{R}_+})\) such that \[\begin{align} \label{eq:fbdy} \left\{ \begin{aligned} &{\@fontswitch\relax\mathcal{L}}v(x)=0,\;\; v'(x)\geq 1\qquad on\;\;(0,b];\\ &{\@fontswitch\relax\mathcal{L}}v(x)\leq 0,\;\; v'(x)= 1\qquad on\;\;(b,\infty);\\ &v(0)=\xi_0, \end{aligned} \right. \end{align}\tag{24}\] as well as \(v''(b)=0\) and \(v'(b)=1\). The rationale behind this is as follows: when \(X_0^{x,D} = x\in (b,\infty)\), then in order to keep the process \(X^{x,D}\) between 0 and \(b\), there is an instantaneous dividend payment of size \(x-b\). When \(x\in(0,b)\), no action is being taken. When \(X^{x,D}\) hits the boundary \(b\), the threshold policy is taking action, leading to the Neumann boundary condition at \(b\). The surplus level will be kept in \((0,b)\), with an initial payment \(0\wedge(X_0^{x,D}-b)\) and then pay out dividend only when \(X^{x,D}\) is at level \(b\).
Assumption 15. \(\mu\) is in \(\in C^1({\mathbb{R}_+})\) with \(\mu(0)>0\). \(\sigma\) is in \(C^1({\mathbb{R}_+})\) and is non-decreasing. There exist \(\underline{b},\hat{b}, \overline{b}\in[0,\infty]\) with \(\underline{b} < \hat{b} < \overline{b}\) and the following conditions hold:
\(\mu(x)^2 \ge 3{\@fontswitch\relax\mathcal{R}}\rho\sigma(x)^2\) on \(x\in[0,\underline{b}]\) and there exists \[\overline{b} = \inf\{x\in(\underline{b},\infty): \mu(x)^2 = 2 {\@fontswitch\relax\mathcal{R}}\rho\sigma(x)^2\},\] with the convention that \(\inf\{\varnothing\} : =\infty\). In addition, \(\mu(x)^2 < 2 {\@fontswitch\relax\mathcal{R}}\rho\sigma(x)^2\) for all \(x>\overline{b}\).
Define \(\psi^\pm:[0,\overline{b}]\ni x\rightarrow \psi^\pm(x) \in \mathbb{R}\) by \(\psi^\pm(x):= \frac{\mu(x)\pm \sqrt{\mu^2(x)-2{\@fontswitch\relax\mathcal{R}}\rho \sigma^2(x)}}{2\rho}.\)
In particular, \(\psi^\pm\) are roots of \(Q(\psi;x):=\rho \psi^2 - \mu(x) \psi + \frac{{\@fontswitch\relax\mathcal{R}}}{2}\sigma^2(x)\). The map \(\psi^+(x)-x\) increases on \([0,\underline{b}]\) and decreases on \((\underline{b},\overline{b})\).
\(\sup_{x\in[0,\overline{b}]} \{\psi^-(x) -x \} < \xi_0 < \psi^+(\underline{b}) - \underline{b}\).
\(\hat{b}\) is the smallest value in \((\underline{b},\overline{b})\) that satisfies \(\psi^+(\hat{b})-\hat{b} = \xi_0\).
Remark 16. Assumption 15 extends the sufficient conditions for solvability and regularity of the optimal dividend problem in [65] to account for ambiguity. Below, we outline the economic implications in the assumption.
Assumption 15i. ensures that \(\mu(x)\) is sufficiently strong relative to \(\sigma(x)\) and \(\mathcal{R}\) in the low-reserve region. On the other hand, it also restricts \(\mathcal{R}\) in the sense that if \(\mathcal{R}\) is too large, immediate liquidation becomes optimal, reflecting extreme ambiguity aversion as discussed at the end of Section 5.
The term \(\rho(\psi^+(x)-x)\) in Assumption 15ii. represents the ambiguity-adjusted net appreciation rate, reducing to \(\mu(x)-\rho x\) when \(\mathcal{R}=0\) as in [65]. This condition requires the rate to be bounded, with dividends paid when it is decreasing, mirroring the classical case.
The condition on the bankruptcy payment \(\xi_0\) (in Assumption 15iii.) ensures that it is neither too low nor too high. This is necessary for a well-posed and meaningful optimal policy and prevents degenerate cases including immediate bankruptcy or indefinite reserve accumulation.
Let us provide some examples satisfying Assumption 15.
Example 17. Define the following two processes.
(Ornstein–Uhlenbeck process) The process \((X_t^x)_{t\geq 0}\) in 17 is given by \[\begin{align} d X_t = \overline{\gamma}(\overline{\mu}-X_t) dt + \overline{\sigma} dW_t,\quad x\in {\@fontswitch\relax\mathcal{X}}:=\mathbb{R}, \end{align}\] where \(\overline{\gamma}>0\), \(\overline{\mu}>0\) and \(\overline{\sigma}>0\) are constants satisfying \[\begin{align} (\overline{\gamma} \overline{\mu})^2 > 3{\@fontswitch\relax\mathcal{R}}\rho\overline{\sigma}^2, \qquad \sqrt{\frac{{\@fontswitch\relax\mathcal{R}}}{2\rho}} < \xi_0 < \frac{\overline{\gamma}\overline{\mu} + \sqrt{(\overline{\gamma}\overline{\mu})^2 -2{\@fontswitch\relax\mathcal{R}}\rho\overline{\sigma}^2}}{2\rho}. \end{align}\] Moreover, let \(\underline{b} = 0\) and \(\overline{b} = \overline{\mu} - \overline{\sigma}\sqrt{2{\@fontswitch\relax\mathcal{R}}\rho}/\overline{\gamma}\).
(Brownian motion with drift) The process \((X_t^x)_{t\geq 0}\) in 17 is given by \[\begin{align} d X_t = \mu dt + \sigma dW_t,\quad x\in {\@fontswitch\relax\mathcal{X}}:=\mathbb{R}, \end{align}\] where \(\mu\) and \(\sigma\) satisfy the following conditions \[\begin{align} \mu^2 > 3{\@fontswitch\relax\mathcal{R}}\rho\sigma^2, \qquad \frac{\mu - \sqrt{\mu^2 -2{\@fontswitch\relax\mathcal{R}}\rho\sigma^2}}{2\rho} < \xi_0 < \frac{\mu + \sqrt{\mu^2 -2{\@fontswitch\relax\mathcal{R}}\rho\sigma^2}}{2\rho}. \end{align}\] Moreover, let \(\underline{b} = 0\) and \(\overline{b} = \overline{\mu} - \overline{\sigma}\sqrt{2{\@fontswitch\relax\mathcal{R}}\rho}/\overline{\gamma}\).
Both processes in i. and ii. satisfy Assumption 15. Case i. represents a mean-reverting surplus model considered by [12].
We now prove that an optimal dividend strategy exists and that it is a threshold control. Moreover, we show that the threshold level and the value function \(J^*\) are characterized by the free-boundary problem 24 . The proofs of the following theorems are presented in Sections 8 and 9.
Theorem 18. Suppose that Condition 2 and Assumption 15 hold. There exists a unique solution \(v^* : \mathbb{R}_+ \to \mathbb{R}_+\) with \(v^* \in C^2(\mathbb{R}_+)\) to the free boundary problem 24 , and the associated free boundary \(b^*\) is unique and lies in \((\underline{b}, \hat{b})\), as defined in Assumption 15.
Theorem 19. Suppose that Condition 2 and Assumption 15 hold. Let \(v^*\) be the unique solution to the free boundary problem 24 with corresponding free boundary \(b^*\) as established in Theorem 18, and let \(J^*\) be defined in 21 . Furthermore, let \(D^*:=D^{({{b}^*})}\) be the \(b^*\)-threshold dividend strategy (see Definition 14). Then, for every \(x \in \mathbb{R}_+\), \[v^*(x) = J^*(x) = \inf_{\theta \in \Theta^{D^*}} J^{\theta}(x; D^*).\] Moreover, \(v^*(x)\) satisfies \[\begin{align} v^*(x) = \sup_{D \in \mathcal{A}^x} \inf_{\theta \in \Theta^D} \mathbb{E}^{\theta} \bigg[ \int_{0}^{\tau^{x,D}} e^{-\rho t} \left( dD_t + \frac{\theta_t^2}{2{\@fontswitch\relax\mathcal{R}}} v^*(X_t^{x,D}) dt \right) + e^{-\rho \tau^{x,D}} \xi_0 \bigg]. \end{align}\] Thus, \(v^*\) coincides with \(J^*\), and \(D^*\) is the optimal robust dividend strategy.
In this section, we analyze the sensitivity of the robust singular criterion and its optimization with respect to the ambiguity aversion parameter \({\@fontswitch\relax\mathcal{R}}\in (0,1)\). All the proof of results are presented in Section 10. Our analysis builds on the following specification which aligns with the primitives given in Definitions 1 and 11 in the previous sections.
Setting 20.
Let \((\tau,D)\in {\@fontswitch\relax\mathcal{T}}\times {\@fontswitch\relax\mathcal{A}}\) be such that \(\int_0^\tau e^{-\rho s}dD_s\in L^2({\@fontswitch\relax\mathcal{F}}_\tau)\).
For every \(R\equiv{\@fontswitch\relax\mathcal{R}}\in(0,1)\) (as in Section 3), let \(\xi_\tau({\@fontswitch\relax\mathcal{R}}):=(\xi_\tau^0)^{1-{\@fontswitch\relax\mathcal{R}}}\), with some \(\xi_\tau^0\in L^2({\@fontswitch\relax\mathcal{F}}_\tau)\) satisfying \(\xi_{\tau}^0>0\) \(\mathbb{P}\)-a.s., so that \(\xi_\tau({\@fontswitch\relax\mathcal{R}})\in L^1({\@fontswitch\relax\mathcal{F}}_\tau)\) and \((\tau,D,\xi_\tau({\@fontswitch\relax\mathcal{R}}))\) satisfies Condition 1.
For every \({\@fontswitch\relax\mathcal{R}}\in (0,1)\), let \(V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}):=(V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}))_{t\geq 0}\) be the Maenhout’s robust singular control process defined as in 9 under setting \((\tau,D,\xi_\tau({\@fontswitch\relax\mathcal{R}}))\) given in i. and ii. (which is well-defined by Proposition 8).
The following theorem shows the sensitivity analysis of \(V^{\text{rob},D}\) w.r.t. \({\@fontswitch\relax\mathcal{R}}\).
Theorem 21. For every \(t \ge 0\), the mapping \((0,1) \ni {\@fontswitch\relax\mathcal{R}}\mapsto V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\in \mathbb{R}_+\) is decreasing and continuous. In addition, for every \(t\geq0\), \(\mathbb{P}\)-a.s., \[\begin{align} \label{eq:Vrob95lim} \lim_{{\@fontswitch\relax\mathcal{R}}\downarrow 0} V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}) = K_t^D\quad\text{and}\quad V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\le K_t^D, \end{align}\qquad{(4)}\] where \(K^D\) is the classical (ambiguity-neutral) criterion in 1 with \((\tau,D,\xi_\tau^0)\).
Remark 22. We note that \(V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\) in 9 is not a priori well-defined at \({\@fontswitch\relax\mathcal{R}}= 0\). However, by ?? , we are able to extend its definition to \({\@fontswitch\relax\mathcal{R}}=0\) by setting \(V_t^{\text{rob},D}(0) := K^D_t\) for \(t\in[0,T]\). Moreover, Theorem 9 ensures that \(V^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}}) = (V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}))^{1-{\@fontswitch\relax\mathcal{R}}}\) for all \({\@fontswitch\relax\mathcal{R}}\in [0,1)\).
We state a corollary on \(J^*\) in Section 4 by noting that \(J^*(x;{\@fontswitch\relax\mathcal{R}}) = V^{\text{rob},*}_0({\@fontswitch\relax\mathcal{R}})\) as in 21 . By Remark 22, we extend the definition of \(J^*(x;\cdot)\) at \({\@fontswitch\relax\mathcal{R}}=0\) by setting \[\begin{align} \label{dfn:robust95J950} J^*(x;0) := \sup_{D \in \mathcal{A}^x} \mathbb{E} \bigg[ e^{-\rho \tau^{x,D}}\xi_0 + \int_0^{\tau^{x,D}} e^{-\rho s} \, dD_s \bigg]. \end{align}\tag{25}\]
Corollary 23. Under Definition 11, for any \(x \in \mathbb{R}_+\), the mapping \([0,1) \ni {\@fontswitch\relax\mathcal{R}}\mapsto J^*(x;{\@fontswitch\relax\mathcal{R}})\) is decreasing, lower semi-continuous and right continuous.
If the conditions of Theorems 18 and 19 are imposed, it is possible to obtain (strong) continuity of \(J^*\) w.r.t \({\@fontswitch\relax\mathcal{R}}\). In particular, this also yields continuity of the optimal threshold, denoted by \(b^*_{{\@fontswitch\relax\mathcal{R}}}\) for each \({\@fontswitch\relax\mathcal{R}}\in[0,1)\).
Theorem 24. Let \(x \in \mathbb{R}_+\) and let \({\@fontswitch\relax\mathcal{I}}\subset[0,1)\) be a closed sub-interval. If Condition 2 and Assumption 15 hold for every \({\@fontswitch\relax\mathcal{R}}\in {\@fontswitch\relax\mathcal{I}}\), then the mappings \({\@fontswitch\relax\mathcal{I}}\ni{\@fontswitch\relax\mathcal{R}}\mapsto J^*(x;{\@fontswitch\relax\mathcal{R}})\in \mathbb{R}_+\) and \({\@fontswitch\relax\mathcal{I}}\ni{\@fontswitch\relax\mathcal{R}}\mapsto b^*_{{\@fontswitch\relax\mathcal{R}}}\in \mathbb{R}_+\) are continuous, i.e., \[\lim_{\substack{|{\@fontswitch\relax\mathcal{R}}_2 - {\@fontswitch\relax\mathcal{R}}_1| \to 0 \\ {\@fontswitch\relax\mathcal{R}}_1,\,{\@fontswitch\relax\mathcal{R}}_2 \in {\@fontswitch\relax\mathcal{I}}}} \left| J^*(x;{{\@fontswitch\relax\mathcal{R}}_1}) - J^*(x;{{\@fontswitch\relax\mathcal{R}}_2}) \right| = 0, \quad \lim_{\substack{|{\@fontswitch\relax\mathcal{R}}_2 - {\@fontswitch\relax\mathcal{R}}_1| \to 0 \\ {\@fontswitch\relax\mathcal{R}}_1,\,{\@fontswitch\relax\mathcal{R}}_2 \in {\@fontswitch\relax\mathcal{I}}}} \left| b^*_{{\@fontswitch\relax\mathcal{R}}_1} - b^*_{{\@fontswitch\relax\mathcal{R}}_2} \right| = 0.\]
Remark 25. The monotonicity and right-continuity results given in Theorem 21 and Corollary 23 align with sensitivity analysis of robust optimizations in continuous time setting, established in Bartl et al. [70], [71], Herrmann and Muhle-Karbe [72], Herrman et al. [73] (under a small ambiguity aversion, i.e., \({\@fontswitch\relax\mathcal{R}}\ll 1\)). Although our paper addressed the well-posedness and sensitivity of the robust singular criterion for small ambiguity aversion, the case for \({\@fontswitch\relax\mathcal{R}}\in (1,\infty)\) suggests interesting avenues for future research.
Let us introduce the following sets: for any \(p\geq1\) and \(t\geq0\),
\(\mathcal{S}_p\) is the set of the set of all real-valued, \(\mathbb{F}\)-progressively measurable processes \(Y\) such that \(\|Y\|_{{\@fontswitch\relax\mathcal{S}}_p}^p:=\mathbb{E}[\sup _{t \geq 0}|Y_{t \wedge \tau}|^p]<\infty\);
\({\@fontswitch\relax\mathcal{S}}_{\infty}\) is the set of all real-valued, \(\mathbb{F}\)-progressively measurable processes \(Y\) such that \(\|Y\|_{{\@fontswitch\relax\mathcal{S}}_\infty}:=\inf\{C\geq 0: |Y_t| \leq C\;for all\;t\geq0\;\mathbb{P}-a.s.\}\);
\(\mathcal{M}_p\) is the set of all real-valued, \(\mathbb{F}\)-progressively measurable processes \(Z\) such that \(\|Z\|_{\mathcal{M}_p}^p:=\mathbb{E}[(\int_0^\tau |Z_t|^2 dt)^{p/2}]<\infty\);
Denote by \(\mathcal{B}_p:=\mathcal{S}_{p} \times \mathcal{M}_p\) the product space equipped with the norm \(\|(Y,Z) \|_{\mathcal{B}_p}^p: =\|Y\|_{{\@fontswitch\relax\mathcal{S}}_p}^p+ \|Z\|_{\mathcal{M}_p}^p\) for every \((Y,Z)\in\mathcal{B}_p\).
We outline the main steps as follows. First, we prove existence of a BSDE solution for any bounded \(D \in \mathcal{A}\) with a Lipschitz aggregator (Step 1). Next, we establish a comparison theorem for aggregators nonincreasing in their second argument, yielding uniqueness (Step 2), and derive a priori bounds for solutions with the non-Lipschitz EZ aggregator. Using these bounds, we approximate the EZ aggregator by a sequence of Lipschitz functions and show convergence of the corresponding BSDE solutions (Step 3). Finally, for general unbounded \(D\), we apply a technique from Aurand and Huang [16] (Step 4).
For any \(0<T<\infty\), we define the spaces \({\@fontswitch\relax\mathcal{S}}_{2}([0,T])\), \(\mathcal{M}_2([0,T])\), and \(\mathcal{B}_2([0,T])\) similarly, with \(\tau\) replaced by \(T\) in the definition of \({\@fontswitch\relax\mathcal{S}}_2\), \({\@fontswitch\relax\mathcal{M}}_2\), and \({\@fontswitch\relax\mathcal{B}}_2\).
Lemma 26. Let \(T<\infty\). Let \(A\in {\@fontswitch\relax\mathcal{S}}_{2}([0,T])\) and \(D\in\mathcal{A}\) be such that \(D_T \le K\) \(\mathbb{P}\)-a.s. for some constant \(K\), and both \(A\) and \(D\) have continuous paths. If the following holds for every \(\zeta \in {\@fontswitch\relax\mathcal{T}}\) such that \(\zeta \leq T\) \(\mathbb{P}\)-a.s., \[\begin{align} \mathbb{E}[A_{\zeta}] \le \alpha + \mathbb{E}\bigg[\int_{\zeta}^{T} A_s dD_s\bigg], \end{align}\] with some constant \(\alpha\), then it holds that \(\mathbb{E}[A_t] \le \alpha e^K\) for every \(t\in[0,T]\).
Proof. For fixed \(t\in[0,T)\) and \(0\le k\le K\), set \(\zeta_k = \inf\{s \ge t: D_s \ge k\}\wedge T\). Then each \(\zeta_k\) is a stopping time bounded by \(T\). Moreover, since \(\{\zeta_k\le s\} = \{D_s\ge k\}\) for \(s\ge t\) and \(\zeta_{K} = T\) holds \(\mathbb{P}\)-a.s., it follows that \[\begin{align} \mathbb{E}[A_{\zeta_k}] \le& \alpha + \mathbb{E}\bigg[ \int_{\zeta_k}^{T} A_s d D_s\bigg] = \alpha + \mathbb{E}\bigg[ \int_{t}^{T} A_s \mathbb{1}_{s\ge \zeta_k}d D_s\bigg]\\ \le & \alpha + \mathbb{E}\bigg[ \int_{t}^{T} A_s \mathbb{1}_{ \zeta_k \le s \le \zeta_{K}}d D_s\bigg] = \alpha + \int_{k}^K \mathbb{E}[A_{\zeta_s}] ds. \end{align}\] Set \(u_s:= \mathbb{E}[A_{\zeta_s}]\) for \(k\in[0,K]\). Then since \(u_k \le \alpha + \int_{k}^K u_s ds\), for \(k\in[0,K]\), the standard (backward) Grönwall’s inequality ensures that \(u_k \le \alpha e^{K-k}\) for \(k\in[0,K]\). We complete the proof by letting \(k=0\). ◻
Theorem 27 (Lipschitz aggregator). Consider a fixed horizon \(0<T < \infty\) and a process \(D\in {\@fontswitch\relax\mathcal{A}}\) such that \(\int_0^{T}e^{-\rho t}d D_t \le K\) \(\mathbb{P}\)-a.s. for some positive constant \(K\). Let \(\xi_T \in L^2(\mathcal{F}_T)\). Assume that \(g\) satisfies the following conditions:
\(\exists\) \(C_g>0\) s.t. \(|g(t,y_1)-g(t,y_2)|\le C_g e^{-\rho t}|y_1-y_2|\) \(\forall y_1,y_2\).
\(\int_0^{T} |g(t,0)|dD_t \in L^2({\@fontswitch\relax\mathcal{F}}_T)\).
Then the BSDE with parameter \((T,\xi_{T},g,D)\) has a unique solution in \(\mathcal{B}_2\).
Proof. The idea is to extend the Picard iteration method given by Pardoux and Peng [74] and then to show that the corresponding sequence is Cauchy in \(\mathcal{B}_2\).
Let \(y^0 = 0\), and \(\{ (y^n_t,z^n_t)_{t\in[0,T]}\}_{n\ge1}\) be a sequence in \(\mathcal{B}_2\) defined recursively as \[\begin{align} y_t^{n+1} = \xi_T+ \int_{t}^T g(s,y_s^n)dD_s - \int_{t}^T z_s^{n} dW_s ,\quad 0\le t \le T. \end{align}\] For any \(y^n\in \mathcal{S}_2\), we construct \(y^{n+1}\) and \(z^n\) as follows. From [item:A1] and [item:A2], \[\begin{align} &\mathbb{E}\bigg[\bigg(\xi_T+\int_0^{T} g(s,y^n_s)d D_s \bigg)^2 \bigg] \le 2 \mathbb{E}\bigg[\xi_T^2+ \bigg(\int_0^{T} \big(|g(s,0)| + C_g e^{-\rho s}|y^n_s|\big)d D_s \bigg)^2 \bigg]\\ &\quad \le2\mathbb{E}[\xi_T^2] + 4 \mathbb{E}\bigg[\int_0^{T} |g(s,0)|dD_s\bigg]^2+ C_g \mathbb{E}\bigg[\sup_{0\le t\le T}|y^n_t|^2 \bigg]K^2<\infty. \end{align}\] This implies that \(\{ \mathbb{E}_t[\xi_T+\int_0^{T} g(s,y^n_s)d D_s ]\}_{t\ge0}\) is a square integrable martingale. By the martingale representation theorem, we construct a unique \(z^{n}\in\mathcal{M}_2\) and \(y^{n+1}\in\mathcal{S}_2\) such that for \(0\le t\le T\) \[\begin{align} &\int_0^t z^{n}_s dW_s = \mathbb{E}_t\bigg[\xi_T+\int_0^{T} g(s,y^n_s)d D_s \bigg] - \mathbb{E}\bigg[\xi_T+\int_0^{T} g(s,y^n_s)d D_s \bigg],\nonumber\\ &\text{ and }y^{n+1}_t: = \mathbb{E}_t\bigg[\xi_T+\int_t^{T} g(s,y^n_s)d D_s \bigg]. \end{align}\] For \(\tau \in \mathcal{T}\), applying Itô’s lemma to \((u_{s\wedge\tau}^{n+1})^2: = |y_{s\wedge\tau}^{n+1} - y_{s\wedge\tau}^{n}|^2\) gives \[\begin{align} &(u_{t\wedge\tau}^{n+1})^2 + \int_{{s\wedge\tau}}^{T\wedge\tau}| z_s^{n}- z_s^{n-1}|^2 ds \\ &= 2\int_{{s\wedge\tau}}^{T\wedge\tau} \big(g(s,y_s^n)- g(s,y_s^{n-1})\big) u_{s}^{n+1}dD_s - 2 \int_{{s\wedge\tau}}^{T\wedge\tau} u_s^{n+1}(z_s^{n}- z_s^{n-1}) dW_s. \end{align}\] It is clear by Aurand and Huang [16] that the above stochastic integral is a uniformly integrable martingale and has zero expectation. By taking expectations on both sides, together with condition [item:A1] it follows that \[\begin{align} &\mathbb{E} \big[(u_{t\wedge\tau}^{n+1})^2\big] + \mathbb{E}\bigg[\int_{t\wedge\tau}^{T\wedge\tau}| z_s^{n}- z_s^{n-1}|^2 ds\bigg]\\ &\quad \le 2\mathbb{E}\bigg[\int_{t\wedge\tau}^{T\wedge\tau} C_g u_{s}^n u_{s}^{n+1}e^{-\rho s}d{D}_s\bigg] \\ &\quad \le C_g \mathbb{E}\bigg[\int_{t\wedge\tau}^{T\wedge\tau} (u_{s}^n)^2 e^{-\rho s}d{D}_s\bigg] + C_g \mathbb{E}\bigg[\int_{t\wedge\tau}^{T\wedge\tau} (u_{s}^{n+1})^2 e^{-\rho s}d{D}_s\bigg]. \end{align}\] Lemma 26 implies \[\begin{align} \label{eq:thmA463pf951} \mathbb{E}[(u_t^{n+1})^2] \le C_g e^{C_g K} \mathbb{E}\bigg[\int_{t}^T (u_{s}^n)^2 e^{-\rho s} d{D}_s\bigg],\quad t\in[0,T]. \end{align}\tag{26}\] Consider a sequence stopping time: \(\zeta_{s}:=\inf\{t\ge0: \int_0^t e^{-\rho \tau}D_{\tau}\ge s \}\wedge T\) for \(s\in[0,K]\), which is bounded. From the proof of Lemma 26, it is obvious that \[\begin{align} \label{eq:thmA463pf952} \mathbb{E}[(u_{\zeta_t}^{n+1})^2] \le C_g e^{C_g K} \mathbb{E}\bigg[\int_{\zeta_t}^T (u_{s}^n)^2 e^{-\rho s} d{D}_s\bigg] \end{align}\tag{27}\] holds for \(t\in[0,K]\). Iterating 26 and 27 gives for \(t\in[0,T]\), \[\begin{align} \mathbb{E}[(u_t^{n+1})^2] &\le C_g e^{C_g K} \mathbb{E}\bigg[\int_{t}^T (u_{s}^n)^2 e^{-\rho s} d{D}_s\bigg]\nonumber\\ &\le C_g e^{C_g K} \mathbb{E}\bigg[\int_{0}^T (u_{s}^n)^2 e^{-\rho s} \mathbb{1}_{ \zeta_0 \le s \le \zeta_K} d{D}_s\bigg]\nonumber\\ &= C_g e^{C_g K} \mathbb{E}\bigg[\int_{0}^{K} (u_{\zeta_s}^n)^2 ds \bigg] = C_g e^{C_g K} \int_{0}^{K} \mathbb{E}[ (u_{\zeta_s}^n)^2] ds\nonumber\\ &\le (C_g e^{C_g K})^2 \int_{0}^{K} \mathbb{E}\bigg[\int_{\zeta_s}^T (u_{t}^{n-1})^2 e^{-\rho t} d{D}_t\bigg] ds\nonumber\\ &= (C_g e^{C_g K})^2 \int_{0}^{K} \int_{t_1}^{K} \mathbb{E}[ (u_{\zeta_{t_2}}^{n-1})^2] d t_2 d t_1 \nonumber\\ &\cdots\nonumber\\ &\le (C_g e^{C_g K})^n \int_{0}^{K} \int_{t_1}^{K}\cdots \int_{t_{n-1}}^K \mathbb{E}[ (u_{\zeta_{t_n}}^{1})^2] d t_{n} dt_{n-1}\cdots d t_1\nonumber\\ &\le (C_g e^{C_g K})^n \mathbb{E}\bigg[\sup_{t\in[0,T]}(u_t^1)^2\bigg] \frac{K^n}{n!}.\nonumber \end{align}\] The above estimate implies that \(\{(y^n,z^n)\}\) is a Cauchy sequence in \(\mathcal{S}_2\times\mathcal{M}_2\), due to the fact that \(y^1\in\mathcal{S}_2\) and so \(\mathbb{E}[\sup_{t\in[0,T]}(u_t^1)^2]\) is finite. It follows that \(\lim_{n\rightarrow\infty}(y^n,z^n) = (y,z) \in \mathcal{B}_2\). By standard arguments, we show that \((y,z)\) solves BSDE 5 with \((T,\xi_{T},g,D)\). The uniqueness of \((y,z)\) follows standard arguments with comparison theorem (see Proposition 29). ◻
Definition 28. Let \((\tau, D,\xi_{\tau})\) satisfy Condition 1. Adapted processes \((Y_t,Z_t)_{t\ge 0}\) are called a supersolution (resp. subsolution) to 5 with \((\tau,\xi_{\tau},g,D)\) if \[\begin{align} Y_t = \xi_{\tau} + \int_t^{\tau} g(s,Y_s)dD_s - \int_t^{\tau}Z_s dW_s + \int_t^{\tau} dA_s\quad \bigg(\text{resp.}-\int_{t}^{\tau} dA_s\bigg), \end{align}\] where \((A_t)_{t\ge0}\) is a right continuous left limit process in \({\@fontswitch\relax\mathcal{P}}\), and it holds that \((Y_t+ \int_0^{t \wedge\tau}g(s,Y_s)dD_s)_{t\ge 0}\) is a local-supermartingale (resp. -submartingale). A solution of 5 with \((\tau,\xi_{\tau},g,D)\) is both a supersolution and a subsolution.
The following result shows a comparison result for 5 without restricting ourselves to the conditions given by Theorem 27. As it is a direct consequence of Kobylanski [75], we omit the proofs here.
Proposition 29. Let \((Y,Z)\) (resp. \((\widetilde{Y},\widetilde{Z})\)) be a supersolution (resp. subsolution) to 5 with \((\tau,\xi_{\tau},g,D)\) (resp. \((\tau,\tilde{\xi}_{\tau},\tilde{g},D)\)). Assume that both \((Y,Z)\) and \((\widetilde{Y},\widetilde{Z})\) are of class \(\mathcal{B}_2\) and one of the following conditions hold:
\(\tilde{g}(t,Y_t) \le g(t,Y_t)\) and \(\tilde{g}(t,y)\) is nonincreasing in \(y\), \(d\mathbb{P}\times dt\)-a.e., or
\(\tilde{g}(t,\widetilde{Y}_t) \le g(t,\widetilde{Y}_t)\) and \(g(t,y)\) is nonincreasing in \(y\), \(d\mathbb{P}\times dt\)-a.e..
If \(\xi_{\tau} \ge \tilde{\xi}_{\tau}\), then \(\widetilde{Y}_t \le Y_t\) for any \(t \ge 0\) \(\mathbb{P}\)-a.s.
We introduce two auxiliary utility processes that act as lower and upper bounds for the solutions of our target BSDE. The well-defined nature of these processe is established in the following lemmas.
Lemma 30. Let \((\tau, D,\xi_{\tau})\) satisfy Condition 1. There exists a process \(\overline{Y}\in {\@fontswitch\relax\mathcal{P}}\) such that \(\mathbb{E}[\sup_{t\ge 0}\overline{Y}_t^{\frac{2}{1-R}}]<\infty\) (in particular \((\overline{Y}_t)_{t\ge 0}\in\mathcal{S}_2\)) given by \[\begin{align} \overline{Y}_t: = \bigg(\mathbb{E}_t\bigg[ \int_t^{\tau} e^{-\rho s} dD_s + (\xi_{\tau})^{\frac{1}{1-R}}\bigg]\bigg)^{1-R}, \quad t \ge 0, \end{align}\] where \(R\in(0,1)\) is the risk-aversion coefficient in \(g_{\text{EZ}}\) defined in 3 .
Proof. Observe that \((\overline{Y})^{\frac{1}{1-R}}\) is the classical dividend-paying utility process, it follows directly that \((\overline{Y})^{\frac{1}{1-R}}\) is the unique solution to BSDE with parameter \((\tau, (\xi_{\tau})^{\frac{1}{1-R}}, \overline{g},D)\), with \(\overline{g}(t,v)=e^{-\rho t}\). Since \((\tau, D,\xi_{\tau})\) satisfies Condition 1, \(\mathbb{E}[\sup_{t\ge 0}\overline{Y}_t^{\frac{2}{1-R}}]<\infty\) by Doob’s maximal inequality. \(\overline{Y}\in\mathcal{S}_2\) follows as a consequence of Jensen’s inequality. ◻
The triplet \((T,D,\xi_T)\) satisfying conditions in Theorem 27 is a special case of the above Lemma.
Lemma 31. For the triplet \((T,D,\xi_T)\) satisfying conditions in Theorem 27, there exists \((\underline{Y}_t)_{t\in[0,T]}\in\mathcal{S}_2\) satisfying \[\begin{align} \underline{Y}_t: = \mathbb{E}_t\bigg[ \int_t^{T} \underline{g}(s,\underline{Y}_s) dD_s + \xi_T \bigg],\quad \underline{g}(t,v) =e^{-\rho t}(1-R v),\quad t\in[0,T]. \end{align}\] In addition, \(\underline{Y}_t \ge e^{-R \|\int_0^T e^{-\rho s} dD_s \|_{\infty}} \mathbb{E}_t[\xi_T]\) \(\mathbb{P}\)-a.s., for all \(t\in[0,T]\).
Proof. Since aggregator \(\underline{g}(\cdot,y)\) satisfies condition in Theorem 27, we obtain the existence and uniqueness of the solution to BSDE 5 with parameters \((\xi,\underline{g},T,D)\). It follows from the comparison theorem (Proposition 29) that \(\underline{Y}_t \ge \tilde{Y}_t\), where \((\tilde{Y},\tilde{Z})\) is the unique solution to BSDE \[\begin{align} \tilde{Y}_t = \xi_T - \int_t^{T} e^{-\rho s} R \tilde{Y}_s dD_s - \int_{t}^{\tau} \tilde{Z}_t dW_s. \end{align}\] Moreover, since \(\tilde{Y}_t = \mathbb{E}_t[e^{-R \int_{t}^T e^{-\rho s}dD_s}\xi_T ] \ge e^{-R \|\int_0^T e^{-\rho s} dD_s \|_{\infty}} \mathbb{E}_t[\xi_T]\) \(\mathbb{P}\)-a.s., this completes the proof. ◻
We start by approximating the EZ aggregator \(g_{\text{EZ}}\) in 3 using Lipschitz aggregators that satisfy conditions [item:A1] and [item:A2]. The proof of the construction of approximation is straightforward, so we omit it in the following lemma.
Lemma 32. The sequence of aggregators \[\begin{align} g_n(\omega,t,y) = \inf_{x > 0 }\{g_{\text{EZ}}(\omega,t,x) + n R e^{-\rho t}(x-y)\} \end{align}\] is well-defined for each \(n\ge 1\), satisfying both [item:A2] and the following:
Monotonicity in \(n\): \(\forall~y \ge 0\), \(g_n(\omega,t,y)\) increase in \(n\), \(d\mathbb{P}\times dt\)-a.s.;
Lipshcitz condition [item:A1]: \(|g_n(\omega,t,y_1)-g_n(\omega,t,y_2)|\le nR e^{-\rho t}|y_1-y_2|\) for all \(y_1, y_2\ge 0\), \(d\mathbb{P}\times dt\)-a.s.;
Monotonicity in \(y\): \(g_n\) is nonincreasing in \(y\), \(d\mathbb{P}\times dt\)-a.s..
Theorem 33. For \(\mathbb{E}[(\xi_{T})^{\frac{2}{1-R}}]<\infty\) and \((T,D)\) satisfying conditions in Theorem 27, further assume that \(\xi_{T}>C\) \(\mathbb{P}\)-a.s. for some constant \(C\). The BSDE 5 with parameter \((T,\xi_T,g_{\text{EZ}},D)\) has a unique solution \((y_t,z_t)_{t\in[0,T]}\in\mathcal{B}_2\).
Proof. For fixed \((\omega,t)\), consider the sequence \(g_n(\omega,t,y)\) associated with \(g_{\text{EZ}}\) in 3 constructed by Lemma 32. It follows by Theorem 27 that, for each \(n\ge1\), the BSDE with parameter \((T,\xi_T,g_n,D)\) has a unique solution \((y_t^n,z_t^n)_{t\in[0,\tau]} \in \mathcal{B}_2\). Lemma 32 [item:prop_agg_i] and [item:prop_agg_iii], Proposition 29 yield that for \(n\ge1\), \(y_t^1(\omega)\le y_t^{n}(\omega)\le y_t^{n+1}\), \(d\mathbb{P}\times dt\)-a.s. Note that \(y^1\) is identical to \(\underline{Y}\) defined in Lemma 31 (with current triplet \((T,D,\xi_T)\)), since \(g_1\) and \(\underline{g}\) are identical. Under the condition that \(\xi_{T}\in L^2(\mathcal{F}_T)_+\) and \(\int_0^T e^{-\rho s} dD_s\le K\) a.s., by Lemma 31, there is a real constant \(C_{\xi}\) such that \(y^1_t = \underline{Y}_t \ge e^{-R K} \mathbb{E}_t[\xi_T] \ge C_{\xi}\) for \(t\in[0,T]\) a.s.
We show that \((y^n)_{n\in \mathbb{N}}\) is bounded from above by \(\overline{Y}\) as defined in Lemma 30 (with current triplet \((T,D,\xi_T)\)). For \(n\ge1\), a direct application of Itô’s formula (cf. Jacod and Shiryaev [76]) to \(\psi:=(y^n)^{\frac{1}{1-R}}\) yields \[\begin{align} d\psi_t =& -\frac{\psi_t^{R} g_n(t,\psi_t^{1-R}) }{1-R}dD_t + \frac{(\psi_{t})^{R}z_t^n }{1-R}dW_t + \frac{R}{2(1-R)^2}(\psi_{t})^{2R-1}|z_t^n|^2 dt. \end{align}\] Note that \(\int_0^{\cdot} \frac{R}{2(1-R)^2}(\psi_{t})^{2R-1}|z_t^n|^2 dt\) is an increasing process, so \(\psi\) is a subsolution of 5 with \((T,(\xi_T)^{\frac{1}{1-R}},h,D)\), where \(h(t,y) = y^{R} g_n(t,y^{1-R})/(1-R )\). Since \(g_n\le g_{\text{EZ}}\), we have \(y^{R}g_n(t,y^{1-R})/(1-R) \le y^{R}g_{\text{EZ}}(t,y^{1-R})/(1-R) = e^{-\rho t}\). Proposition 29 yields that \(\psi \le (\overline{Y})^{\frac{1}{1-R}}\) so that \(y^n \le \overline{Y}\).
Now for each \(n\ge1\), \(C_{\xi} \le y_t^1(\omega)\le y_t^{n}(\omega)\le y_t^{n+1}\le \overline{Y}_t\), \(d\mathbb{P}\times dt\)-a.s., there exists an \(\mathcal{F}_t\)-progressively measurable process \(y\) satisfying \(\lim_{n\rightarrow\infty} y_t^n(\omega) = y_t(\omega)\), \(d\mathbb{P}\times dt\)-a.s. Hence, \(\mathbb{E}[\sup_{s\in[0,T]}|y_s|^2] \le \mathbb{E}[\sup_{s\in[0,T]}|\overline{Y}_s|^2]<\infty.\) In order to take limit of \((z^n_t)_{t\in[0,T]}\), we derive the following estimate by applying Itô’s lemma to \((y_t^n)^2\): \[\begin{align} \mathbb{E}\bigg[ \int_{0}^{T} |z^n_s|^2 ds\bigg]&= \mathbb{E}[\xi_T^2] + 2 \mathbb{E}\bigg[\int_0^{T} y_s^ng_n(s,y_s^n)d D_s \bigg] \nonumber \\ &\le \mathbb{E}[\xi_T^2] + 2 \mathbb{E}\bigg[\int_0^{T} y_s^ng_{\text{EZ}}(s,y_s^1)d D_s \bigg] \\ &\le \mathbb{E}[\xi_T^2]+ 2C_{\xi}^{\frac{-R}{1-R}} \mathbb{E}\bigg[\sup_{s\in[0,T]}|\overline{Y}_s|^2\bigg] \bigg\|\int_0^T e^{-\rho s} dD_s \bigg\|_{\infty} <\infty, \nonumber \end{align}\] where we use the fact that \(0\le g^{n}(t, y^n) \le g_{\text{EZ}}(t,y^n) \le g_{\text{EZ}}(t,y^1) \le C_{\xi}^{\frac{-R}{1-R}}e^{-\rho t}\) for any \(n\) and \(t\in[0,T]\). Therefore, there exists \(z\in\mathcal{M}_2\) and a sub sequence \((z^{n_j})_j\) of \((z^n)_n\) such that \(z^{n_j} \rightharpoonup z\) weakly in \(\mathcal{M}_2.\)
Passing to a subsequence if necessary, we can show that the whole sequence \((z^n)_n\) converges strongly to \(z\) in \(\mathcal{M}_2\) by Lebesgue’s DCT. Moreover, \(y\) is a continuous process because \(y^n\) has continuous paths by construction. We can now pass the limit in \[\begin{align} y_t^n = \xi_T + \int_{t}^T g_n(s,y^n_s)d D_s - \int_t^T z_s^n dW_s, \end{align}\] so that \((y_t,z_t)_{t\in[0,T]}\in\mathcal{B}_2\) is a solution to BSDE with \((T,\xi_T,g_{\text{EZ}},D)\). ◻
Theorem 34. For \((\tau, D,\xi_{\tau})\) satisfying Condition 1, assume that there is a real constant \(C>0\) such that \(\xi_{\tau}>C\) \(\mathbb{P}\)-a.s. Then the BSDE 5 with parameter \((\tau,\xi_{\tau},g_{\text{EZ}},D)\) has a unique solution \((Y_t,Z_t)_{t\ge 0}\in\mathcal{B}_2\). In addition, it holds that \(Y\) is continuous, \(Y\ge C\) \(\mathbb{P}\)-a.s. and \(\mathbb{E}[\sup_{t\ge 0} (Y_t)^{\frac{2}{1-R}}]<\infty\).
Proof. We aim to extend the result in Theorem 33 to the case with random horizon and unbounded controls.
For each \(n\in\mathbb{N}\), we construct a solution \((Y_t^n,Z_t^n)_{t\ge0} \in \mathcal{B}_2\) to the BSDE \[\begin{align} \label{eq:Y95n95construct} Y_t^n = \xi_{\tau} + \int_{t \wedge {\tau}}^{{\tau}} \mathbb{1}_{s\in[0,n\wedge\tau^n]} g_{\text{EZ}}(s,Y_s^n)dD_s - \int_{t \wedge {\tau}}^{{\tau}} Z_s^n dW_s, \quad t\ge0, \end{align}\tag{28}\] where \(\tau^n\in\mathcal{T}\) is defined as \(\tau^n: = \inf\{t\ge 0: \int_0^{t}e^{-\rho s}d D_s \ge n \}\). We denote by \(\overline{\tau}^n := {\tau}\wedge\tau^n\). Theorem 33 implies that there exits a unique solution \((y_t,z_t)_{t\in[0,n]}\in\mathcal{B}_2\) for fixed horizon \([0,n]\) BSDE \[\begin{align} y_t = \mathbb{E}_n[ \xi_{\tau}] + \int_{t}^{n} \mathbb{1}_{s\in[0,\overline{\tau}^n]} g_{\text{EZ}}(s,y_s)dD_s - \int_t^n z_s dW_s. \end{align}\] The conditions in Theorem 33 are satisfied since \(\int_{0}^n \mathbb{1}_{s\in[0,\overline{\tau}^n]} e^{-\rho s}dD_s \le n\) a.s. On the other hand, under the condition that \(\xi_{\tau}\in L^2_+(\mathcal{F}_{\tau})\), martingale representation theorem implies that there exists \((\eta_t)_{t\ge 0} \in \mathcal{M}_2\) such that \[\begin{align} \mathbb{E}_{t}[\xi_{\tau}] = \xi_{\tau}- \int_t^{{\tau}} \eta_s dW_s,\text{ on }\{t<{\tau}\};\quad \eta_t = 0 \text{ on }\{t >{\tau} \}. \end{align}\] So we construct \((Y_t^n,Z_t^n)_{t\ge0}\in \mathcal{B}_2\) as \[\begin{align} Y_t^n = y_t \mathbb{1}_{t\in[0,n\wedge\overline{\tau}^n]} + \mathbb{E}_{t}[\xi_{\tau}]\mathbb{1}_{t\in(n\wedge\overline{\tau}^n,\infty)},\quad Z_t^n = z_t \mathbb{1}_{t\in[0,n\wedge\overline{\tau}^n]} + \eta_t\mathbb{1}_{t\in(n\wedge\overline{\tau}^n,\infty)}. \end{align}\] Next, we aim to show that \((Y^n,Z^n)_{n\in\mathbb{N}}\) is Cauchy in \(\mathcal{B}_2\).
For any \(m>n\), let \(\Delta Y_t: =Y_t^m - Y_t^n\), \(\Delta Z_t : =Z_t^m-Z_t^n\). Then it suffices to show that \(|(\Delta Y, \Delta Z)|_{\mathcal{B}_2}\rightarrow 0\) as \(m,n\rightarrow \infty\).
For \(0\le t \le n\), by 28 and the fact that \(\tau^{n}(\omega)\le \tau^{m}(\omega)\) we have \[\begin{align} &\Delta Y_{t}-\Delta Y_{n\wedge {\tau}}\\ &\quad = \int_{t\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^n} \Delta g_{\text{EZ}}^{m,n}(s)dD_s + \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} g_{\text{EZ}}(s,Y_s^m) dD_s- \int_{t\wedge {\tau}}^{n\wedge {\tau}} \Delta Z_s dW_s, \end{align}\] where \(\Delta g_{\text{EZ}}^{m,n}(s): =g_{\text{EZ}}(s,Y_s^m) - g_{\text{EZ}}(s,Y_s^n)\). It follows from Itô’s lemma that \[\begin{align} &|\Delta Y_{t\wedge{\tau}}|^2 + \int_{t\wedge{\tau}}^{n\wedge {\tau}} |\Delta Z_s|^2 ds = 2 \int_{t\wedge\overline{\tau}^n}^{n\wedge \overline{\tau}^n} \Delta Y_s \Delta g_{\text{EZ}}^{m,n}(s) dD_s \\ &\quad + 2 \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} \Delta Y_s g_{\text{EZ}}(s,Y_s^m) dD_s- 2 \int_{t\wedge{\tau}}^{n\wedge{\tau}} \Delta Y_s \Delta Z_s dW_s +| \Delta Y_{n\wedge {\tau}}|^2 . \end{align}\] By the fact that \(g_{\text{EZ}}\) is non-increasing in its second argument, so that \(\Delta g_{\text{EZ}}^{m,n}(s)\) should have opposite sign of \(\Delta Y_s\). This, together with the non-decreasing property of \(D\), gives \(\int_{t\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^n} \Delta Y_s \Delta g_{\text{EZ}}^{m,n}(s)dD_s\le0\). \[\begin{align} &\int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} \Delta Y_s g_{\text{EZ}}(s,Y_s^m) dD_s = \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} \Delta Y_s \big(\Delta g_{\text{EZ}}^{m,n}(s)+ g_{\text{EZ}}(s,Y^n_s)\big) dD_s \\ &\quad \le \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} \Delta Y_s g_{\text{EZ}}(s,Y^n_s) dD_s = \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} \Delta Y_s g_{\text{EZ}}(s,\mathbb{E}_s[\xi]) dD_s\\ &\quad \le 2 \sup_{t\in[0,{\tau}]}|\overline{Y}_t| C_4^{\frac{-R}{1-R}} \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} e^{-\rho s} dD_s , \end{align}\] where the second line follows by construction that \(Y^n_s = \mathbb{E}_{s}[\xi]\) for \(s>n\wedge\tau^n\), the last line follows by the fact that \(\xi_{\tau} \ge C_4\) \(\mathbb{P}\)-a.s. for some constant \(C_4>0\), as well as the priory bound \(Y^m\), \(Y^n\le \overline{Y}\) given in Lemma 30. Hence, \[\begin{align} &|\Delta Y_{t\wedge{\tau}}|^2 + \int_{t\wedge {\tau}}^{n\wedge {\tau}} |\Delta Z_s|^2 ds\\ &\quad \le | \Delta Y_{n\wedge {\tau}}|^2 +2 \sup_{t\in[0,{\tau}]}|\overline{Y}_t| C_4^{\frac{-R}{1-R}} \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} e^{-\rho s} dD_s - 2\int_{t\wedge {\tau}}^{n\wedge {\tau}} \Delta Y_s \Delta Z_s dW_s, \end{align}\] where the above local martingale is a uniformly integrable martingale. Hence, \[\begin{align} \mathbb{E}\bigg[\int_{0}^{n\wedge {\tau}} |\Delta Z_s|^2 ds \bigg] &\le \mathbb{E}[ | \Delta Y_{n\wedge {\tau}}|^2] + 2 C_4^{\frac{-R}{1-R}} \mathbb{E}\bigg[ \sup_{t\in[0,{\tau}]}|\overline{Y}_t| \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m} e^{-\rho s} dD_s\bigg] \nonumber \\ &=: \mathbb{E}[ | \Delta Y_{n\wedge {\tau}}|^2] + 2C_4^{\frac{-R}{1-R}} \operatorname{I}^{n,m}. \label{eq:temp95est95YD} \end{align}\tag{29}\] Moreover, it holds that \[\begin{align} \operatorname{I}^{n,m}&\leq \mathbb{E}\bigg[\sup_{t\in[0,{\tau}]}|\overline{Y}_t|^2\bigg]^{\frac{1}{2}} \mathbb{E}\bigg[\bigg( \int_{n\wedge \overline{\tau}^n}^{n\wedge \overline{\tau}^m}e^{-\rho s} dD_s \bigg)^2\bigg]^{1/2}\leq C_5(m-n).\label{eq:temp95est95YD951} \end{align}\tag{30}\] where the constant \(C_5>0\) (not depending on \(m\),\(n\),\(\Delta Y\),\(\Delta Z\)) exists because \(D\in{\@fontswitch\relax\mathcal{A}}\) has continuous path. In a similar manner, we have that \[\begin{align} &\mathbb{E}\bigg[ \sup_{t\in[0,n]} |\Delta Y_{t\wedge {\tau}}|^2 \bigg] \le \mathbb{E}[ | \Delta Y_{n\wedge {\tau} }|^2]+2C_4^{\frac{-R}{1-R}} \operatorname{I}^{n,m}+ 2\mathbb{E}\bigg[\sup_{t\in[0,n]}\int_{0}^{t\wedge {\tau}} \Delta Y_s \Delta Z_s dW_s \bigg]\nonumber\\ &\quad =: \mathbb{E}[ | \Delta Y_{n\wedge {\tau} }|^2]+2C_4^{\frac{-R}{1-R}} \operatorname{I}^{n,m}+2 \operatorname{II}^{n,m},\label{eq:temp95est95YD952} \end{align}\tag{31}\] where \(\operatorname{I}^{n,m}\) is given in 29 , and that \[\begin{align} \operatorname{II}^{n,m} &\leq C_6 \mathbb{E}\bigg[\bigg(\int_{0}^{n\wedge {\tau}} |\Delta Y_s|^2 |\Delta Z_s|^2 ds \bigg)^{1/2} \bigg]\nonumber \\ &\leq \frac{1}{4}\mathbb{E}\Big[ \sup_{t\in[0,n]} |\Delta Y_{t\wedge {\tau}}|^2 \Big] + C_6^2 \mathbb{E}\bigg[\int_{0}^{n\wedge {\tau}} |\Delta Z_s|^2 ds \bigg],\label{eq:temp95est95YD953} \end{align}\tag{32}\] where the first inequality follows from the inequality of arithmetic and geometric means, and the constant \(C_6>0\) (not depending on \(m\),\(n\),\(\Delta Y\),\(\Delta Z\)) exists by the Burkholder-Davis-Gundy’s inequality. Combining 29 , 30 , 31 , and 32 , we have some constant \(C_7>0\) (that does not depend on \(m\),\(n\)) such that \[\begin{align} \mathbb{E}\bigg[ \sup_{0\le t\le n}| \Delta Y_{t\wedge {\tau}}|^2+\int_{0}^{n\wedge {\tau}} |\Delta Z_s|^2 ds \bigg] \le C_7 \Big(\mathbb{E}\big[ | \Delta Y_{n\wedge {\tau}}|^2\big]+(m-n)\Big). \end{align}\] Next, for \(n<t \le m\), we have \[\begin{align} \Delta Y_{t} = \int_{t \wedge \overline{\tau}^n}^{m\wedge \overline{\tau}^m} g_{\text{EZ}}(s,Y_s^m)dD_s - \int_{t\wedge {\tau}}^{m\wedge {\tau}} {\Delta}Z_s dW_s, \end{align}\] where the equality holds by the fact that \(\Delta Z_s = 0\) on \(s > m\) since \(Z^m_s = Z^n_s =\eta_s\) for \(s>m\wedge{\tau} \ge m \wedge \overline{\tau}^m\).
It then follows from Itô’s lemma and similar arguments as above that \[\begin{align} &|\Delta Y_{t\wedge{\tau}}|^2 + \int_{t\wedge {\tau}}^{m\wedge{\tau}} |{\Delta}Z_s|^2 ds\\ &\quad = 2 \int_{t\wedge \overline{\tau}^m}^{m\wedge\overline{\tau}^m} \Delta Y_s g_{\text{EZ}}(s,Y^m_s) dD_s - 2\int_{t\wedge {\tau}}^{m\wedge {\tau}} \Delta Y_s {\Delta}Z_s dW_s\\ &\quad \le 2 \int_{t\wedge \overline{\tau}^m}^{m\wedge\overline{\tau}^m} |\Delta Y_s| g_{\text{EZ}}(s,\mathbb{E}_s[\xi_{\tau}]) dD_s - 2\int_{t\wedge {\tau}}^{m\wedge {\tau}} \Delta Y_s {\Delta}Z_s dW_s. \end{align}\] Note that \(n\wedge \overline{\tau}^m \le t\wedge \overline{\tau}^m \le m\wedge \overline{\tau}^m\), therefore similar argument as for the case \(0\le t \le n\) gives the existence of \(C_8>0\) independent of \(m\), \(n\) such that \[\begin{align} \mathbb{E}\bigg[ \sup_{n\le t\le m}| \Delta Y_{t\wedge {\tau}}|^2+\int_{n \wedge {\tau}}^{m \wedge {\tau}} |\Delta Z_s|^2 ds \bigg] \le C_8 (m-n). \end{align}\] Finally, \(Y_t^m = Y_t^n = \mathbb{E}_{t\wedge{\tau}}[\xi_{\tau}]\) and \(Z_t^m = Z_t^n = \eta_t\) for \(t>m\wedge{\tau}\ge m\wedge\overline{\tau}^m\), so \[\begin{align} \label{eq:mn95cauchy951} \mathbb{E}\bigg[ \sup_{t\ge 0}| \Delta Y_{t\wedge {\tau}}|^2+\int_{0}^{{\tau}} |\Delta Z_s|^2 ds \bigg] \le C_7 \mathbb{E}\big[ | \Delta Y_{n\wedge {\tau}}|^2\big] + C_9(m-n), \end{align}\tag{33}\] with \(C_9:=(C_7+C_8)\). By the DCT, we have \[\begin{align} \lim\limits_{n,m\rightarrow\infty} \mathbb{E}\big[ | \Delta Y_{n\wedge {\tau}}|^2\big] &= \mathbb{E}\Big[ \lim\limits_{n,m\rightarrow\infty}|Y^m_{n\wedge {\tau}} - Y^n_{n\wedge {\tau}}|^2 \Big] \\ &=\mathbb{E}\Big[ \lim\limits_{n\rightarrow\infty}|Y_{n\wedge\tau}-Y^n_{n\wedge {\tau}}|^2\Big] = \mathbb{E}\big[ | \xi_{\tau} - \xi_{\tau}|^2\big] = 0. \end{align}\] Since the second term in 33 vanishes as \(m,n\rightarrow\infty\), we conclude that \(\|(\Delta Y,\Delta Z)\|_{\mathcal{B}_2} \rightarrow 0\) as \(m\), \(n\rightarrow\infty\), i.e., \((Y^n,Z^n)_n\) is Cauchy in \(\mathcal{B}_2\). Since \(\mathcal{B}_2\) is complete, \(\lim_{n\rightarrow\infty}(Y^n,Z^n) = (Y,Z)\in\mathcal{B}_2\) exists. Finally, we show the limit \((Y,Z)\in\mathcal{B}_2\) solves BSDE 5 with parameter \((\tau,\xi_{\tau},g_{\text{EZ}},D)\). For each \(n\in\mathbb{N}\), \((Y_t^n,Z_t^n)_{t\ge0} \in \mathcal{B}_2\) solves the BSDE 28 , \[\begin{align} Y_t^n = \xi_{\tau} + \int_{t \wedge {\tau}}^{{\tau}} \mathbb{1}_{s\in[0,n\wedge\tau^n]} g_{\text{EZ}}(s,Y_s^n)dD_s - \int_{t \wedge {\tau}}^{{\tau}} Z_s^n dW_s, \quad t\ge0. \end{align}\] It is not difficult to pass the limit \(n\rightarrow\infty\) in the above equation and show that each term of it converges to a corresponding term in 5 for almost all \(\omega\in\Omega\) uniformly in \(t\ge0\), thus we omit the details here. ◻
We start by showing that the value function \(J^*\) is bounded.
Lemma 35. Under Condition 2, we have \(\xi_0 \le J^*(x) \le x+ \xi_0 + {\overline{\mu}}/{\rho}\), \(x\in\mathbb{R}_+\), where \(\overline{\mu}\) is the constant that dominates function \(\mu(x)-\rho x\) in Condition 2.
Proof. Define the process \(\theta^0: = (\theta^0(t)\equiv 0 )_{t\ge0}\). It is easy to verify that \(\theta_0 \in \Theta\) and the induced measure \(\mathbb{Q}^{\theta^0}\) coincides with the reference measure \(\mathbb{P}\). It then follows by definition of \(J^*\) in 21 that \(J^*(x)\le \sup_{D\in \mathcal{A}_x} \mathbb{E}[ \int_{0}^{\tau^{x,D}} e^{-\rho t}dD_t ] + \xi_0.\) Moreover, we observe that \[\begin{align} \mathbb{E}\bigg[ \int_{0}^{\tau^{x,D}} e^{-\rho t}dD_t \bigg] = x + \mathbb{E}\bigg[\int_{0}^{\tau_N}e^{-\rho t} \big(\mu(X_t^D)-\rho X_t^D\big)dt- e^{-\rho \tau_N} X^D_{\tau_N} \bigg], \end{align}\] where \(\tau_N: = \tau^{x,D}\wedge N \wedge \inf\{t\ge0: X_t^D \ge N \}\) is an almost surely finite stopping time. Further invoking the requirement that admissible \(D\) keep the controlled process \(X^{x,D}\) non-negative, we get the following inequality of \(J^*\) \[\begin{align} J^*(x)\le x + \xi_0 + \sup_{D\in \mathcal{A}_x} \mathbb{E}\bigg[ \int_{0}^{\tau^{x,D}} e^{-\rho t}\big(\mu(X_t^D)-\rho X_t^D\big)^+ dt \bigg] \le x+ \xi_0 + \frac{\overline{\mu}}{\rho}, \end{align}\] where \((y)^+ = \max(y,0)\). The proof is complete. ◻
Next, we present a technical result for uncontrolled diffusion \(X\) in 17 in order to establish the boundary condition ?? for optimal value function. Formally we define the following stopping time for \(x\in \mathcal{X}\) \[\begin{align} \label{dfn:zeta95y124x} \zeta_{y|x} = \inf\{t\ge 0, X_t^x = y \}, \quad y\in\mathcal{X} \cup \{-\ell,\infty\}. \end{align}\tag{34}\] Condition 2 guarantees that \(\mathbb{P}(\zeta_{0|0} = 0) = 1\), \(\mathbb{P}(\zeta_{0|x}<\infty) = 1\), and \(\mathbb{P}(\zeta_{\infty|x} <\infty) =0\), for all \(x\in\mathcal{X}/\{0\}\).
Lemma 36. Under Condition 2, for any \(\delta>0\), we have \[\begin{align} \label{eq:lem95bound950} \lim_{x\rightarrow h} \mathbb{P}(\zeta_{h|x}<\delta) = 1, \end{align}\qquad{(5)}\] where stopping time \(\zeta_{y|x}\) is defined in 34 . In particular, for fixed \(y>0\), \[\begin{align} \lim_{x \downarrow 0} \mathbb{P}(\zeta_{0|x} < \delta) = 1, \quad \lim_{x\downarrow 0} \mathbb{E}\Big[\max_{0\le s\le \delta\wedge{\zeta}_{0|x}\wedge{\zeta}_{y|x}} X^x_s\Big] = 0.\label{eq:lem95bound952} \end{align}\qquad{(6)}\]
Proof. The proof involves a transformed process \(U_t\) of \(X_t\), which is a time-changed Brownian motion, and we prove that auxiliary versions of ?? , ?? hold with the process \(U_t\). We introduce the scale function of \(X\) in 17 as \[\begin{align} \mathfrak{s}(x): = \int_{0}^{x}\exp\Big(-2\int_0^z\frac{\mu(y)}{\sigma^2(y)}dy\Big)dz, \end{align}\] which is strictly increasing, continuously differentiable bijection of the interval \(\mathcal{X}\) onto the interval \(I=(\iota_{l},\iota_{\infty})\) with endpoints \(\iota_l:= \mathfrak{s}(\ell+)\) and \(\iota_{\infty}:= \mathfrak{s}(\infty-)\); see, e.g., Borodin and Salminen [77]. Denote by \(\mathfrak{s}^{-1}\) the inverse mapping of \(\mathfrak{s}\). We define \(U_t : = \mathfrak{s}(X_t)\), with state space \(I\), and its dynamic is given by \[\begin{align} \label{pf:lem:X95bound950} U_t = \mathfrak{s}(X_0) + \int_0^t \varsigma(U_s) dW_s, \end{align}\tag{35}\] where the dispersion function \(\varsigma(u): = (\mathfrak{s}'\cdot\sigma)(\mathfrak{s}^{-1}(u))\); see e.g., Karatzas and Shreve [78]. To further remove the dependency on the function \(\varsigma\) in 35 , we consider a time change for transformation. According to the Dambis–Dubins–Schwarz theorem (see Revuz and Yor [79]), one can rewrite the stochastic integral \(\int_{0}^{\cdot} \varsigma(U_s) dW_s\) of Brownian motion \(W\) as a time-changed Brownian motion on (an extension of) the probability space, i.e., \(\hat{W}_{\Lambda^U(\cdot)}\) that we define as follows. Define a time scale \(\Gamma^U(r)\) for \(r \in{\mathbb{R}_+}\) as \[\begin{align} \Gamma^{U}(r):=\int_0^{r} \frac{dt}{\varsigma^2\big(\mathfrak{s}(X_0)+\hat{W}_t\big)} \end{align}\] in terms of a standard Brownian motion \(\hat{W}\), and denote by \(\Lambda^{U}\) the inverse mapping of \(\Gamma^{U}\), i.e. \(\Lambda^U(t): = \inf\{r\ge 0: \Gamma^U(r)>t\}\). The representation \[\begin{align} \label{pf:lem:X95bound951} U_t = \mathfrak{s}(X_0) + \hat{W}_{\Lambda^{U}(t)} \end{align}\tag{36}\] holds. Due to the strict monotonicity of the scale function \(\mathfrak{s}\), it is sufficient to prove auxiliary properties of ?? ?? for the process \(U\) given by 36 . For convenience of presentation, we adopt the notation \[\begin{align} &\zeta^U_{y|x} :=\inf\{t\ge 0: U_t = \mathfrak{s}(y) \text{ given }U_0 =\mathfrak{s}(x)\}. \end{align}\] From the continuous path properties of standard Brownian motion together with the fact that \(\Lambda^{U}(t)>0\) holds for all \(t>0\), it is proved (see e.g.[77]) \[\begin{align} &\mathbb{P}\Big(\big \{\omega\in\Omega: \exists \epsilon(\omega)>0 \text{ s.t. } U_t \le \mathfrak{s}(x),\forall t\in[0,\epsilon(\omega)] \big \} \big| X_0 = x \Big)=0,\\ & \mathbb{P}\Big(\big \{\omega\in\Omega: \exists \epsilon(\omega)>0 \text{ s.t. } U_t \ge \mathfrak{s}(x),\forall t\in[0,\epsilon(\omega)] \big \} \big| X_0 = x \Big)=0. \end{align}\] Given the above statements, we have for any fixed \(\delta>0\), \[\begin{align} &\mathbb{P}( \zeta^U_{a|x} \ge \delta) = \mathbb{P}\Big(\inf_{t\in[0,\delta]} U_t \ge \mathfrak{s}(a) \big| X_0=x \Big) \rightarrow \mathbb{P}\Big(\inf_{t\in[0,\delta]} U_t > \mathfrak{s}(x) \big| X_0=x \Big) = 0,\\ & \mathbb{P}( \zeta^U_{b|x} \ge \delta) = \mathbb{P}\Big(\sup_{t\in[0,\delta]} U_t \ge \mathfrak{s}(b) \big| X_0=x \Big) \rightarrow \mathbb{P}\Big(\sup_{t\in[0,\delta]} U_t < \mathfrak{s}(x) \big| X_0=x \Big) = 0, \end{align}\] as \(a\uparrow x\) and \(b\downarrow x\) respectively. Hence, we conclude ?? .
To show ?? , we recognize from the time change property that \[\begin{align} \max_{0\le s\le \delta\wedge{\zeta}^U_{0|x}\wedge{\zeta}^U_{y|x}} U_s &= \max_{0\le r \le \Lambda^{U}(\delta)\wedge\Lambda^{U}(\zeta^U_{0|x})\wedge\Lambda^{U}(\zeta^U_{y|x})} \mathfrak{s}(x)+ \hat{W}_r\\ &= \max_{0\le r \le \Lambda^{U}(\delta)\wedge\hat{\zeta}_{0|x}\wedge\hat{\zeta}_{y|x}} \mathfrak{s}(x)+ \hat{W}_r, \end{align}\] where \(\hat{\zeta}_{y|x} = \inf\{ r\ge 0: B_x(r)= \mathfrak{s}(y) \}\), with \(B_x(r):= \mathfrak{s}(x)+ \hat{W}_r\).
By the monotonicity of \(\mathfrak{s}\), \(B_x(t,\omega)\) decreases as \(x\downarrow 0\). By the law of iterated logarithm (see Revuz and Yor [79]), we have \(\hat{\zeta}_{0|x}\rightarrow0\) as \(x\downarrow 0\). By definition, \(\Lambda^{\Upsilon}(\delta)\) is finite a.s. for \(\delta\in[0,\infty)\). Since \(B_x(r)\) has continuous path, considering a decreasing sequence \((x_n)_{n\in\mathbb{N}}\) with \(x_n<y\) and \(\lim_{n\rightarrow\infty} x_n=0\), we have \(\lim_{n\rightarrow\infty} \max_{0\le r \le \Lambda^{U}(\delta)\wedge\hat{\zeta}_{0|x_n}\wedge\hat{\zeta}_{y|x_n}} B_{x_n}(r) = \mathfrak{s}(0),\) and \(\max_{0\le r \le \Lambda^U(\delta)\wedge\hat{\zeta}_{y|x}}B_x(r)\le \mathfrak{s}(y)\); hence ?? holds. ◻
Now we turn to the proof of Proposition 13.
Proof of Proposition 13.. The boundedness of \(J^*\) follows by Lemma 35. We first show the non-decreasing property of \(J^*\). If \(y>x\), then for any admissible pair \((D,\theta)\in\mathcal{A}^x \times \Theta^D\), we can put \(\hat{D}_t = D_t + y-x\), corresponding to immediately paying dividends in the amount of \(y-x\),
thereby instantly changing the initial reserve from \(y\) to \(x\) and then following strategy \(D\). Obviously, \(\hat{D}\in\mathcal{A}^y\), and that \(J^{\theta}(x;\hat{D}) = J^{\theta}(x;D) + (y-x)\), it follows that \[J^*(y) \ge J^*(x)+ (y-x),\] thus, \(J^*\) is nondecreasing.
Next, we prove ?? . Fix \(\delta>0\) and \(y>0\), in view of Lemma 36, for any \(\epsilon\in(0,1)\), choose \(x>0\) such that \[\begin{align} \mathbb{P}(\zeta_{0|x} < \delta) \ge 1-\epsilon,\qquad \mathbb{E} \bigg[\max_{0\le s\le
\delta\wedge{\zeta}_{{0|x}}\wedge\zeta_{y|x}} X_s\bigg] \le \epsilon.
\end{align}\] Let \(\hat{\tau} = \tau^{x,D} \wedge \zeta_{y|x} \wedge \delta\), where \(\tau^{x,D}\) defined in 19 . Since \(0\le X^{x,D}\le X\), we have \(\tau^{x,D}\le \zeta_0\) and \(\mathbb{P}(\tau^{x,D} <\zeta_{y|x} \wedge \delta) \ge \mathbb{P}(\zeta_{0|x} <\zeta_{y|x} \wedge
\delta) \ge 1-\epsilon\). Due to the requirement that \(- X^D_t \ge 0\) for all \(t\ge 0\), we have and \(D_{\hat{\tau}}\le X_{\hat{\tau}} \le \max_{0\le
s\le \hat{\tau}}X_s\le \max_{0\le s\le \delta\wedge{\zeta}_{{0|x}}\wedge\zeta_{y|x}} X_s\). Therefore \(\mathbb{E}_x[D_{\hat{\tau}}]\le \epsilon\). Consider the objective function \(J^{\theta^0}(x;D)\) in 20 with \(\theta^0= (\theta^0(t)\equiv 0 )_{t\ge0}\), we have \[\begin{align} &J^{\theta^0}(x;D) =
\mathbb{E}\Big[\int_0^{\hat{\tau}} e^{-\rho t} dD_t + \mathbb{1}_{\hat{\tau}<\tau^{x,D}}\int_{\hat{\tau}}^{\tau^{x,D}} e^{-\rho t} dD_t + \xi e^{-\rho \tau^{x,D}} \Big]\\ &\le \mathbb{E}[D_{\hat{\tau}}] + \mathbb{E}\bigg[
\mathbb{1}_{\hat{\tau}<\tau^{x,D}} \mathbb{E}\bigg[\int_{\hat{\tau}}^{\tau^{x,D}} e^{-\rho t} dD_t \bigg| \mathcal{F}_{\hat{\tau}} \bigg]\bigg] + \xi \\ &\le \epsilon + \mathbb{E}[ \mathbb{1}_{\hat{\tau}<\tau^{x,D}} J^*(y)] + \xi \le \epsilon +
J^*(y)\mathbb{P}(\hat{\tau}<\tau^{x,D})+\xi \le \xi + \big(1+J^*(y)\big)\epsilon.
\end{align}\] Due to the arbitrariness of \(\epsilon\), \(\inf_{\theta}J^{\theta}(0+;\hat{D})\le J^{\theta^0}(0+;\hat{D}) \le \xi\), thus by arbitrariness of \(D\), \(J^*(0+) \le \xi\); together with \(J^*(0+) \ge \xi\) by Lemma 35, we
conclude the validity of ?? .
Lastly, we prove that \(J^*\) is continuous at any \(x>0\). For arbitrary \(\theta\in\Theta\), let the measure \(\mathbb{Q}^{\theta}\) be defined in 8 and Brownian motion \(W^{\theta}\) under measure \(\mathbb{Q}^{\theta}\). Under \(\mathbb{Q}^{\theta}\), the uncontrolled surplus \(X\) satisfies \[\begin{align} dX_t = \big(\mu(X_t) + \sigma(X_t)\theta(X_t) \big)dt + \sigma(X_t) dW^{\theta}_t,\quad
X_0 = x.
\end{align}\] For arbitrary \(x>0\), define the stopping time \(\zeta^{\mathbb{Q}}_{0,y|x}:=\inf\{t\ge0: X_t \notin (0,y)\}\) under measure \(\mathbb{Q}\). It is well-known from the classical theory of diffusion that \[\begin{align}
\label{pf:prop95v95cts951} \mathbb{Q}(X_{\zeta^{\mathbb{Q}}_{0,y|x}} = y) = \frac{\mathfrak{s}^{\mathbb{Q}}(x) - \mathfrak{s}^{\mathbb{Q}}(0)}{ \mathfrak{s}^{\mathbb{Q}}(y) - \mathfrak{s}^{\mathbb{Q}}(0) },
\end{align}\tag{37}\] where \(\mathfrak{s}^{\mathbb{Q}}\) is the scale function of \(X_t\) under measure \(\mathbb{Q}\), that is,
\[\mathfrak{s}^{\mathbb{Q}}(x) : = \int_{0}^{x}\exp\bigg(-2\int_0^z\frac{\mu(y)+\sigma(y)\theta(y)}{\sigma^2(y)}dy\bigg)dz.\] 37 implies that for arbitrary \(\epsilon>0\), there exists \(y({\epsilon}) > x\) such that \[\mathbb{Q}(X_{\zeta^{\mathbb{Q}}_{0,y|x}} = y) \ge 1-\epsilon~for~x< y<y(\epsilon).\]
Then, following similar arguments for proving ?? , we have for any \(\epsilon>0\), there exists \(y'(\epsilon)>x\) such that \(\mathbb{Q}(\zeta^{\mathbb{Q}}_{y|x}<\zeta^{\mathbb{Q}}_{0|x}\wedge \epsilon) \ge 1-\epsilon\) for \(x<y<y'(\epsilon)\). For arbitrary \(D\in
\mathcal{A}^y\) with \(x<y<y'(\epsilon)\), we construct a dividend strategy for surplus process \(X\) with \(X_0 = x\) as follows: \(D'_t = 0\) when \(t< \zeta^{\mathbb{Q}}_{0,y|x}\wedge\epsilon\) and for \(t\ge \zeta^{\mathbb{Q}}_{0,y|x}\wedge\epsilon\), \(D'_t = X_{\epsilon}\) on \(\{\epsilon< \zeta^{\mathbb{Q}}_{0,y|x}\}\), \(D'_t = D_{t-\zeta^{\mathbb{Q}}_{y|x}}\) on \(\{\zeta^{\mathbb{Q}}_{y|x}<\epsilon\wedge\zeta^{\mathbb{Q}}_{0|x}\}\), \(D_t \equiv 0\) otherwise. We have for \(x<y<y'(\epsilon)\), \[\begin{align} &J^{\theta}(x,{D}') = \mathbb{E}^{\theta} \bigg[ \int_{0}^{\tau^{x,D'}} e^{\int_{0}^{s} \frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du -\rho s} dD'_s + e^{\int_{0}^{\tau^{x,D'}}
\frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}-\rho du}\xi_{\tau^{x,D'}}\bigg] \\ &\ge \mathbb{E}^{\theta} \bigg[ \mathbb{1}_{\zeta^{\mathbb{Q}}_{y|x}<\zeta^{\mathbb{Q}}_{0|x}}\bigg(\int_{0}^{\zeta^{\mathbb{Q}}_{y|x}} e^{\int_{0}^{s}
\frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du -\rho s} dD_s \\ &\qquad\qquad\qquad\qquad\quad +\int_{\zeta^{\mathbb{Q}}_{y|x}}^{\zeta^{\mathbb{Q}}_{0|y}} e^{\int_{0}^{s} \frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du -\rho s} dD_s+
e^{\int_{0}^{\zeta^{\mathbb{Q}}_{y|x}} \frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du}\xi_{\tau^{x,D'}}\bigg)\bigg]\\ &\ge \mathbb{E}^{\theta} \bigg[ \mathbb{1}_{\zeta^{\mathbb{Q}}_{y|x}<\zeta^{\mathbb{Q}}_{0|x}}\mathbb{E}^{\theta}\bigg[
\int_{\zeta^{\mathbb{Q}}_{y|x}}^{\zeta^{\mathbb{Q}}_{0|y}} e^{\int_{0}^{s} \frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du -\rho s} dD_s + e^{\int_{0}^{\zeta^{\mathbb{Q}}_{y|x}}
\frac{\theta^2_u}{2{\@fontswitch\relax\mathcal{R}}}du}\xi_{\tau^{x,D'}}\bigg|\mathcal{F}_{\zeta^{\mathbb{Q}}_{y|x}}\bigg] \bigg]\\ & \ge \mathbb{Q}(\zeta^{\mathbb{Q}}_{y|x}<\zeta^{\mathbb{Q}}_{0|x}\wedge
\epsilon)e^{-\rho\epsilon}J^{\theta}(y,{D}) \ge (1-\epsilon)e^{-\rho\epsilon}J^{\theta}(y,{D}).
\end{align}\] Hence, \[\begin{align} &\inf_{\theta} J^{\theta}(y,{D}) - J^{*}(x) \le \inf_{\theta} J^{\theta}(y,{D}) - \inf_{\theta} J^{\theta}(x;D') \\ &\quad \le \inf_{\theta} J^{\theta}(y,{D})
-(1-\epsilon)e^{-\rho\epsilon}J^{\theta}(y,{D}) \le \big(1-(1-\epsilon)e^{-\rho \epsilon}\big) J^*(h),
\end{align}\] for some \(h>y\) sufficiently large. Therefore \(J^*(x)\) is continuous at \(x>0\). ◻
We start with some preliminary results on the free-boundary problem 24 .
Recall the function \(\psi^+\) (see Assumption 15), we claim that for every \(b>0\) and \(\gamma\in \mathbb{R}\) that there is a function \(v_{b,\gamma}\in C^2((0,b)\cup(b,\infty))\cap C^1(\mathbb{R}_+)\) (with \(\gamma=0\), \(v_{b,0}\in C^2({\mathbb{R}_+})\)) solving the following nonlinear ODE: for \(x\in\mathbb{R}_+\), \[\begin{align} \left\{
\begin{aligned} &\dfrac{\sigma^2(x)}{2}v''_{b,\gamma}(x)+ \mu(x)v_{b,\gamma}'(x) - {\@fontswitch\relax\mathcal{R}}\dfrac{\sigma^2(x)}{2}\dfrac{\big(v_{b,\gamma}'(x)\big)^2}{v_{b,\gamma}(x)} = (\rho+\gamma) v_{b,\gamma}(x) ,\\
&v_{b,\gamma}'(b)=1, \quad v_{b,\gamma}(b)= \psi^+(b). \label{eq:HJB95ODE1} \end{aligned} \right.
\end{align}\tag{38}\] Indeed, for any given \(b>0\) and \(\gamma\in \mathbb{R}\), one-to-one correspondence \[\begin{align}
\label{eq:1to195cor} h_{b,\gamma}:=(v_{b,\gamma})^{1-{\@fontswitch\relax\mathcal{R}}}
\end{align}\tag{39}\] enables to consider the following linear ODE given by \[\begin{align}
\label{eq:ODE95h} \left\{ \begin{aligned} &\dfrac{1}{2}\sigma^2(x)h''_{b,\gamma}(x)+ \mu(x)h_{b,\gamma}'(x) = (1-{\@fontswitch\relax\mathcal{R}})(\rho+\gamma) h_{b,\gamma}(x)\quad on\;\;[a_1,a_2],\\
&h_{b,\gamma}'(b)=(1-{\@fontswitch\relax\mathcal{R}})\big(\psi^+(b)\big)^{-{\@fontswitch\relax\mathcal{R}}}, \quad h_{b,\gamma}(b)= \big(\psi^+(b)\big)^{1-{\@fontswitch\relax\mathcal{R}}}, \end{aligned} \right.
\end{align}\tag{40}\] with an interval \([a_1,a_2]\subset \mathbb{R}\) with \(a_1<b<a_2\). Since the linear ODE 40 admits a unique solution
\(h_{b,\gamma}\in C^2([a_1,b)\cup(b,a_2])\cap C^1([a_1,a_2])\) (\(h_{b,0}\in C^2([a_1,a_2])\) when \(\gamma=0\)) for any \([a_1,a_2]\subset \mathbb{R}\) with \(a_1<b<a_2\) (The Picard-Lindelöf’s Existence and Uniqueness Theorem, see e.g., Nagle et al. [80]), this ensures our claim to hold.
Define \(g_{b,\gamma}:\mathbb{R}_+\ni x\rightarrow g_{b,\gamma}(x)\in\mathbb{R}\) for every \(b>0\) and \(\gamma\in \mathbb{R}\) by \[\begin{align} g_{b,\gamma}(x) := \frac{v'_{b,\gamma}(x)}{v_{b,\gamma}(x)}. \end{align}\] Since \(v_{b,\gamma}\in C^2((0,b)\cup(b,\infty))\cap C^1(\mathbb{R}_+)\) (when \(\gamma=0\), \(v_{b,0}\in C^2(\mathbb{R}_+)\)) is the solution of 38 and satisfies \[\begin{align} \label{eq:equiv95deriv952} v_{b,\gamma}(x) = \psi^+(b)\exp\left(\int_{b}^x g_{b,\gamma}(u) du \right), \end{align}\tag{41}\] it follows that \(g_{b,\gamma}\) is of \(C^1((0,b)\cup(b,\infty))\cap C(\mathbb{R}_+)\) (when \(\gamma=0\), \(g_{b,0}\in C^1({\mathbb{R}_+})\)) solves the following ODE: for \(x\in \mathbb{R}_+\) \[\begin{align} \left\{ \begin{aligned} &\dfrac{\sigma^2(x)}{2}g_{b,\gamma}'(x)+\mu(x)g_{b,\gamma}(x) + \dfrac{(1-{\@fontswitch\relax\mathcal{R}})}{2}\sigma^2(x)g_{b,\gamma}^2(x) = \rho+\gamma,\\ &g_{b,\gamma}(b) = {1}/{\psi^+(b)}.\label{eq:ODE95g} \end{aligned} \right. \end{align}\tag{42}\] For notational simplicity, we set \(v_b:=v_{b,0}\), \(h_{b}:=h_{b,0}\), and \(g_{b}:=g_{b,0}\) for every \(b>0\). Targeting at proving Theorem 18, our main task is to show the following proposition. The proof of this result is based on a sequence of preliminary lemmas (Lemmas 38-42), which we will discuss in next subsection.
Proposition 37. Suppose that Assumption 15 holds. For every \(b>0\), let \(v_{b}\in C^2({\mathbb{R}_+})\) be the solution of the nonlinear ODE given in 38 (when \(\gamma=0\)). Then there exists a unique threshold \(b^{\ast}\in (\underline{b},\hat{b})\) such that \[\begin{align} v_{b^{*}}(0) =\xi_0,\quad v'_{b^{*}}(x)\ge 1 \quad on\;\;(0,b^{*}], \end{align}\] with \(\underline{b},\hat{b}\) appearing in Assumption 15. We adopt the convention that \(v_{b^{*}}\) is extended to 0 by continuity and the value of \(v_{b^{*}}\) at 0 is its limit from the right.
At this point, we see that \(v^*(x) = v_{b^{*}}(x)\) to the left of \(b^*\) satisfies the first line and the third line of 24 . Setting up \((v^*)^{\prime}(x)=1\) on \((b^*,\infty)\), it remains to show the second line of 24 holds. We present this final piece of the proof for Theorem 18 here.
Proof of Theorem 18.. From Proposition 37, define \(v^*:{\mathbb{R}_+}\rightarrow \mathbb{R}\) by \[\begin{align} \label{eq:const95v} v^*(x):=\left\{ \begin{aligned} &v_{b^*}(x)\quad &&if\;\;x\in [0,b^*];\\ &x-b^*+\psi^+(b^*)\quad && if\;\;x\in (b^*,\infty). \end{aligned} \right. \end{align}\tag{43}\] The function \(v_{b^*}\) is in \(\mathcal{C}^2(\mathbb{R}_+)\) and satisfies 38 with \(\gamma = 0\). In particular, \(v_{b^*}''(b^*) = 0\) by construction. Indeed, substituting the boundary conditions \(v_{b^*}'(b^*) = 1\) and \(v_{b^*}(b^*) = \psi^+(b^*)\) into the ODE at \(x = b^*\) gives \[\frac{\sigma^2(b^*)}{2} v_{b^*}''(b^*) + \mu(b^*) - {\@fontswitch\relax\mathcal{R}}\frac{\sigma^2(b^*)}{2} \frac{1}{\psi^+(b^*)} - \rho \psi^+(b^*) = 0.\] Since \(\psi^+(b^*)\) solves the quadratic equation \(Q(\psi; b^*) = 0\), it follows that \(v_{b^*}''(b^*) = 0\). By the definition given in 43 , the function \(v^*\) is in \(\mathcal{C}^2({\mathbb{R}_+})\) and satisfies the free boundary problem 24 to the left of \(b^*\). It remains to validate for \(x\in(b^*,\infty)\), \({\@fontswitch\relax\mathcal{L}} v^*\le 0\) as \(({v^*})^{\prime}(x)=1\) is trivial.
For \(x \in (b^*, \overline{b})\), note that \(\psi^+(x) - x\) is decreasing, as stated in Assumption 15 ii.. Consequently, we have \(v^*(x) = x - b^* + \psi^+(b^*) \ge \psi^+(x)\). Hence \(\mathcal{L}v^*(x)\) is given by \[\mathcal{L}v^*(x) = -\frac{Q(v^*(x); x)}{v^*(x)} \le 0,\] where \(Q(\psi; x)\) is defined in Assumption 15 ii..
If \(\overline{b} = \infty\), the proof concludes here. Otherwise, for \(x \in [\overline{b}, \infty)\), we have \(Q(v^*(x); x) < 0\), which implies \(\mathcal{L}v^*(x) < 0\). Thus, we conclude that \(v^*(x)\) satisfies \(\mathcal{L}v^* \le 0\) on the interval \((b^*, \infty)\). This means it solves the free boundary problem 24 with the free boundary \(b^* \in (\underline{b}, \hat{b})\). This completes the proof. ◻
In what follows, we often make use of the following elementary properties that are based on Cohen et al. [20].
Lemma 38. Let \((a_1,a_2)\subseteq \mathbb{R}\) and \(a_3\in (a_1,a_2)\), and let \(f:(a_1,a_2)\ni x \rightarrow f(x)\in \mathbb{R}\) be a function of class \(C^1((a_1,a_2))\). Then the following hold:
If \(f(a_3)>c\) (resp. \(<c\)) with some \(c\in\mathbb{R}\) and there exist \(y_1: = \sup\{y\in(a_1,a_3):f(y)=c\}\) and \(y_2 : =\inf\{y\in(a_3,a_2):f(y)=c \}\), then \[f'(y_1)\ge0\;\;(resp.~\le 0),\quad f'(y_2)\le 0\;\;(resp.~\ge0).\]
If \(f'(a_3)>0\) (resp., \(<0\)) and there exist \(y_3:=\sup\{y\in(a_1,a_3):f(y)=f(a_3)\}\) and \(y_4:=\inf\{y\in(a_3,a_2):f(y) = f(a_3)\}\), then \[f'(y_3)\le 0\;\;(resp.~\ge0),\quad f'(y_4)\le 0\;\; (resp.~\ge0).\]
We start with a perturbation result which helps to get estimations for \(g_b\) via the estimates of \(g_{b,\gamma}\).
Lemma 39. Suppose that Assumption 15 holds. For every \(b>0\), let \(v_{b,\gamma}\) be the solution of 38 and \(g_{b,\gamma}\) be the solution of 42 . Then the following hold:
Let \(b\in (\underline{b},\overline{b})\) and \(\gamma<0\) (with \(\underline{b},\overline{b}\) appearing in Assumption 15). Define \[y_{b,\gamma}:=\sup\{x\in(0,b):g_{b,\gamma}(x) \le 1/\psi^+(x)\},\] with the usual convention \(\sup\{\varnothing\} : =0\). Then \(y_{b,\gamma}\in[0,\underline{b}]\), and for every \(x\in(y_{b,\gamma},b)\), \(g_{b,\gamma}(x)> 1/\psi^+(x)\).
Let \(b_1,b_2\in (\underline{b},\overline{b})\) with \(b_1\leq b_2\) and \(\gamma_1,\gamma_2 \in (-\infty,0]\) with \(\gamma_1>\gamma_2\). Then the following holds: for every \(x\in(0,b_1)\), \[g_{b_1,\gamma_1}(x)<g_{b_2,\gamma_2}(x).\]
As \(\gamma\rightarrow0\), we have \(\sup_{x\in[0,b]} |g_{b,\gamma}(x) - g_b(x)| = O(\gamma),\) where \(O(\cdot)\) denotes the Landau symbol. Moreover, for every \(x\in[0,b]\), \[\begin{align} \lim_{\gamma\rightarrow0}|v_{b,\gamma}(x) - v_{b}(x)|=0,\qquad \lim_{\gamma\rightarrow0}|v'_{b,\gamma}(x) - v'_{b}(x)|=0. \end{align}\]
Proof of of i.. We first note that by 42 , the following inequality holds: \[\begin{align} \label{eq:est95g1} g_{b,\gamma}'(b)= \frac{2}{(\sigma(b)\psi^+(b))^2}Q(\psi^+(b))+\frac{2\gamma}{(\sigma(b))^2}-\frac{1}{(\psi^+(b))^2}<-\frac{1}{(\psi^+(b))^2}, \end{align}\tag{44}\] where we use that \(Q(\psi^+(b))=0\) (defined in Assumption15ii.) and \(\gamma<0\).
Define \(\phi_{b,\gamma}:\mathcal{X}\rightarrow \mathbb{R}\) as \(\phi_{b,\gamma}(x) := g_{b,\gamma}(x) - 1/\psi^+(x)\). Since \(g_{b,\gamma}(b)={1}/{\psi^+(b)}\) and \((\psi^+)'(b)\le 1\) (because \(\psi^+(x)-x\) is decreasing on \([\underline{b},\overline{b})\) and \(b\) is in \((\underline{b},\overline{b})\); see Assumption15ii.), the inequality in 44 gives \[\begin{align} \label{eq:est95g2} \phi_{b,\gamma}(b)=0,~\phi_{b,\gamma}'(b)=g'_{b,\gamma}(b) + \frac{(\psi^+)'(b)}{(\psi^+(b))^2} < \frac{(\psi^+)'(b)-1}{(\psi^+(b))^2} \le0 . \end{align}\tag{45}\]
We claim that \(y_{b,\gamma}\leq \underline{b}\). Since the case with \(y_{b,\gamma}=0\) is trivial, we can and do consider only the case with \(y_{b,\gamma}=\sup\{x\in(0,b):\phi_{b,\gamma}(x)=0\}\). Arguing by contradiction, assume that \(y_{b,\gamma}>\underline{b}\). Then by the continuity of \(\phi_{b,\gamma}\) (due to that of \(g_{b,\gamma}\) and \(\psi^+\)), we have that \(\phi_{b,\gamma}(y_{b,\gamma})= g_{b,\gamma}(y_{b,\gamma}) - {1}/{\psi^+(y_{b,\gamma})}=0.\) Furthermore, using the same arguments devoted for the second property given in 45 (along with \(y_{b,\gamma}>\underline{b}\)), we have \[(\phi_{b,\gamma})'(y_{b,\gamma})=g'_{b,\gamma}(y_{b,\gamma}) + \frac{(\psi^+)'(y_{b,\gamma})}{(\psi^+(y_{b,\gamma}))^2} <0,\] which contradicts Lemma 38ii.. Hence, the claim holds true.
Furthermore, from the definition of \(y_{b,\gamma}\) and the fact that \((\phi_{b,\gamma})'(b)=(g_{b,\gamma}- 1/\psi^+)'(b)<0\) (see 45 ), the last statement of i. is true. This completes the proof.
To that end, let \(y_{b_2,\gamma_2}\) be defined as Lemma 39i. (by assigning \(b=b_2\) and \(\gamma=\gamma_2\)). Define \(\Delta g: [0,b]\ni x\rightarrow \Delta g(x)\in \mathbb{R}\) by \(\Delta g(x):= g_{b_1,\gamma_1}(x)-g_{b_2,\gamma_2}(x)\) for \(x\in [0,b_1]\). Since \(g_{b_i,\gamma_i}\), \(i=1,2\), is the solution of 42 (when \(b=b_i\), \(\gamma=\gamma_i\)), the following holds \[\begin{align} &\dfrac{1}{2}\sigma^2(x)(\Delta g)'(x) + \Delta g(x) \Big( \mu(x)+ \frac{(1-{\@fontswitch\relax\mathcal{R}})}{2}\sigma^2(x)\left({\Delta g(x)} + 2g_{b_2,\gamma_2}(x)\right)\Big) \nonumber\\ &= \gamma_1-\gamma_2,\quadfor\;\;x\in [0,b_1]\label{eq:est95g4952} \end{align}\tag{46}\] and \(\Delta g(b_1) = 1/{\psi^+(b_1)}-g_{b_2,\gamma_2}(b_1)<0\) which follows from Lemma 39i. with fact that \(b_1\in (y_{b_2,\gamma_2},b_2)\) and \(g_{b_1,\gamma_1}(b_1)=1/{\psi^+(b_1)}\).
Here, we claim that \(\Delta g(x)<0\) on \((0,b_1)\). Arguing by contradiction, assume it does not hold; then there exists \(y_{3}:=\sup\{x\in(0,b_1): \Delta g(x) = 0\}\). Since \(\Delta g(y_3)=0\) (by its continuity) and \(y_3\in[0,b_1]\), assigning \(x=y_3\) into 46 ensures that \(\Delta g'(y_3)=2(\gamma_1-\gamma_2)/\sigma^2(y_3)>0\). This together with the fact that \(\Delta g(b_1)<0\) contradicts Lemma 38i. This completes the proof.
This proof is similar to the one given in Cohen et al. [20], so we omit the details here. ◻
Lemma 40. Suppose that Assumption 15 holds. For every \(b>0\), let \(v_b\in C^2(\mathbb{R}_+)\) be the solution of 38 (when \(\gamma=0\)) and \(g_{b}\in C^1(\mathbb{R}_+)\) be the solution of 42 (when \(\gamma=0\)). Then the following hold:
Let \(b\in(\underline{b},\overline{b})\). Then there exists \(y_b^*\in [0,\underline{b}]\) satisfying that for all \(x\in [y_b^*,b]\), \(g_{b}(x)\ge 1/\psi^+(x),\) with \(\underline{b}\) appearing in Assumption 15.
Let \(b_1,b_2\in (\underline{b},\overline{b})\) with \(b_1<b_2\). Then the following holds: for all \(x\in (0,b_1]\), \(g_{b_1}(x)\le g_{b_2}(x).\)
Let \(b\in (\underline{b},\overline{b})\) and \(\gamma\le 0\). As \(\varepsilon \rightarrow 0\), \(\sup_{x\in[0,b]}|g_{b+\varepsilon,\gamma}(x) - g_{b,\gamma}(x)| =O(\varepsilon)\). Furthermore, for every \(x\in[0,b]\), \[\begin{align} \label{eq:lem95cts95b95g952} \lim_{\varepsilon \rightarrow 0} |v_{b+\varepsilon,\gamma}(x) - v_{b}(x)| =0,\qquad \lim_{\varepsilon \rightarrow 0} |v_{b+\varepsilon, \gamma}'(x) - v_{b,\gamma}'(x)| =0. \end{align}\qquad{(7)}\]
Proof of i.. Let \(y_{b,\gamma}\in[0,\underline{b}]\) for every \(\gamma<0\) be defined as in Lemma 39i. Then \(y_{b,\gamma}\) increases in \(\gamma<0\). Indeed, Lemma 39ii. ensures that for any \(\gamma_1,\gamma_2\in(-\infty,0)\) such that \(\gamma_1>\gamma_2\), the following holds that \(g_{b,\gamma_1}(x) <g_{b,\gamma_2}(x)\) for every \(x\in(0,b)\). Combining this with the definition of \(y_{b,\gamma_i}\), \(i=1,2\), (see Lemma 39i.), we have \(y_{b,\gamma_1}>y_{b,\gamma_2}\). This ensures that the (increasing) monotonicity of \(y_{b,\gamma}\) with resepect to \(\gamma<0\).
Consider an increasing sequence \((\gamma_n)_{n\in \mathbb{N}}\subseteq (-\infty,0)\) such that \(\gamma_n\uparrow 0\) as \(n\rightarrow\infty\). Then the monotonicity of \((y_{b,\gamma_n})_{n\in\mathbb{N}}\) together with uniformly boundedness within \([0,\underline{b}]\) ensures that there exists \(y_{b}^*:=\lim_{n\rightarrow\infty}y_{b,\gamma_n}\) satisfying \(y_{b}^*\in[0,\underline{b}]\) and \(y_{b}^*\geq y_{b,\gamma_n}\) for every \(n\in \mathbb{N}\). Combined with Lemma 39i., this ensures that for every \(n\in\mathbb{N}\) and every \(x\in[y_{b}^*,b)\), \(g_{b,\gamma_n}(x)\geq 1/\psi^+(x).\) Therefore combined with Lemma 39iii. and letting \(n\rightarrow \infty\), this completes the proof.
Lemma 39ii. ensures that for every \(\gamma<0\) and every \(x\in(0,b_1]\), \(g_{b_1,0}(x)=g_{b_1}(x)< g_{b_2,\gamma}(x).\) Combined with Lemma 39iii. and letting \(\gamma\uparrow 0\), this completes the proof.
We establish the proof for the case where \(\gamma = 0\), as the analysis can be readily extended to accommodate the case where \(\gamma\neq 0\). We start with proving the first convergence therein. Let \(\varepsilon\in [0,1\wedge(\overline{b}-b))\). Since \(g_{b+\varepsilon}\in C^2(\mathbb{R}_+)\) (resp. \(g_{b}\in C^2(\mathbb{R}_+)\)) is the solution of 42 (when \(\gamma=0\) with \(b+\varepsilon\) (resp. \(b\))), the following holds: for every \(x\in\mathbb{R}_+\), \[\begin{align} &g_{b+\varepsilon}(x) - g_{b}(x) -\big(g_{b+\varepsilon}(b) - g_{b}(b)\big)= - \int_{x}^b \big(g'_{b+\varepsilon}(y) - g'_{b}(y) \big) dy \nonumber\\ &= \int_{x}^b \Big( (1-{\@fontswitch\relax\mathcal{R}})\sigma^2(x)\big(g_{b+\varepsilon}(y) +g_b(y)\big) + \frac{2\mu(x)}{\sigma^2(x)} \Big)\big(g_{b+\varepsilon}(y) - g_b(y)\big) dy. \label{eq:est95g5} \end{align}\tag{47}\]
Lemma 40ii. ensures that \(g_b(x)\le g_{b+\varepsilon}(x)\le g_{b+1}(x)\le C_0\) for every \(x\in[0,b]\), where \(C_0>0\) is a constant (that depends on \(1\wedge (\overline{b}-b)\) but not on \(\varepsilon\)). Furthermore, \(\sigma\) and \(\mu\) are uniformly bounded on \([0,b]\) (see Assumption 15). Hence, combined with 47 these properties ensure that there is a constant \(C_1>0\) (that depends on \(C_0\) but not on \(\varepsilon\)) such that for every \(x\in[0,b)\), \(0\leq g_{b+\varepsilon}(x) - g_{b}(x) \leq g_{b+\varepsilon}(b) - g_{b}(b) + C_1 \int_b^x \big(g_{b+\varepsilon}(y) - g_b(y)\big) dy.\) An application of Grönwall’s inequality gives the existence of a constant \(C_2>0\) satisfying \[\begin{align} \label{eq:est95g6} 0\leq \sup_{x\in[0,b]}(g_{b+\varepsilon}(x) - g_{b}(x)) \leq C_2 (g_{b+\varepsilon}(b) - g_{b}(b)). \end{align}\tag{48}\]
Furthermore, we claim that there is a constant \(C_3\) (that depends on \(1\wedge (\overline{b}-b)\) but not on \(\varepsilon\)) such that \(0\leq g_{b+\varepsilon}(b) - g_{b}(b) \leq C_3 \varepsilon.\) Indeed, \[\begin{align} \left|g_{b+\varepsilon}(b) - g_{b}(b) \right| \le |g_{b+\varepsilon}(b+\varepsilon) -g_{b}(b) | + |g_{b+\varepsilon}(b+\varepsilon) - g_{b+\varepsilon}(b)|: = \operatorname{I} + \operatorname{II}. \end{align}\] \(\operatorname{I}\) is bounded by \(\varepsilon \sup_{x\in[b,b+\varepsilon]}|{(\psi^+)'(x)}/{(\psi^+)^2(x)}|\), which is of \(O(\varepsilon)\) by the fact that \(g_{b+\varepsilon}(b+\varepsilon)=1/\psi^+(b+\varepsilon)\), \(g_{b}(b)=1/\psi^+(b)\), and \(\psi^+\in C^1([0,\overline{b}])\). \(\operatorname{II}\) is of \(O(\varepsilon)\) with similar arguments devoted for 47 .
Combined with 48 , this ensures that \[\begin{align} \lim_{\varepsilon\downarrow 0}\sup_{x\in[0,b]}\Big(g_{b+\varepsilon}(x) - g_{b}(x)\Big) =0. \end{align}\] ?? holds because \(v_{b}(x) = \psi^+(b)\exp(\int_{b}^x g_{b}(u) du )\) and \(v_b'(x) = v_b(x)g_b(x)\).
The proof in the case \(\varepsilon\uparrow 0\) is similar therefore we omit it. ◻
Lemma 41. Suppose that Assumption 15 holds. Let \(b\in (0,\underline{b}]\) and \(v_b\in C^2(\mathbb{R}_+)\) be the solution of 38 (when \(\gamma=0\)), with \(\underline{b}\) appearing in Assumption 15. Then for every \(x\in(0,b]\), \(v_{b}'(x) \le 1.\)
Proof. Let \(\gamma>0\) and \(v_{b,\gamma}\) be the solution of 38 . Since \(v_{b,\gamma}'(b)=1\) and \(v_{b,\gamma}(b)= \psi^+(b)\), the nonlinear ODE given in 38 ensures that \(v''_{b,\gamma}(b) = 2\gamma\psi^+(b)/\sigma^2(b)>0\).
We claim that for every \(x\in(0,b)\), \(v'_{b,\gamma}(x)<1\). Arguing by contradiction, we assume that it does not hold. Then together with \(v_{b,\gamma}'(b)=1\), the following supremum is attained: \(x_1:=\sup\{x\in(0,b): v'_{b,\gamma}(x) = 1 \}\). By definition of this, it is clear that for every \(x\in[x_1,b]\), \(v'_{b,\gamma}(x)\le1,\) which ensures that \(v_{b,\gamma}(x) - v_{b,\gamma}(b)\geq x-b\) for every \(x\in[x_1,b]\).
Furthermore, from Assumption 15ii. (with \(b<\underline{b}\)) and \(v_{b,\gamma}=\psi^+(b)\), it
follows that for every \(x\in[x_1,b]\), \(v_{b,\gamma}(x_1) \ge \psi^+(b)-(b-x_1) \ge \psi^+(x_1)\). This implies that \(Q(v_{b,\gamma}(x_1)) \ge
Q(\psi^+(x_1))=0\) (see Assumption 15ii.). From the nonlinear ODE 38 , it follows that \[\begin{align}
\frac{1}{2}\sigma^2(x_1) v''_{b,\gamma}(x_1)=Q\big(v_{b,\gamma}(x_1)\big)\frac{1}{v_{b,\gamma}(x_1)}+\gamma v_{b,\gamma}(x_1) \ge \gamma \psi^+(x_1)>0.
\end{align}\] Combining this with the fact that \(v''_{b,\gamma}(b)>0\) contradicts Lemma 38ii. Hence we have shown that the claim
holds.
Using Lemma 39iii. and letting \(\gamma\downarrow0\), we complete the proof. ◻
Lemma 42. Suppose that Assumption 15 holds. For every \(b>0\) and \(\gamma\in\mathbb{R}\), let \(v_{b,\gamma}\in C^2((0,b)\cup(b,\infty))\cap C^1(\mathbb{R}_+)\) (\(v_{b}=v_{b,0}\in C^2(\mathbb{R}_+)\) when \(\gamma=0\)) be the solution of 38 . Then the following hold:
Let \(b\le \hat{b}\) (with \(\underline{b},\hat{b}\) appearing in Assumption 15) and \(\gamma\le 0\). If \(v_{b,\gamma}(0) \ge \xi_0\), then for every \(x\in[0,b]\), \(v_{b,\gamma}(x)\ge x + \xi_0.\)
Let \(\gamma < 0\). Then there exists a unique threshold \(b^*(\gamma)\) defined by \[b^*(\gamma) := \inf\{b \in (\underline{b}, \hat{b}) : v_{b,\gamma}(0) = \xi_0\}.\] Moreover, \(b^*(\gamma)\) is monotonically increasing in \(\gamma < 0\). Consequently, \[\begin{align} \label{def:b42gamma} b^* := \lim_{\gamma \uparrow 0} b^*(\gamma) \end{align}\qquad{(8)}\] is well-defined and unique, with \(b^* \in (\underline{b}, \hat{b}]\) and \(v_{b^*}(0) = \xi_0\).
Proof of i.. Arguing by contradiction, we assume that it does not hold. Recalling the one-to-one correspondence \(h_{b,\gamma}=(v_{b,\gamma})^{1-{\@fontswitch\relax\mathcal{R}}}\) (see 39 ), we obtain that \[\max_{x\in[0,b]}\left\{ (x+\xi_0)^{1-{\@fontswitch\relax\mathcal{R}}} - h_{b,\gamma}(x) \right\} > 0.\] We note that from Assumption15iii. (with \(b\leq \hat{b}\)) and the condition \(v_{b,\gamma}(0) \ge \xi_0\), it follows that \(h_{b,\gamma}(x)\ge(x+\xi_0)^{1-{\@fontswitch\relax\mathcal{R}}}\) when \(x=0\) and \(x=b\). Hence this implies that \((x+\xi_0)^{1-\epsilon} - h_{b,\gamma}(x)\) takes its maximum at \(x_0\in(0,b)\). We have \[\begin{align} \label{pf:lem95comp951} -{\@fontswitch\relax\mathcal{R}}(1-{\@fontswitch\relax\mathcal{R}})(x_0+\xi)^{-1-{\@fontswitch\relax\mathcal{R}}} \le h_{b,\gamma}''(x_0),~(1-{\@fontswitch\relax\mathcal{R}})(x_0+\xi)^{-{\@fontswitch\relax\mathcal{R}}} =h_{b,\gamma}'(x_0). \end{align}\tag{49}\] From the linear ODE given in 40 , we get to a contradiction \[\begin{align} &0 = \dfrac{1}{2}\sigma^2(x_0)h_{b,\gamma}''(x_0)+ \mu(x_0)h_{b,\gamma}'(x_0) - (1-{\@fontswitch\relax\mathcal{R}})(\rho+\gamma) h_{b,\gamma}(x_0)\\ > & (1-{\@fontswitch\relax\mathcal{R}})(x_0+\xi_0)^{-1-{\@fontswitch\relax\mathcal{R}}}\Big(-\frac{{\@fontswitch\relax\mathcal{R}}}{2}\sigma^2(x_0) + \mu(x_0)(x_0+\xi) -(\rho+\gamma)(x_0+\xi)^2\Big)\ge0, \end{align}\] where the first inequality follows from 49 and that \(h_{b,\gamma}(x_0) < (x_0+\xi)^{1-{\@fontswitch\relax\mathcal{R}}}\), and the second inequality follows from Assumption15ii. and \(\gamma\le 0\). This completes the proof of i.
This statement is proved in three steps.
We claim that \(v_{\underline{b},\gamma}(0)> \xi_0\) and \(v_{\hat{b},\gamma}(0)<\xi_0\).
Lemma 41 (when \(b=\underline{b}\)) ensures that \(v_{\underline{b},\gamma}(0)\geq v_{\underline{b},\gamma}(\underline{b})-(\underline{b}-x)\) for every \(x\in[0,\underline{b}]\). In particular, \(v_{\underline{b},\gamma}(0)>v_{\underline{b},\gamma}(\underline{b})-\underline{b}\) because it cannot be the case that \(v_{\underline{b},\gamma}'(x)\equiv 0\) for \(x\in[0,\underline{b}]\). Furthermore, from the fact that \(v_{\underline{b},\gamma}(\underline{b})=\psi^+(\underline{b})\) and Assumption 15ii.iii. (with \(\underline{b}<\hat{b}\)), it follows that \(v_{\underline{b},\gamma}(0) > \psi^+(\underline{b})-\underline{b}> \psi^+(\hat{b})-\hat{b} = \xi_0\).
It remains to show \(v_{\hat{b},\gamma}(0)<\xi_0\). Arguing by contradiction, we assume that it does not hold. From Lemma 42i., it follows that \(v_{\hat{b},\gamma}(x) \ge x + \xi_0\) for every \(x\in[0,\hat{b}]\). Since \(v_{\hat{b},\gamma}\) is the solution of 38 , assigning \(b=\hat{b}\) into 38 ensures that \(v''_{\hat{b},\gamma}(\hat{b}) = 2\gamma\sigma^2(\hat{b})\psi^+(\hat{b})<0\).
By using a proof by contradiction, standard arguments shows that \(v_{\hat{b},\gamma}'(x)>1\), for \(x\in(\underline{b},\hat{b})\). From the claim, the fact that \(v_{\hat{b},\gamma}(\hat{b})=\psi^+(\hat{b})\), and Assumption 15ii. (with \(\hat{b}\in[\underline{b},\overline{b}]\)), it follows that \(v_{\hat{b},\gamma}(\underline{b}) < \psi^+(\hat{b}) - (\hat{b}-\underline{b}) = \xi_0+\underline{b}\). This contradicts with
Lemma 42i. Therefore, we conclude that \(v_{\hat{b},\gamma}(0)< \xi_0\).
We claim that \(b^*(\gamma): = \inf\{b\in(\underline{b},\hat{b}): v_{b,\gamma}(0) = \xi_0\}\) is well-defined and that \(v_{b,\gamma}(0) < \xi_0\) for every \(b\in( b^*(\gamma), \hat{b})\).
By Lemma 40iii., we have the continuity of \(b\mapsto v_{b,\gamma}(0)\) with fixed \(\gamma\le 0\). There \(b^*(\gamma): = \inf\{b\in(\underline{b},\hat{b}): v_{b,\gamma}(0) = \xi_0\}\) is well-defined.
Now we claim that \(v_{b,\gamma}(0) < \xi_0\) for every \(b\in( b^*(\gamma), \hat{b})\). From Lemma 40i. together with \(b^*(\gamma)>\underline{b}\), it follows that \(g_{b,\gamma}(x) \ge 1/\psi^+(x)\) for every \(x\in [b^*(\gamma),b]\). Combining this with the fact that \((\psi^+)'(x) < 1\) for every \(x\in[b^*(\gamma),b]\) (see Assumption 15ii.) ensures that \(g_{b,\gamma}(x) > (\psi^+)'(x)\psi^+(x)\) for every \(x\in [b^*(\gamma),b]\).
Furthermore, as \(b>b^*(\gamma)\), Lemma 40ii. ensures that \(g_{b,\gamma}(x) > g_{b^*(\gamma),\gamma}(x)\) for \(x\in(0,b^*(\gamma)]\). Hence we conclude that \[\begin{align} v_{b,\gamma}(0) &=\psi^+(b) \exp\bigg({\int_{b}^{b^*(\gamma)}g_{b,\gamma}(y)dy + \int_{b^*(\gamma)}^{0}g_{b,\gamma}(y)dy }\bigg)\\ & <\psi^+(b^*(\gamma)) \exp \bigg(\int_{b^*(\gamma)}^{0}g_{b^*(\gamma),\gamma}(y)dy \bigg) = v_{b^*(\gamma),\gamma}(0) = \xi_0. \end{align}\] . We claim that \(b^*(\gamma)\) increases in \(\gamma<0\), which ensures that \(b^*\) defined in ?? is well-defined by monotone convergence. Furthermore, we claim that \(b^*\in (\underline{b},\hat{b})\) and \(v_{b^*}(0) = \xi_0.\)
Let \(b\in(\underline{b},\overline{b})\) and \(\gamma_1,\gamma_2\in(-\infty,0]\) with \(\gamma_1\ge\gamma_2\). Then Lemma 39ii. ensures that \(g_{b,\gamma_1}(x) < g_{b,\gamma_2}(x)\) for every \([0,b]\). From this and the relationship given in 41 , it follows that \(v_{b,\gamma_1}(x) \ge v_{b,\gamma_2}(x)\) on \([0,b]\). In particular, since it cannot be the case that \(g_{b,\gamma_1} \equiv v_{b,\gamma_2}\) for \(x\in[0,b]\), hence \(v_{b,\gamma_1}(0) > v_{b,\gamma_2}(0)\).
From this and the fact that \(b^*(\gamma_1)=\inf\{b\in(\underline{b},\hat{b}): v_{b,\gamma_1}(0) = \xi_0\}\), it follows that \(\xi_0 = v_{b^*(\gamma_1),\gamma_1}(0) > v_{b^*(\gamma_1),\gamma_2}(0)\). Furthermore, Step 2 ensures that \(v_{b,\gamma_2}(0) < \xi_0\) for every \(b\in( b^*(\gamma_2), \hat{b})\). We hence have \(b^*(\gamma_1) > b^*(\gamma_2)\).
By the monotonicity of \(b^*(\gamma)\) and its uniform boundedness, we can apply the MCT to have the existence of \(b^*\) in ?? .
It remains to show that \(v_{b^*}(0) = \xi_0\) and \(b^*\in (\underline{b},\hat{b})\). To that end, consider an increasing sequence \((\gamma_n)_{n\in \mathbb{N}}\) such that \(\gamma_n\in (-\infty,0)\) for every \(n\in\mathbb{N}\) and \(\lim_{n\rightarrow \infty } \gamma_n = 0\).
Since \(\lim_{n\rightarrow \infty} b^*(\gamma_n) = b^*\), Lemma 40iii. ensures that for any \(\varepsilon>0\), there exists \(N_\varepsilon \in \mathbb{N}\) satisfying that for every \(n\geq N_\varepsilon\), \(|v_{b^*}(0) - v_{b^*(\gamma_n)}(0)|<\varepsilon.\) Furthermore Lemma 39iii. ensures that there exists \(\widetilde{N}_\varepsilon\in \mathbb{N}\) satisfying that for every \(n\geq \widetilde{N}_\varepsilon\), \(|v_{b^*(\gamma_n),\gamma_n}(0) - v_{b^*(\gamma_n)}(0)|<\varepsilon.\) Since \(v_{b^*(\gamma_n),\gamma_n}(0) = \xi_0\) for every \(n\in\mathbb{N}\) (see Step 2.), we hence obtain that for every \(n\geq \widetilde{N}_\varepsilon \vee N_\varepsilon\), \[\begin{align} |v_{b^*}(0) - \xi_0 |\leq |v_{b^*}(0) - v_{b^*(\gamma_n)}(0)|+|v_{b^*(\gamma_n),\gamma_n}(0) - v_{b^*(\gamma_n)}(0)| < 2\varepsilon. \end{align}\] By letting \(\varepsilon\downarrow 0\), we show \(v_{b^*}(0) = \xi_0\). Finally, we know that \(\underline{b}<b^*(\gamma)< \hat{b}\) for every \(\gamma<0\) by Step 1, therefore \(\underline{b}\le \lim_{\gamma\uparrow 0}b^*(\gamma) = b^*\le \hat{b}\). The strict inequality \(\underline{b}< b^*\) follows by the increasing monotonicity of \(b^*(\gamma)\). ◻
Proof. Given the existence of \(b^*\in(\underline{b},\hat{b}]\) satisfying \(v_{b^*}(0)=\xi_0\) by Lemma 42ii., we first prove that \(v_{b^*}'(x)\geq 1\) for every \(x\in[0,b^*]\).
From Lemma 42ii., define \(b^{\gamma}\) for every \(\gamma<0\) by
\[\begin{align}
\label{eq:quad95perturb950} b^{\gamma}:=b^*(\gamma)= \inf\{b\in(\underline{b},\hat{b}): v_{b,\gamma}(0) = \xi_0\}.
\end{align}\tag{50}\] Furthermore, denote by \(v_{b^{\gamma},\gamma}\in C^2((0,b^{\gamma})\cup(b^{\gamma},\infty))\cap C^1(\mathbb{R}_+)\) the solution of 38 (when \(b=b^{\gamma}\)).
The proof is achieved in two steps.
We star with perturbation analysis with a fixed \(\gamma\in(-\rho,0)\). We claim \[\begin{align}
\label{eq:claim95195shooting} v_{b^\gamma,\gamma}'(b^\gamma)(x)\ge 1,\forall x\in(0,b^{\gamma}].
\end{align}\tag{51}\] From the boundary conditions \(v_{b^{\gamma},\gamma}'(b^{\gamma})=1\) and \(v_{b^{\gamma},\gamma}(b^{\gamma})= \psi^+(b^{\gamma})\), the \(C^2\) property of \(v_{b^{\gamma},\gamma}\) and nonlinear ODE 38 gives \(v''_{b^{\gamma},\gamma}(b^{\gamma}) =
2\gamma\sigma^2(b^{\gamma})\psi^+(b^{\gamma})<0.\) Argue by contradiction, assume 51 doesn’t hold. Then combined with the fact that \(v_{b^\gamma,\gamma}'(b^\gamma)=1\),
we have the existence of \[x_1:=\sup\{ x\in(0,b^\gamma): v'_{b^\gamma,\gamma}(x) = 1 \}.\] In the following, we prove that the existence of \(x_1\) leads to a contradiction. i. Case
\(x_1 \in (\underline{b},{b}^\gamma)\). Since \(v_{b^\gamma,\gamma}'(x_1)= 1\) (noting that \(v_{b^\gamma,\gamma}\in C^1(\mathbb{R}_+)\)), it
follows that \(v_{b^\gamma,\gamma}'(x) > 1\) for every \(x\in(x_1,b^\gamma)\). Furthermore, since \(\psi^+(x)-x\) decreases on \([x_1,b^\gamma]\subset [\underline{b},\hat{b}]\) (see Assumption 15ii. and \(x_1 \in
(\underline{b},{b}^\gamma)\)) and \(v_{b^\gamma,\gamma}(b^\gamma)=\psi^+(b^\gamma)\), hence \[\begin{align} v_{b^\gamma,\gamma}(x_1) \le \psi^+(b^\gamma)-(b^\gamma-x_1)\le \psi^+(x_1).
\end{align}\] Moreover, since \(x_1 \in (\underline{b},\hat{b})\) and \(v_{b^\gamma,\gamma}(0)= \xi_0\) (by definition of \(b^\gamma\)), Lemma42i. ensures that \(v_{b^\gamma,\gamma}(x_1) \ge x_1 +\xi_0 \ge \psi^-(x_1)\), so that \(Q(v_{b^\gamma,\gamma}(x_1);x_1)\le 0\) by recalling that \(Q(\cdot;x_1)\) defined in Assumption 15ii. This
implies that \[\begin{align} v''_{b^\gamma,\gamma}(x_1) = \frac{2}{\sigma^2(x_1)}\Big(\gamma v_{b^\gamma,\gamma}(x_1) +\frac{Q(v_{b^\gamma,\gamma}(x_1);x_1)}{v_{b^\gamma,\gamma}(x_1)}\Big)\le \frac{2}{\sigma^2(x_1)}\gamma
(x_1+\xi_0)<0.
\end{align}\] This together with the fact that \(v''_{b^{\gamma},\gamma}(b^{\gamma})<0\) contradicts Lemma 38ii.
ii. Case \(x_1\in(0,\underline{b}]\). Recalling \(x_1=\sup\{ x\in(0,b^\gamma): v'_{b^\gamma,\gamma}(x) = 1 \}\) and the fact that \(v_{b^\gamma,\gamma}'(b^\gamma)=1\) and \(v_{b^\gamma,\gamma}''(b^\gamma)<0\), we employ Lemma 38ii. to obtain that \(v''_{b^\gamma,\gamma}(x_1)\geq 0\), which can be rewritten by \[\begin{align} \label{eq:quad95perturb} v''_{b^\gamma,\gamma}(x_1)= \frac{2}{\sigma^2(x_1)}\frac{1}{v_{b^\gamma,\gamma}(x_1)}\widetilde{Q}_{\gamma}(v_{b^\gamma,\gamma}(x_1);x_1)\geq 0, \end{align}\tag{52}\] where \(\widetilde{Q}_\gamma(\psi;x_1):=(\rho+\gamma) \psi^2-\mu(x_1)\psi + \frac{\@fontswitch\relax\mathcal{R}}{2}\sigma^2(x_1)\) (which is well-defined by Assumption 15ii. with the fact that \(\rho>\rho+\gamma>0\)) and we have used the nonlinear ODE of \(v_{b^\gamma,\gamma}\) given in 38 .
Lemma 42i. together with Assumption 15iii. (and \(x_1\in(0,\underline{b}]\subset [0,\hat{b}]\)) ensures that \(v_{b^\gamma,\gamma}(x_1) \geq x_1+\xi_0\geq \psi^-(x_1)\). Furthermore, recalling that \(\gamma\) satisfies \(\rho>\rho+\gamma>0\), we have \[\begin{align} v_{b^\gamma,\gamma}(x_1) \geq& \frac{\mu(x_1) - \sqrt{\mu^2(x_1)-2{\@fontswitch\relax\mathcal{R}}\rho \sigma^2(x_1)}}{2\rho}\tag{53} \\ >&\frac{\mu(x_1) - \sqrt{\mu^2(x_1)-2{\@fontswitch\relax\mathcal{R}}(\rho+\gamma) \sigma^2(x_1)}}{2(\rho+\gamma)} =:\widetilde{\psi^-}(x_1),\tag{54} \end{align}\] where we note that RHS of 53 equals \(\psi^-(x_1)\) and that the last term \(\widetilde{\psi^-}(x_1)\) is the smaller one among two roots of the quadratic function \(\widetilde{Q}_\gamma(\cdot;x_1) = 0\).
Since \(\widetilde{Q}_{\gamma}(v_{b^\gamma,\gamma}(x_1);x_1)\geq0\) (see 52 ), from 54 it follows that \[\begin{align} \label{eq:quad95perturb3} v_{b^\gamma,\gamma}(x_1)\geq \frac{\mu(x_1) + \sqrt{\mu^2(x_1)-2{\@fontswitch\relax\mathcal{R}}(\rho+\gamma) \sigma^2(x_1)}}{2(\rho+\gamma)}=:\widetilde{\psi^+}(x_1), \end{align}\tag{55}\] where the right-hand side \(\widetilde{\psi^+}(x_1)\) is the other (larger) root of \(\widetilde{Q}_\gamma(\cdot;x_1)\).
Now we claim that for \(x\) fixed, if \((\psi^+)'(x) \ge 1\) then \((\widetilde{\psi^+})'(x) \ge 1\). By taking derivative of \(\psi^+(x)\) w.r.t. \(x\), we have \[\begin{align} (\psi^{+})'(x) = \frac{1}{2\rho}\mu'(x) + \frac{\mu(x)\mu'(x)}{2\rho\sqrt{\mu(x)^2-2{\@fontswitch\relax\mathcal{R}}\rho\sigma^2(x)}} - \frac{{\@fontswitch\relax\mathcal{R}}\sigma(x)\sigma'(x)}{\sqrt{\mu(x)^2-2{\@fontswitch\relax\mathcal{R}}\rho\sigma^2(x)}}. \end{align}\] When \((\psi^+)'(x) \ge 1\), it is clear that \(\mu'(x)>0\) since \(\sigma'(x)\ge 0\) by Assumption 15. Therefore the first term and the last term in above equation are decreasing w.r.t \(\rho\). It remains to show the same monotonicity for the second term in \((\psi^{+})'(x)\). Take derivative of the function \(h(\rho;x): = \rho\sqrt{\mu(x)^2-2{\@fontswitch\relax\mathcal{R}}\rho\sigma^2(x)}\) w.r.t \(\rho\), we have \[\begin{align} \partial_{\rho} h(\rho;x) = \sqrt{\mu(x)^2-2{\@fontswitch\relax\mathcal{R}}\rho\sigma^2(x)} - \frac{{\@fontswitch\relax\mathcal{R}}\rho \sigma^2(x)}{\sqrt{\mu(x)^2-2{\@fontswitch\relax\mathcal{R}}\rho\sigma^2(x)}} \ge 0, \end{align}\] where the inequality is followed by Assumption 15i. with \((x_2,x_1)\subset [0,\underline{b}]\). As a result, we show that \(1 \le (\psi^{+})'(x)\le (\widetilde{\psi^+})'(x)\) since \(\rho+\gamma < \rho\) for \(\gamma<0\). Combined with Assumption 15ii., which says that \(\psi^+(x)-x\) increases on \([0,\underline{b}]\), we have \(\widetilde{\psi^+}(x)-x\) increases on \([0,\underline{b}]\).
To proceed, recalling that \(v'_{b^\gamma,\gamma}(x_1) = 1\) and \(v''_{b^\gamma,\gamma}(x_1)\geq 0\) (see 52 ), standard arguments by using a proof by contradiction shows that there exists \(x_2\) defined as \(x_2:=\sup\{ x\in(0,x_1): v'_{b^\gamma,\gamma}(x) = 1 \}\). Since \(v'_{b^\gamma,\gamma}(x)<1\) for every \(x\in(x_2,x_1)\) (from the definition of \(x_2\) and the fact that \(v'_{b^\gamma,\gamma}(x_1) = 1\) and \(v''_{b^\gamma,\gamma}(x_1)\geq 0\)), the following hold: \(v'_{b^\gamma,\gamma}(x_2)=1\) and \[\begin{align} v_{b^\gamma,\gamma}(x_2)&>v_{b^\gamma,\gamma}(x_1)-(x_1-x_2)\nonumber\\ &\geq \frac{\mu(x_1) + \sqrt{\mu^2(x_1)-2{\@fontswitch\relax\mathcal{R}}(\rho+\gamma) \sigma^2(x_1)}}{2(\rho+\gamma)}-(x_1-x_2)\nonumber\\ &\geq \frac{\mu(x_2) + \sqrt{\mu^2(x_2)-2{\@fontswitch\relax\mathcal{R}}(\rho+\gamma) \sigma^2(x_2)}}{2(\rho+\gamma)}, \label{eq:quad95perturb4} \end{align}\tag{56}\] where the second inequality follows from 55 and the last inequality follows from the fact that \(\widetilde{\psi^+}(x)-x\) increases on \((x_2,x_1)\) since \((x_2,x_1)\subset [0,\underline{b}]\).
Recalling the nonlinear ODE 38 , the inequality given in 56 ensures that \[v''_{b^\gamma,\gamma}(x_2)=\frac{2}{\sigma^2(x_2)}\frac{1}{v_{b^\gamma,\gamma}(x_2)}\widetilde{Q}_{\gamma}(v_{b^\gamma,\gamma}(x_2);x_2)>0.\] This together with 52 contradicts Lemma 38. (Lemma 2.2ii. requires \(v''_{b^\gamma,\gamma}(x_1)>0\), but we can see it cannot be the case that \(v''_{b^\gamma,\gamma}(x_2)>0\), \(v''_{b^\gamma,\gamma}(x_1)=0\) and \(v''_{b^\gamma,\gamma}(b^{\gamma})<0\)).
By deriving contradictions for the above two cases, we have 51 .
Recall \(b^\gamma\) for every \(\gamma<0\) given in 50 and let \(b^*\) be defined by \(b^*=\lim_{\gamma\uparrow 0} b^\gamma\) (see Lemma 42ii.). We claim that \(v'_{b^*}(x)\ge 1\) for every \(x\in (0,b^*]\).
To that end, let us consider an increasing sequence \((\gamma_n)_{n\in\mathbb{N}}\) such that \(-\rho <\gamma_n<0\) for every \(n\in\mathbb{N}\) and \(\lim_{n\rightarrow \infty } \gamma_n = 0\). Since \(\lim_{j\rightarrow \infty} b^{\gamma_n} = b^*\) and \((b^{\gamma_n})_{n\in\mathbb{N}}\) is a increasing sequence (see Lemma 42ii.), the following hold: \(b^*\geq b^{\gamma_n} \geq b^{\gamma_0}\) for every \(n\in \mathbb{N}\).
Combining this with Step 1. (along with \(v_{b^{\gamma_n},\gamma_n}\in C^1(\mathbb{R}_+)\) for every \(n\in\mathbb{N}\)) ensures that for every \(n\in\mathbb{N}\) and every \(x\in (0,b^*]\), \[\begin{align} \label{eq:quad95perturb5} v_{b^{\gamma_n},\gamma_n}'(x) \ge 1. \end{align}\tag{57}\]
Fix \(y\in(0,b^{\gamma_0}]\subset (0,b^*]\). For arbitrary \(\varepsilon>0\), then Lemma 39iii. ensures that there exists \({N}_\varepsilon\in \mathbb{N}\) such that \(|v'_{b^{\gamma_n},\gamma_n}(x) - v'_{b^{\gamma_n}}(x)|<\varepsilon\) for every \(n\geq {N}_\varepsilon\). Furthermore, Lemma 40iii. ensures that there exists \(\widetilde{N}_\varepsilon\in \mathbb{N}\) such that \(|v'_{b^*}(x) - v'_{b^{\gamma_n}}(x)|<\varepsilon\) for every \(n\geq \widetilde{N}_\varepsilon\). Therefore, from 57 , it follows that for every \(n\geq \widetilde{N}_\varepsilon\vee N_\varepsilon\), \(v'_{b^*}(x) >v'_{b^{\gamma_n}}(x) -\varepsilon> v'_{b^{\gamma_n},\gamma_n}(x)- 2\varepsilon\geq 1- 2\varepsilon.\) By letting \(\varepsilon\rightarrow0\), we conclude that \(v_{b^*}'(x)\ge1\) for \(y\in (0, b^{\gamma_0}]\).
It remains to show \(v_{b^*}'(y)\ge1\) on \((b^{\gamma_0},b^*]\).
Assume the contrary: \(\min_{y\in[b^{\gamma_0},b^*]} v_{b^*}'(y) <1\). Then \(v_{b^*}'(y) -1\) takes its minimum at \(x_0\in(b^{\gamma_0},b^*)\)
because \(v'_{b^*}(b^*) = 1\) and \(v'_{b^*}(b^{\gamma_0})\ge1\) by above analysis. Then we have that \(v_{b^*}''(x_0) = 0\) and \(v'_{b^*}(x_0)<1\). By the continuity of \(v'_{b^*}\), let \(\epsilon=(1-v'_{b^*}(x_0))/2\) there exists \(x_0'\) in a local neighborhood of \(x_0\) such that \(v_{b^*}''(x_0') > 0\) and \(v_{b^*}'(x_0')<
v'_{b^*}(x_0) + \epsilon<1\). Hence \[g_{b^*}(x_0') = \dfrac{v'_{b^*}(x_0')}{v_{b^*}(x_0')} < \frac{1}{x_0'+\xi} < \frac{1}{\psi^{-}(x_0')},\] where the first inequality follows
Lemma 42i. that \(v_{b^*}(x_0')\ge x_0'+\xi\) since \(v_{b^*}(0)=\xi\), while the last
inequality follows Assumption 15iii. Recall the ODE 38 of \(v_{b^*}\), we have \[\begin{align} \frac{1}{2}\sigma^2(x_0) v''_{b^*}(x_0') = \frac{1}{v_{b^*}(x_0')}Q(v_{b^*}(x_0');x_0') > 0.
\end{align}\] Combined with the above arguments, we conclude that \(g_{b^*}(x_0') < 1/\psi^+(x_0')\). This contradicts with Lemma 40i. which gives \(g_{b^*}(x_0')\ge{1}/{\psi^+(x_0')}\) as \(x_0>b^{\gamma_0}>\underline{b}\). As a result, we see \(v_{b^*}'(y)\ge1\) on \((b^{\gamma_0},b^*]\).
Finally, we show that \(b^*<\hat{b}\), where \(\hat{b}\) defined in Assumption 15iii. We only need to show \(b^* \neq \hat{b}\) since \(b^*\le \hat{b}\) by Lemma 42ii. When \(b^* = \hat{b}\), by definition of \(\hat{b}\), we have \(v_{b^*}(0)=\xi_0\) while \(v_{b^*}(b^*) = \psi^+(b^*) = b^* + \xi_0\). Together with our previous result that \(v_{b^*}'(x)\ge 1\) on \((0,b^*]\), we conclude \(v_{b^*}'(x) \equiv 1\) on \((0,b^*]\), which cannot be the case due to the structure of ODE 38 . Therefore, we have \(b^*<\hat{b}\). ◻
Proof. Theorem 19 is proved in two steps. As the proof for the case \(x=0\) is obvious, we assume \(x\in{\mathbb{R}_+}\). For simplicity, we omit the superscript \(x\) in \(X^{x,D}\) and \(\tau^{x,D}\), as defined in 18 and 19 .
Recall that \(J^*\) is defined in 21 , we claim that for every \(x\in{\mathbb{R}_+}\), \[\begin{align} \tag{58} v^*(x)\geq& \sup_{D\in {\@fontswitch\relax\mathcal{A}}^x}\inf_{\theta \in \Theta_D}\mathbb{E}^{\theta}\bigg[\int_{0}^{\tau^{D}}e^{-\rho t} \Big( dD_t + \frac{\theta_t^2}{2{\@fontswitch\relax\mathcal{R}}} v^*(X_t^{D}) dt\Big) + e^{-\rho \tau^{D}}\xi_0 \bigg],\\ v^*(x) \ge& J^*(x).\tag{59} \end{align}\]
Set an arbitrary \(D\in {\@fontswitch\relax\mathcal{A}}^x\) and define the stopping times \((\hat{\tau}_n)_{n\in \mathbb{N}}\) by \[\begin{align} \label{eq:stop95bdd95X} \hat{\tau}_n:= \inf\{t\ge 0: X^{D}_t\notin [1/n,n] \}\wedge n. \end{align}\tag{60}\] Recall the structure of the operator \(\mathcal{L}\) from 22 , and define the measure \(\mathbb{Q}^*\) via 8 with the Girsanov kernel \(\theta^*:=(\theta^*(D)_s)_{s\geq 0}\) given by \[\begin{align} \theta^*_s:= -{\@fontswitch\relax\mathcal{R}}\frac{\sigma(X_s^{D})(v^*)'(X_s^{D})}{v^*(X_s^{D})}. \label{eq:worst95kernel} \end{align}\tag{61}\] Note that by 43 and Proposition 37, \((v^*)'\) is bounded on \([0,b^*]\) and is identical to \(1\) on \((b^*,\infty)\) while \(v^*\) is bounded below by \(\xi_0\) on \({\mathbb{R}_+}\). Following similar arguments of in Theorem 9, it follows that \(\theta^*\in \Theta^D\). From this, we can denote by \(\mathbb{Q}^*\) the corresponding probability measure obtained from \(\theta^*\).
An application of Dynkin formula into the process \(e^{-\rho t} v^*(X_t^D)\) on the random time interval \([0,\tau^D\wedge\hat{\tau}_n]\) under \(\mathbb{Q}^*\) (corresponding to \(\theta^*\)) ensures that \[\begin{align} &e^{-\rho (\tau^D\wedge \hat{\tau}_n )} v^*(X_{\tau^D\wedge \hat{\tau}_n }^D) -v^*(x) =\int_0^{\tau^D\wedge \hat{\tau}_n } e^{-\rho s} \; \hat{\@fontswitch\relax\mathcal{L}}^{\theta^*}v^*(X_s^D)ds\nonumber\\ & +\int_0^{\tau^D\wedge \hat{\tau}_n } e^{-\rho s} \sigma(X_{s}^D)(v^*)'(X_{s}^D) d W^{\mathbb{Q}^*}_s- \int_0^{\tau^D\wedge \hat{\tau}_n }e^{-\rho s} (v^*)'(X_{s}^D) dD_s,\label{eq:thm95veri95pf951} \end{align}\tag{62}\] where \(\hat{\@fontswitch\relax\mathcal{L}}^{\theta^*}\) is the infinitesimal operator under \(\mathbb{Q}^*\), i.e., \[\begin{align} \hat{\@fontswitch\relax\mathcal{L}}^{\theta^*}v^*(X_s^D):=\Big(\frac{\sigma^2}{2}(v^*)''+ (\mu+\sigma\theta_s^*)(v^*)' -\rho v^*\Big)(X_{s}^D). \end{align}\] Based on the definitions of the nonlinear operator \({\@fontswitch\relax\mathcal{L}}\) in 22 , the process \(\theta^*\) in 61 , and given that \(v^*(X_s^D)>0\) for \(s\in[0,\tau^D)\) and \(v^*\) solves 24 , it follows that for every \(s\in[0,\tau^D)\): \[\begin{align} \hat{\@fontswitch\relax\mathcal{L}}^{\theta^*}v^*(X_s^D)= {\@fontswitch\relax\mathcal{L}} v^*(X_s^D)-\frac{(\theta_s^*)^2}{2{\@fontswitch\relax\mathcal{R}}}v^*(X_s^D)\leq -\frac{(\theta_s^*)^2}{2{\@fontswitch\relax\mathcal{R}}}v^*(X_s^D). \end{align}\] Furthermore, Since \(v^*\in C^2({\mathbb{R}_+})\) (see Theorem 18), \(\sigma\in C^1(\mathbb{R}_+)\) (see Assumption 15i.) and \(X^{x,D}\) is bounded in \([0,\tau^D\wedge \hat{\tau}_n ]\) (see 60 ), the Brownian local martingale term given in 62 is martingales. Lastly, since \((v^*)'(x)\geq 1\) for \(x\in {\mathbb{R}_+}\) and \(v^*(0)=\xi_0\), it follows that \(v^*(X_{\tau_D\wedge \hat{\tau}_n }^D)\geq \xi_0\). Taking the expectation over 62 under \(\mathbb{Q}^*\) gives \[\begin{align} v^*(x) \geq & \mathbb{E}^{\theta^*}\bigg[\int_0^{(\tau_D\wedge \hat{\tau}_n )} e^{-\rho s}d D_s +\int_0^{(\tau_D\wedge \hat{\tau}_n )}e^{-\rho s}\frac{(\theta^*_s)^2}{2{\@fontswitch\relax\mathcal{R}}}v^*(X^D_s) ds +e^{-\rho (\tau_D\wedge \hat{\tau}_n )} \xi_0 \bigg]. \end{align}\] Since \(\tau_D\wedge \hat{\tau}_n \rightarrow \tau_D\) as \(n\rightarrow \infty\) (see 60 ), an application of Fatou’s lemma together with \(\theta^* \in \Theta^D\) (see 61 ) ensures that \[\begin{align} v^*(x) \geq & \mathbb{E}^{\theta^*}\bigg[\int_0^{\tau_D} e^{-\rho s}d D_s+\int_0^{\tau_D} e^{-\rho s}\frac{(\theta_s^*)^2}{2{\@fontswitch\relax\mathcal{R}}}v^*(X^D_s) ds+e^{-\rho \tau_D} \xi_0 \bigg] \\ \geq & \inf_{\theta \in \Theta_D} \mathbb{E}^{\theta}\bigg[\int_0^{\tau_D} e^{-\rho s}d D_s+\int_0^{\tau_D} e^{-\rho s}\frac{\theta_s^2}{2{\@fontswitch\relax\mathcal{R}}}v^*(X^D_s) ds+e^{-\rho \tau_D} \xi_0 \bigg]. \nonumber \end{align}\] As the inequality holds for every \(D\in {\@fontswitch\relax\mathcal{A}}^x\), the claim 58 holds. To show 59 , the proof proceeds similarly by applying Dynkin formula to the process \(e^{-\rho t + \int_0^{t}({\theta}^*_s)^2/2{\@fontswitch\relax\mathcal{R}}ds} v^*(X_t^D)\) on the random time interval \([0,\tau^D\wedge\hat{\tau}_n]\) and take expectation. We obtain that \[\begin{align} v^*(x) &\geq \mathbb{E}^{\theta^*}\bigg[\int_0^{\tau_D} e^{-\rho s+\frac{1}{2{\@fontswitch\relax\mathcal{R}}}\int_0^{s} (\theta_u^*)^2 du}d D_s+e^{-\rho \tau_D+\frac{1}{2{\@fontswitch\relax\mathcal{R}}}\int_0^{\tau_D} (\theta_u^*)^2 du} \xi_0 \bigg]\geq J^{\theta^*}(x;D), \end{align}\] where \(J^{\theta}\) is defined in 20 . The last term is greater than \(\inf_{\theta\in \Theta^D} J^{\theta}(x;D)\). Taking supreme over \({D\in\mathcal{A}^x}\) on both sides of the above equation gives 59 .
Let \(D^*=D(b^*)\) be the \(b^*\)-threshold strategy. We show that \[\begin{align} \tag{63} v^*(x)\leq& \inf_{\theta \in \Theta_{D^*}}\mathbb{E}^{\theta}\bigg[\int_{0}^{\tau^{D^*}}e^{-\rho t} \Big(dD^*_t + \frac{\theta_t^2}{2{\@fontswitch\relax\mathcal{R}}} v^*(X_t^{D^*}) dt \Big) + e^{-\rho \tau^{D^*}}\xi_0 \bigg],\\ \tag{64} v^*(x) \le& J^*(x). \end{align}\] The analysis is divided into two cases, i.e., \(x\in(0,b^*]\) or \(x\in(b^*,\infty)\). Before starting the proof, we first note that from the definition of \(b^*\)-threshold strategy (see Definition 14), the following properties hold for every \(s\in(0,\tau^{D^*}]\): \[\begin{align} \label{eq:skorokhod95reflect} X^{D^{*}}_s \in [0, b^*],\;\;\quad dD^*_s>0\quad only when X_s^{D^*}=b^*. \end{align}\tag{65}\] Furthermore, if \(x\in(0,b^*]\), the above properties hold for every \(s\in[0,\tau^{D^*}]\). We first consider the case \(x\in(0,b^*]\). Let \(\theta \in \Theta({D^*})\) and \(\mathbb{Q}\) be the corresponding probability measure induced by the kernel \(\theta\). Furthermore, as in 60 , we define for every \(n\in \mathbb{N}\) by \(\hat{\tau}_n:= \inf\{t\ge 0: X^{D^*}_t\notin [1/n,n] \}\wedge\tau^{D^*}.\) An application of Dynkin formula to the process \(e^{-\rho t} v^*(X_t^{D^*})\) on the random time interval \([0,\hat{\tau}_n]\) under \(\mathbb{Q}\) gives that \[\begin{align} \label{eq:thm95veri95pf9510} \begin{aligned} v^*(x) &=\mathbb{E}^\theta \bigg[e^{-\rho (\hat{\tau}_n)} v^*(X_{\hat{\tau}_n}^{D^*})-\int_0^{\hat{\tau}_n} e^{-\rho s} (\hat{\@fontswitch\relax\mathcal{L}}^{\theta}v^*(X_s^{D^*})ds-dD^*_s ) \bigg], \end{aligned} \end{align}\tag{66}\] where \(\hat{\@fontswitch\relax\mathcal{L}}^{\theta}\) is the infinitesimal operator under \(\mathbb{Q}^*\), i.e., \[\begin{align} \hat{\@fontswitch\relax\mathcal{L}}^{\theta}v^*(X_s^{D^*}):=\Big(\frac{1}{2}\sigma^2(v^*)''(X_{s}^{D^*}) + (\mu+\sigma\theta_s)(v^*)' -\rho v^*\Big)(X_{s}^{D^*}), \end{align}\] and we use the fact that \(\mathbb{E}^{\theta}[\int_0^{\tau^{D^*}\wedge \hat{\tau}_n } e^{-\rho s} \sigma(X_{s}^{D^*})(v^*)'(X_{s}^{D^*}) d W^{\mathbb{Q}}_s]=0\) and that \(dD^*_s>0\) only when \((v^*)'(X_{s}^{D^*})=1\) for \(s\in[0,\tau^{D^*}]\). From 65 and the fact that \({\@fontswitch\relax\mathcal{L}}v^*(x)=0\) for every \(x\in(0,b^*]\) (see 24 with the free boundary \(b^*\)), it follows that for every \(s\in [0,\tau^{D^*}]\), \({\@fontswitch\relax\mathcal{L}}v^*(X_s^{D^*})=0\). Recall the definition of \({\@fontswitch\relax\mathcal{L}}\) in 22 , we obtain that for every \(s\in [0,\tau^{D^*}]\) \[\begin{align} \label{eq:claim3} \hat{\@fontswitch\relax\mathcal{L}}^{\theta}v^*(X_s^{D^*})\geq {\@fontswitch\relax\mathcal{L}}v^*(X_s^{D^*})- \frac{(\theta_s)^2}{2{\@fontswitch\relax\mathcal{R}}} v^*(X_s^{D^*}) = - \frac{(\theta_s)^2}{2{\@fontswitch\relax\mathcal{R}}} v^*(X_s^{D^*}). \end{align}\tag{67}\] By 67 and 66 , we hence obtain that \[\begin{align} \label{eq:thm95veri95pf952} \begin{aligned} v^*(x)\leq \mathbb{E}^\theta \bigg[\int_0^{\hat{\tau}_n} e^{-\rho s} \left(\frac{(\theta_s)^2}{2{\@fontswitch\relax\mathcal{R}}} v^*(X_s^{D^*})ds+dD^*_s \right)+e^{-\rho (\hat{\tau}_n)} v^*(X_{\hat{\tau}_n}^{D^*}) \bigg]. \end{aligned} \end{align}\tag{68}\] From \(v^*\in C^2({\mathbb{R}_+})\) and the fact that \(X^{D^{*}}_s \in [0, b^*]\) for every \(s\in [0,\tau^{D^*}]\) (see 65 ), it follows that \(\sup_{s\in [0,\tau^{D^*}]}|v^*(X_s^{D^*})|<\infty.\) Hence, an application of the DCT (applicable to the last term of right-hand side of 68 ) and monotone convergence theorem (applicable to the first term of right-hand side of 68 ), together with \(\tau^{D^*}\wedge \hat{\tau}_n\rightarrow \tau^{D^*}\) as \(n\rightarrow \infty\) into 68 ensures that \[\begin{align} \label{eq:thm95veri95pf953} v^*(x) &\leq \mathbb{E}^\theta \bigg[\int_0^{\tau^{D^*}} e^{-\rho s} \left(\frac{(\theta_s)^2}{2{\@fontswitch\relax\mathcal{R}}} v^*(X_s^{D^*})ds+dD^*_s \right)+e^{-\rho \tau^{D^*}} \xi_0 \bigg], \end{align}\tag{69}\] where we have used that \(v^*(X_{\tau^{D^*}}^{D^*})=v^*(0)=\xi_0\) (see 24 ). As the inequality holds for every \(\theta \in \Theta_{D^*}\), the claim 63 holds for every \(x\in(0,b^*]\).
It remains to prove the case \(x\in (b^*,\infty)\). Note that for every \(x\in (b^*,\infty)\), \(v^*(x)=x-b^*+\psi^+(b^*)\) (see 43 ). Furthermore, since the strategy \(D^*=D(b^*)\) starts with an instantaneous increase with amount \(x-b^*\), it follows that \[\begin{align} &\mathbb{E}^\theta \bigg[\int_0^{\tau^{D^*}} e^{-\rho s} \Big(\frac{(\theta_s)^2}{2{\@fontswitch\relax\mathcal{R}}} v^*(X_s^{x,D^{*}})ds+dD^*_s \Big)+e^{-\rho \tau^{D^*}} \xi_0 \bigg]\geq v^*(x) \end{align}\] holds for every \(\theta \in \Theta_{D^*}\), where we have used the inequality 69 (since it holds when \(x=b^*\)) and \(v^*(b^*)=\psi^+(b^*)\).
To show 64 , similar to the analysis in Step 1, an application of Dynkin formula to the process \(e^{-\rho t \int_0^{t}({\theta}_s)^2/(2{\@fontswitch\relax\mathcal{R}}) ds} v^*(X_t^{D^*})\) gives \[\begin{align} v^*(x) \le \inf_{\theta\in{\Theta_{D}^*}} J^{\theta}(x;D^*) \le \sup_{D\in\mathcal{A}^x} \inf_{\theta\in{\Theta_{D}}} J^{\theta}(x;D) \le J^*(x),\quad x\in{\mathbb{R}_+}. \end{align}\] This completes the proof. ◻
We provide some observations used for the proof of Theorem 21.
Lemma 43. Let \((\tau,D)\) satisfy the condition in Setting 20 i. For \({\@fontswitch\relax\mathcal{R}}\in(0,1)\), let \(V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\) be defined as in Setting 20 iii. Moreover, let \(V^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}})\) be the EZ singular control utility defined as in 4 under setting \((\tau,D,\xi_\tau({\@fontswitch\relax\mathcal{R}}))\) given in Setting 20. If \(\xi^0_\tau\) given in Setting 20ii satisfies \(\xi^0_{\tau} > C\) \(\mathbb{P}\)-a.s. with some constant \(C > 0\), then the following properties hold for every \({\@fontswitch\relax\mathcal{R}}\in(0,1)\):
\((V^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}}), Z^D({\@fontswitch\relax\mathcal{R}}))\) is the unique \(L^2\)-solution to 5 under \((\tau,D,\xi_\tau({\@fontswitch\relax\mathcal{R}}))\); \(V^{\text{EZ},D}_t({\@fontswitch\relax\mathcal{R}}) \ge C\) \(\mathbb{P}\)-a.s.for \(t \ge 0\); and \(\mathbb{E}[\sup_{t \ge 0} (V_t^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}}))^{\frac{2}{1-{\@fontswitch\relax\mathcal{R}}}}] < \infty\).
\(V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}) = (V_t^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}}))^{\frac{1}{1-{\@fontswitch\relax\mathcal{R}}}}\) for \(t \ge 0\). In addition, \((V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}),{\mathbb{Z}}^D({\@fontswitch\relax\mathcal{R}}))\) with \({\mathbb{Z}}^D({\@fontswitch\relax\mathcal{R}}): = (V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}))^{\@fontswitch\relax\mathcal{R}}Z^D({\@fontswitch\relax\mathcal{R}})/(1-{\@fontswitch\relax\mathcal{R}})\) solves \[\begin{align} \label{eq:Vrob95BSDE} V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}) = \xi^0_{\tau} + \int_{t\wedge\tau}^{\tau} e^{-\rho s} dD_s - \frac{{\@fontswitch\relax\mathcal{R}}}{2} \int_{t\wedge\tau}^{\tau} \frac{({\mathbb{Z}}^D_s({\@fontswitch\relax\mathcal{R}}))^2}{V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}})} ds - \int_{0}^{\tau} \mathbb{Z}^D_s({\@fontswitch\relax\mathcal{R}}) dW_s. \end{align}\qquad{(9)}\]
\(V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\) admits the following representation: for \(t\in[0,T]\) \[\begin{align} \label{eq:vrobproof951} V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}) = \mathop{\mathrm{ess\,inf}}_{\theta\in\Theta} \mathbb{E}^{\theta}_t\bigg[ \int_{t\wedge\tau}^{\tau}\bigg( e^{-\rho s} dD_s+ \frac{V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}) \theta_s^2}{2{\@fontswitch\relax\mathcal{R}}} ds \bigg) + \xi^0_{\tau}\bigg], \end{align}\qquad{(10)}\] where \(\Theta\) is defined in Definition 7 and does not depend on \({\@fontswitch\relax\mathcal{R}}\).
Let \(t\ge0\). There exists some \({\@fontswitch\relax\mathcal{F}}_t\)-measurable \(\overline{c}'_t>0\) (which does not depend on \({\@fontswitch\relax\mathcal{R}}\)) such that \(\mathbb{E}[\sup_{t\ge0} \overline{c}'_t]<\infty\) and \[\begin{align} \label{eq:Vrob95lb} \ln\big(V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\big) \ge \frac{\ln(K_t^D) - {\@fontswitch\relax\mathcal{R}}\overline{c}'_t}{1-{\@fontswitch\relax\mathcal{R}}},\quad \text{ \mathbb{P}-a.s. for }t \ge 0. \end{align}\qquad{(11)}\]
Proof. Part i. follows from Theorem 5 ii., while Part ii. follows from Theorem 9 ii.. Hence, we start with proving Part iii..
Let \(t\ge0\). For any \(\theta\in\Theta^D({\@fontswitch\relax\mathcal{R}})\), we have \(V_t^{\theta,D}({\@fontswitch\relax\mathcal{R}})\ge V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\). By the definition 9 of \(V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\), \[\begin{align} V^{\text{rob}}_t({\@fontswitch\relax\mathcal{R}}) \geq \mathop{\mathrm{ess\,inf}}_{\theta\in\Theta^D({\@fontswitch\relax\mathcal{R}})} \mathbb{E}^{\theta}_t\bigg[ \int_{t\wedge\tau}^{\tau}\bigg( e^{-\rho s} dD_s+ \frac{V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}) \theta_s^2}{2{\@fontswitch\relax\mathcal{R}}} ds \bigg) + \xi^0_{\tau}\bigg]. \end{align}\] We get the inequality in the \(\ge\) direction of ?? by noticing that \(\Theta^D({\@fontswitch\relax\mathcal{R}})\subset\Theta\). For the reverse inequality, we note that for every \(\theta \in \Theta\) \[\begin{align} -\frac{{\@fontswitch\relax\mathcal{R}}}{2} \frac{\mathbb{Z}^D({\@fontswitch\relax\mathcal{R}})^2}{V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})} = \inf_{\theta } \Big\{ \frac{V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\theta^2}{2{\@fontswitch\relax\mathcal{R}}} + \mathbb{Z}^D({\@fontswitch\relax\mathcal{R}})\theta \Big\}\leq \frac{V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\theta^2}{2{\@fontswitch\relax\mathcal{R}}} + \mathbb{Z}^D({\@fontswitch\relax\mathcal{R}})\theta. \end{align}\]
Substituting this into ?? , taking expectations under \(\mathbb{Q}_t^{\theta}\) for each \(\theta \in \Theta\), and then taking the essential infimum over \(\theta\in\Theta\) yield the inequality in the \(\le\) direction of ?? . This completes the argument.
We now prove Part iv. Let \(t\ge0\). By definitions 4 and 10 , we have \[\begin{align} &K_t^D - V^{\text{EZ},D}_t({\@fontswitch\relax\mathcal{R}})\nonumber\\ &\quad = \mathbb{E}_t\bigg[\xi^0_{\tau} - (\xi^0_{\tau})^{1-{\@fontswitch\relax\mathcal{R}}} + \int_{t\wedge\tau}^{\tau} e^{-\rho s} \big( 1 - (1-{\@fontswitch\relax\mathcal{R}}) (V^{\text{EZ},D}_s({\@fontswitch\relax\mathcal{R}}))^{\frac{-{\@fontswitch\relax\mathcal{R}}}{1-{\@fontswitch\relax\mathcal{R}}}}\big) dD_s \bigg] \nonumber\\ &\quad \le \mathbb{E}_t\bigg[\xi^0_{\tau} - (\xi^0_{\tau})^{1-{\@fontswitch\relax\mathcal{R}}} + \int_{t\wedge\tau}^{\tau} e^{-\rho s} {\@fontswitch\relax\mathcal{R}}V^{\text{EZ},D}_s({\@fontswitch\relax\mathcal{R}}) \, dD_s \bigg] \nonumber\\ &\quad \le {\@fontswitch\relax\mathcal{R}}\bigg( \mathbb{E}_t[(\xi^0_{\tau})^2] + \mathbb{E}_t\bigg[\sup_{s \ge t} (V_s^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}}))^2\bigg] \mathbb{E}_t\bigg[ \bigg(\int_0^{\tau} e^{-\rho s} dD_s \bigg)^2 \bigg] \bigg). \end{align}\] The first inequality uses the concavity of \(f(x) = 1 - (1-{\@fontswitch\relax\mathcal{R}})x^{\frac{-{\@fontswitch\relax\mathcal{R}}}{1-{\@fontswitch\relax\mathcal{R}}}}\), which implies \(f(x) \le f(1) + f'(1)(x-1) = {\@fontswitch\relax\mathcal{R}}+ {\@fontswitch\relax\mathcal{R}}(x-1)\). The second inequality follows from \(x - x^{1-{\@fontswitch\relax\mathcal{R}}} \le {\@fontswitch\relax\mathcal{R}}x^2\) for \(x > 0\) and \({\@fontswitch\relax\mathcal{R}}\in (0,1)\). By assumption, the RHS is finite \(\mathbb{P}\)-a.s., so there exists some \({\@fontswitch\relax\mathcal{F}}_t\)-measurable \(\overline{c}_t\) (not depending on \({\@fontswitch\relax\mathcal{R}}\)) such that \(K_t^D - V_t^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}}) \le {\@fontswitch\relax\mathcal{R}}\overline{c}_t,\) \(\mathbb{P}\)-a.s.. Furthermore, \(\overline{c}_t\) can be chosen so that \(\mathbb{E}[\sup_{t\ge0} \overline{c}_t] < \infty\). Indeed, it holds that \(\mathbb{E}[\sup_{t\ge0} |K^D_t - V^{\text{EZ},D}_t({\@fontswitch\relax\mathcal{R}})|] < \infty\) (by the bounds \(\mathbb{E}[\sup_{t\ge0} |K^D_{t \wedge \tau}|^2] < \infty\) and \(\mathbb{E}[\sup_{t\ge0} |V^{\text{EZ},D}_{t \wedge \tau}({\@fontswitch\relax\mathcal{R}})|^2] < \infty\), as well as \(K^D_t=V^{\text{EZ},D}_t({\@fontswitch\relax\mathcal{R}})=\xi_{\tau}^0\) on \(\{t\ge\tau\}\).)
Next, using \(V^{\text{EZ},D}_t = (V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}))^{1-{\@fontswitch\relax\mathcal{R}}}\), we obtain for any \(\lambda \in (0,1)\), \[\begin{align} &\lambda \ln(K_t^D) + (1-\lambda)(1-{\@fontswitch\relax\mathcal{R}}) \ln\big(V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}})\big)\\ &\quad\le \ln\Big( \lambda K^D_t + (1-\lambda) \big(V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}})\big)^{1-{\@fontswitch\relax\mathcal{R}}} \Big). \end{align}\] Applying Taylor’s expansion and taking the limit as \(\lambda \downarrow 0\), we deduce \[\ln(K_t^D) - \ln(V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}})) \le \frac{{\@fontswitch\relax\mathcal{R}}\overline{c}_t}{V_t^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}})} - {\@fontswitch\relax\mathcal{R}}\ln\big(V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}})\big),\quad \text{ \mathbb{P}-a.s.}\] Setting \(\overline{c}'_t = \overline{c}_t/C\) (with the lower bound \(C\) of \(V_t^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}})\)), we have ?? . ◻
Proof of Theorem 21.. We first prove the theorem under the additional assumption that \(\xi^0_{\tau} > C\) \(\mathbb{P}\)-a.s. with \(C > 0\) (as imposed in Lemma 43). The general case follows by considering the truncated sequence \(\tilde{\xi}_{\tau}^n = \frac{1}{n} \vee \xi^0_{\tau}\) for each \(n \in \mathbb{N}\), and then letting \(n \to \infty\). We present the proof in two steps.
Let \(0 < {\@fontswitch\relax\mathcal{R}}_1 < {\@fontswitch\relax\mathcal{R}}_2 < 1\). For any \(\theta \in \Theta\), \[\frac{1}{2{\@fontswitch\relax\mathcal{R}}_1} \int_0^t \theta_s^2 \, ds \ge \frac{1}{2{\@fontswitch\relax\mathcal{R}}_2} \int_0^t \theta_s^2 \, ds, \quad \forall t \ge 0.\] By the representation of \(V^{\theta,D}\) in ?? (see Proposition 8), the inequality \(V^{\theta,D}_t({\@fontswitch\relax\mathcal{R}}_1) \ge V^{\theta,D}_t({\@fontswitch\relax\mathcal{R}}_2),\) \(t\ge0,\) holds for all \(\theta\in\Theta\) (even if not in the admissible set \(\Theta^D({\@fontswitch\relax\mathcal{R}}_1)\) of \({\@fontswitch\relax\mathcal{R}}_1\)). Therefore, \[\begin{align} V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_1) =& \mathop{\mathrm{ess\,inf}}_{\theta\in\Theta^D({\@fontswitch\relax\mathcal{R}}_1)} V_t^{\theta,D}({\@fontswitch\relax\mathcal{R}}_1) \\ \ge& \mathop{\mathrm{ess\,inf}}_{\theta\in\Theta^D({\@fontswitch\relax\mathcal{R}}_2)} V_t^{\theta,D}({\@fontswitch\relax\mathcal{R}}_1) \ge \mathop{\mathrm{ess\,inf}}_{\theta\in\Theta^D({\@fontswitch\relax\mathcal{R}}_2)} V_t^{\theta,D}({\@fontswitch\relax\mathcal{R}}_2) = V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_2),~t\ge0. \end{align}\] The first inequality holds because \(\Theta^D({\@fontswitch\relax\mathcal{R}}_1) \subseteq \Theta^D({\@fontswitch\relax\mathcal{R}}_2)\); as \({\@fontswitch\relax\mathcal{R}}\) increases, the integrability conditions in 11 become less restrictive. On the other hand, note that \(\theta^0 := (\theta^0_t)_{t \ge 0} \equiv 0 \in \Theta^D({\@fontswitch\relax\mathcal{R}})\), for any \({\@fontswitch\relax\mathcal{R}}\in(0,1)\). Thus, \[V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}) = \inf_{\theta\in \Theta^D({\@fontswitch\relax\mathcal{R}})} V_t^{\theta,D}({\@fontswitch\relax\mathcal{R}}) \le V_t^{\theta^0,D}({\@fontswitch\relax\mathcal{R}}) \equiv K_t^D, \quad t\ge0,\;{\@fontswitch\relax\mathcal{R}}\in(0,1).\]
Let \(0 < {\@fontswitch\relax\mathcal{R}}_1 < {\@fontswitch\relax\mathcal{R}}_2 < 1\) and let \(\theta^{*,2} \in \Theta^D({\@fontswitch\relax\mathcal{R}}_2) \subset \Theta\) denote the worst-case kernel for \({\@fontswitch\relax\mathcal{R}}_2\) so that \(V^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_2) = V^{\theta^{*,2},D}({\@fontswitch\relax\mathcal{R}}_2)\). Specifically, \[\begin{align} \label{eq:theta422} \theta^{*,2}_t := - {\@fontswitch\relax\mathcal{R}}_2 \frac{{\mathbb{Z}}^D_t({\@fontswitch\relax\mathcal{R}}_2)}{V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_2)} = -\frac{{\@fontswitch\relax\mathcal{R}}_2}{1-{\@fontswitch\relax\mathcal{R}}_2} \frac{Z^D_t({\@fontswitch\relax\mathcal{R}}_2)}{V^{\text{EZ},D}_t({\@fontswitch\relax\mathcal{R}}_2)}, \quad t\ge0. \end{align}\tag{70}\] Let \(\eta^{\theta^{*,2}}\) be defined as in 7 with \(\theta=\theta^{*,2}\). By ?? in Lemma 43iii, \[\begin{align} V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_1) \le & \mathbb{E}_t^{\theta^{*,2}}\bigg[ \int_{t\wedge\tau}^{ \tau} e^{-\rho s} dD_s + \xi^0_{\tau}+ \frac{1}{2{\@fontswitch\relax\mathcal{R}}_1} \int_{t\wedge\tau}^{\tau} V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_1) (\theta^{*,2}_s)^2 ds\bigg]\\ = &V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_2) +\frac{1}{2}\Big(\frac{1}{{\@fontswitch\relax\mathcal{R}}_1}-\frac{1}{{\@fontswitch\relax\mathcal{R}}_2}\Big) \mathbb{E}_t^{\theta^{*,2}}\bigg[ \int_{t\wedge\tau}^{\tau} (\theta^{*,2}_s)^2 V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_1) ds \bigg]\\ &+ \mathbb{E}_t^{\theta^{*,2}}\bigg[ \int_{t\wedge\tau}^{\tau} \frac{(\theta^{*,2}_s)^2}{2{\@fontswitch\relax\mathcal{R}}_2} \big( V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_1)-V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_2)\big)ds \bigg]. \end{align}\] Since \(V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_1)\ge V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_2)\) by monotonicity, Grönwall’s inequality gives \[\begin{align} &V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_1)-V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_2) \\ &\quad\le \frac{{\@fontswitch\relax\mathcal{R}}_2-{\@fontswitch\relax\mathcal{R}}_1}{2{\@fontswitch\relax\mathcal{R}}_1{\@fontswitch\relax\mathcal{R}}_2}\mathbb{E}_t^{\theta^{*,2}}\bigg[\int_{t\wedge\tau}^{\tau} e^{\int_{t\wedge\tau}^{s}\frac{(\theta^{*,2}_u)^2}{2{\@fontswitch\relax\mathcal{R}}_2}du} (\theta^{*,2}_s)^2 V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_1) ds \bigg]\\ &\quad = \frac{{\@fontswitch\relax\mathcal{R}}_2-{\@fontswitch\relax\mathcal{R}}_1}{2{\@fontswitch\relax\mathcal{R}}_1{\@fontswitch\relax\mathcal{R}}_2}\mathbb{E}_t\bigg[\int_{t\wedge\tau}^{\tau} \eta_s^{\theta^{*,2}} e^{\int_{t\wedge\tau}^s\frac{(\theta^{*,2}_u)^2}{2{\@fontswitch\relax\mathcal{R}}_2}du} (\theta^{*,2}_s)^2 V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_1) ds \bigg]:=\mathrm{I}_t. \end{align}\] Using Itô’s formula for \(\ln V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_2)\), direct calculation gives \[\begin{align} \eta_s^{\theta^{*,2}} e^{\int_{t}^s\frac{(\theta^{*,2}_u)^2}{2{\@fontswitch\relax\mathcal{R}}_2}du} &= \exp\bigg\{{\@fontswitch\relax\mathcal{R}}_2\bigg(\ln\Big(\frac{V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_2)}{V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_2)}\Big)-\int_{t}^s\frac{e^{-\rho u}}{V^{\text{rob},D}_u({\@fontswitch\relax\mathcal{R}}_2)}dD_u \bigg)\bigg\}\\ &\le \Big(\frac{ V_{t}^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_2) }{ V_s^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_2)}\Big)^{{\@fontswitch\relax\mathcal{R}}_2}. \end{align}\] Hence, using the explicit form of \(\theta^{*,2}\) in 70 , we have \[\begin{align} \mathrm{I}_t\le& \frac{{\@fontswitch\relax\mathcal{R}}_2-{\@fontswitch\relax\mathcal{R}}_1}{2{\@fontswitch\relax\mathcal{R}}_1{\@fontswitch\relax\mathcal{R}}_2} \mathbb{E}_t\bigg[\int_{t\wedge\tau}^{\tau}\bigg( \frac{{\@fontswitch\relax\mathcal{R}}_2^2}{(1-{\@fontswitch\relax\mathcal{R}}_2)^2}\frac{|Z_s^D({\@fontswitch\relax\mathcal{R}}_2)|^2}{V_s^{\text{EZ},D}({\@fontswitch\relax\mathcal{R}}_2) \big(V^{\text{rob},D}_{s}({\@fontswitch\relax\mathcal{R}}_2)\big)^{1-{\@fontswitch\relax\mathcal{R}}_2}} \\ &\quad\quad\quad\quad\quad\quad\quad\quad\quad \times \Big(\frac{ V_{t}^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_2) }{ V_s^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_2)}\Big)^{{\@fontswitch\relax\mathcal{R}}_2} V^{\text{rob},D}_s({\@fontswitch\relax\mathcal{R}}_1)\bigg) ds \bigg]\\ \le&\frac{({\@fontswitch\relax\mathcal{R}}_2-{\@fontswitch\relax\mathcal{R}}_1){\@fontswitch\relax\mathcal{R}}_2}{2{\@fontswitch\relax\mathcal{R}}_1(1-{\@fontswitch\relax\mathcal{R}}_2)^2} \frac{\big(V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}_2)\big)^{{\@fontswitch\relax\mathcal{R}}_2}}{C^{\frac{1}{1-{\@fontswitch\relax\mathcal{R}}_2}}} \mathbb{E}_t\bigg[ \sup_{t\ge 0} \frac{ K^D_t}{ V^{\text{EZ},D}_t({\@fontswitch\relax\mathcal{R}}_2)} \int_{t\wedge\tau}^{\tau} |Z^D_s({\@fontswitch\relax\mathcal{R}}_2)|^2 ds \bigg], \end{align}\] where \(C^{\frac{1}{1-{\@fontswitch\relax\mathcal{R}}_2}}\) is a lower bound for \(V_s^{\text{rob}}({\@fontswitch\relax\mathcal{R}}_2)\), and we have used the fact that \(V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_1) \le K^D_t\). To bound the expectation, we use ?? , which yields \[\begin{align} \sup_{t\ge0} \frac{ K_t^D}{ V^{\text{EZ},D}_t({\@fontswitch\relax\mathcal{R}}_2)} \le \sup_{t\ge0} \Big(1+{\@fontswitch\relax\mathcal{R}}_2 \frac{\overline{c}_t}{C}\Big) =\sup_{t\ge0} (1+{\@fontswitch\relax\mathcal{R}}_2 \overline{c}_t')<\infty, \text{ {\mathbb{P}}-a.s.,} \end{align}\] where \(\overline{c}_t'\) is the constant from ?? , and we use \(\mathbb{E}[\sup_{t\ge0} \overline{c}_t'] < \infty\). Together with \(\int_0^{\tau} |Z_t^D({\@fontswitch\relax\mathcal{R}}_2)|^2 dt \in L^2({\@fontswitch\relax\mathcal{F}}_{\tau})\), there exists a constant \(C'_t>0\) such that \[\begin{align} 0 \le V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_1) - V^{\text{rob},D}_t({\@fontswitch\relax\mathcal{R}}_2) \le C'_t ({\@fontswitch\relax\mathcal{R}}_2 - {\@fontswitch\relax\mathcal{R}}_1), \text{ {\mathbb{P}}-a.s.} \end{align}\] This establishes the continuity of the mapping \((0,1) \ni {\@fontswitch\relax\mathcal{R}}\mapsto V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}})\). In addition, \(\lim_{{\@fontswitch\relax\mathcal{R}}\downarrow 0} V_t^{\text{rob},D}({\@fontswitch\relax\mathcal{R}}) = K^D_t\) follows from ?? , so ?? holds. ◻
Proof of Corollary 23.. Since \(J^*(x;{\@fontswitch\relax\mathcal{R}}) = \sup_{D\in {\@fontswitch\relax\mathcal{A}}^x} V^{\text{rob},D}_0({\@fontswitch\relax\mathcal{R}})\), the monotonicity and lower semi-continuity from left follows directly from Theorem 21: \[\liminf_{{\@fontswitch\relax\mathcal{R}}\uparrow \tilde{{\@fontswitch\relax\mathcal{R}}}} J^*(x;{\@fontswitch\relax\mathcal{R}}) = \liminf_{{\@fontswitch\relax\mathcal{R}}\uparrow \tilde{{\@fontswitch\relax\mathcal{R}}}} \sup_{D\in {\@fontswitch\relax\mathcal{A}}^x} V^{\text{rob},D}_0 ({\@fontswitch\relax\mathcal{R}}) \ge \sup_{D\in {\@fontswitch\relax\mathcal{A}}^x} V^{\text{rob},D}_0 (\tilde{{\@fontswitch\relax\mathcal{R}}}) = J^*(x;\tilde{{\@fontswitch\relax\mathcal{R}}}).\] For the right continuity, we observe that \(\lim_{{\@fontswitch\relax\mathcal{R}}\downarrow \tilde{{\@fontswitch\relax\mathcal{R}}}} J^*(x;{\@fontswitch\relax\mathcal{R}}) =\sup_{{\@fontswitch\relax\mathcal{R}}\le \tilde{{\@fontswitch\relax\mathcal{R}}} }J^*(x;{\@fontswitch\relax\mathcal{R}})\), and the interchange of supremum arguments yields the desired result. ◻
Proof of Theorem 24.. We denote by \(\underline{b}_{{\@fontswitch\relax\mathcal{R}}}\), \(\overline{b}_{{\@fontswitch\relax\mathcal{R}}}\), and \(\hat{b}_{{\@fontswitch\relax\mathcal{R}}}\) the constants from Assumption 15, emphasizing their dependence on \({\@fontswitch\relax\mathcal{R}}\). We set \(\hat{\boldsymbol{b}} := \max_{{\@fontswitch\relax\mathcal{R}}\in {\@fontswitch\relax\mathcal{I}}} \hat{b}_{{\@fontswitch\relax\mathcal{R}}}\). By Theorem 19, \(J^*(x;{\@fontswitch\relax\mathcal{R}})=v^*(x;{\@fontswitch\relax\mathcal{R}})\) for all \(x \in \mathbb{R}_+\), where \(v^*(x;{\@fontswitch\relax\mathcal{R}})\) is constructed in 43 (by rewriting \(\psi^+(b^*)\) as \(\psi^+(b^*_{{\@fontswitch\relax\mathcal{R}}};{\@fontswitch\relax\mathcal{R}})\) for each \({\@fontswitch\relax\mathcal{R}}\) therein; see also Assumption 15ii.). Throughout, let \((v^*)'(\cdot;{\@fontswitch\relax\mathcal{R}})\) denote the derivative of \(v^*(\cdot;{\@fontswitch\relax\mathcal{R}})\) w.r.t. \(x\). For clarity, we organize the proof into following steps.
Consider \({\@fontswitch\relax\mathcal{R}}_1<{\@fontswitch\relax\mathcal R}_2\in{\@fontswitch\relax\mathcal{I}}\).
Let \(D_1^*\) and \(\theta^{*,2}\) denote the optimal dividend for \({\@fontswitch\relax\mathcal{R}}_1\) and the worst-case kernel for \({\@fontswitch\relax\mathcal{R}}_2\) respectively. From the proof of Theorem 19 in Section 9, we have \[\begin{align} & v^*(x ; {\@fontswitch\relax\mathcal{R}}_1) \leq J^{\theta^{*, 2}}(x; D_1^*, v^*(\cdot; {\@fontswitch\relax\mathcal{R}}_1), {\@fontswitch\relax\mathcal{R}}_1),\\ &v^*(x,{\@fontswitch\relax\mathcal{R}}_2) \geqslant J^{\theta^{*, 2}}(x; D_1^*,v^*(\cdot; {\@fontswitch\relax\mathcal{R}}_2),{\@fontswitch\relax\mathcal{R}}_2), \end{align}\] where for any Borel measurable \(f:\mathbb{R}_+\to\mathbb{R}\), \[J^{\theta^{*, 2}}(x; D, f, {\@fontswitch\relax\mathcal{R}}):=\mathbb{E}^{\theta^{*, 2}}\bigg[\int_0^{\tau^{x,D}} e^{-\rho s}\Big(d D_s+\frac{({\theta_s^{*, 2}})^2}{2 {\@fontswitch\relax\mathcal{R}}}f(X_s^{D}) d s\Big)+e^{-\rho\tau^{x,D}} \xi_0\bigg].\] Following similar arguments as in Theorem 21 (Step 2), we obtain \[\begin{align} &0\le v^*(x;{\@fontswitch\relax\mathcal{R}}_1) - v^*(x;{\@fontswitch\relax\mathcal{R}}_2)\\ &\le\frac{({\@fontswitch\relax\mathcal{R}}_2-{\@fontswitch\relax\mathcal{R}}_1){\@fontswitch\relax\mathcal{R}}_2}{2{\@fontswitch\relax\mathcal{R}}_1}v^*(x;{\@fontswitch\relax\mathcal{R}}_2)^{{\@fontswitch\relax\mathcal{R}}_2}\mathbb{E}\bigg[\int_0^{\tau^{x,D_1^*}} e^{-\rho s}\tilde{v}_{{\@fontswitch\relax\mathcal{R}}_1,{\@fontswitch\relax\mathcal{R}}_2}(X_s^{D_1^*}) d s\bigg]\le C_{\@fontswitch\relax\mathcal{I}}({\@fontswitch\relax\mathcal{R}}_2-{\@fontswitch\relax\mathcal{R}}_1), \end{align}\] with \(\tilde{v}_{{\@fontswitch\relax\mathcal{R}}_1,{\@fontswitch\relax\mathcal{R}}_2}(\cdot):=((v^*)'(\cdot;{\@fontswitch\relax\mathcal{R}}_1)\sigma(\cdot))^2\frac{v^*(\cdot;{\@fontswitch\relax\mathcal{R}}_1)}{v^*(\cdot;{\@fontswitch\relax\mathcal{R}}_2)}(v^*(\cdot;{\@fontswitch\relax\mathcal{R}}_2))^{-(1-{\@fontswitch\relax\mathcal{R}}_2)}\) and some constant \(C_{\@fontswitch\relax\mathcal{I}}>0\) (not depending on \({\@fontswitch\relax\mathcal{R}}_1\) and \({\@fontswitch\relax\mathcal{R}}_2\)). This uniform bound holds since \(X^{D_{{\@fontswitch\relax\mathcal{R}}}^*}\) is valued in \([0, \hat{\boldsymbol{b}}]\) and \(\xi_0 + x \leq v^*(x; {\@fontswitch\relax\mathcal{R}}) \leq \xi_0 + x + \overline{\mu}/\rho\) for all \({\@fontswitch\relax\mathcal{R}}\in {\@fontswitch\relax\mathcal{I}}\). Therefore, \(v^*(x; {\@fontswitch\relax\mathcal{R}})\) is continuous in \({\@fontswitch\relax\mathcal{R}}\).
Note that \(v^*(0;{\@fontswitch\relax\mathcal{R}}) = \xi_0\), and by Theorem 21, the mapping \({\@fontswitch\relax\mathcal{R}}\mapsto v^*(x;{\@fontswitch\relax\mathcal{R}})\) is decreasing. Consequently, for any \({\@fontswitch\relax\mathcal{R}}_1 < {\@fontswitch\relax\mathcal{R}}_2\in{\@fontswitch\relax\mathcal{I}}\), we have \((v^*)'(0;{\@fontswitch\relax\mathcal{R}}_1) \geq (v^*)'(0;{\@fontswitch\relax\mathcal{R}}_2)\). Suppose, to the contrary, that \({\@fontswitch\relax\mathcal{R}}\mapsto (v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}})\) is not continuous. Then there exists \(\tilde{{\@fontswitch\relax\mathcal{R}}} \in \operatorname{int}{\@fontswitch\relax\mathcal{I}}\) such that \[\begin{align} \alpha_1 := \liminf_{\epsilon \rightarrow 0+} (v^*)^{\prime}(0;\tilde{{\@fontswitch\relax\mathcal{R}}} - \epsilon) > \limsup_{\epsilon \rightarrow 0+} (v^*)^{\prime}(0;\tilde{{\@fontswitch\relax\mathcal{R}}} + \epsilon) := \alpha_2 \ge 1, \end{align}\] where the inequality \(\alpha_2 \ge 1\) follows from the fact that \((v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}}) \ge 1\), as it satisfies the HJB-VI 23 . For a sufficiently small \(\delta_e>0\), we have \[\begin{align} (v^*)^{\prime}(0;\tilde{{\@fontswitch\relax\mathcal{R}}} - \epsilon) > (v^*)^{\prime}(0;\tilde{{\@fontswitch\relax\mathcal{R}}} + \epsilon) + \frac{\alpha_1 - \alpha_2}{2},\quad for all \epsilon\in(0,\delta_e). \end{align}\] Since \(v^*(x;{\@fontswitch\relax\mathcal{R}})\) is twice continuously differentiable w.r.t. \(x\) on \(\mathbb{R}_+\), the derivative \((v^*)'(x;{\@fontswitch\relax\mathcal{R}})\) is continuous in \(x\). Thus, for a sufficiently small \(\delta_x > 0\), \[\begin{align} (v^*)'(x;\tilde{{\@fontswitch\relax\mathcal{R}}} - \epsilon) > (v^*)'(x;\tilde{{\@fontswitch\relax\mathcal{R}}} + \epsilon) + \frac{\alpha_1 - \alpha_2}{2},\quad for all x \in [0, \delta_x], \epsilon \in (0, \delta_e). \end{align}\] Integrating both sides from \(x = 0\) to \(x = \delta_x\) yields \[\begin{align} v^*(\delta_x;\tilde{{\@fontswitch\relax\mathcal{R}}} - \epsilon) > v^*(\delta_x;\tilde{{\@fontswitch\relax\mathcal{R}}} + \epsilon) + \frac{\alpha_1 - \alpha_2}{2} \delta_x, \quad for all\epsilon \in (0, \delta_{e}). \end{align}\] Taking the limit as \(\epsilon \rightarrow 0^+\), we obtain \(v^*(\delta_x;\tilde{{\@fontswitch\relax\mathcal{R}}}^{-}) > v^*(\delta_x;\tilde{{\@fontswitch\relax\mathcal{R}}}^{+})\), which contradicts to the continuity of \({\@fontswitch\relax\mathcal{R}}\mapsto v^*(x;{\@fontswitch\relax\mathcal{R}})\) in Corollary 23. We conclude that \({\@fontswitch\relax\mathcal{I}} \ni {\@fontswitch\relax\mathcal{R}}\mapsto (v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}})\) must be continuous.
Recall \(g_{b^*_{{\@fontswitch\relax\mathcal{R}}}}(x) \equiv g_{b^*_{{\@fontswitch\relax\mathcal{R}}},0}(x)\) satisfying \[\begin{align} \left\{ \begin{aligned} &\frac{\sigma^2(x)}{2} (g_{b^*_{{\@fontswitch\relax\mathcal{R}}}})^{\prime}(x) + \mu(x) g_{b^*_{{\@fontswitch\relax\mathcal{R}}}}(x) + \frac{(1-{\@fontswitch\relax\mathcal{R}})}{2} \sigma^2(x) (g_{b^*_{{\@fontswitch\relax\mathcal{R}}}})^2(x) = \rho, \\ &g_{b^*_{{\@fontswitch\relax\mathcal{R}}}}(b^*_{{\@fontswitch\relax\mathcal{R}}}) = 1 / \psi^+(b^*_{{\@fontswitch\relax\mathcal{R}}}), \end{aligned} \right. \end{align}\] as in 42 . Fix \({\@fontswitch\relax\mathcal{R}}_1,{\@fontswitch\relax\mathcal{R}}_2\in{\@fontswitch\relax\mathcal{I}}\), for notational simplicity, let \(g_i\) denote \(g_{b^*_{{\@fontswitch\relax\mathcal{R}}_i}}\) for \(i=1,2\). Then, by 42 , \[\begin{align} g_i(x) &= g_i(0) - \int_0^{x} g'_i(y) \, dy = \frac{(v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}}_i)}{v^*(0;{\@fontswitch\relax\mathcal{R}}_i)} - \int_0^{x} g'_i(y) \, dy \\ &= \frac{(v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}}_i)}{\xi_0} + \int_0^{x} \left( \frac{2\mu(y)}{\sigma^2(y)} g_i(y) + (1-{\@fontswitch\relax\mathcal{R}}_1) g_i^2(y) - \rho \right) dy. \end{align}\] Therefore, for \(x \in [0, \hat{\boldsymbol{b}}]\), \[\begin{align} |g_1(x) - g_2(x)| \le & \frac{1}{\xi_0} \left| (v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}}_1) - (v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}}_2) \right| + C_{\hat{\boldsymbol{b}}} \int_{0}^{x} |g_1(y) - g_2(y)| dy \\ &+ \int_{0}^{x} ({\@fontswitch\relax\mathcal{R}}_2 - {\@fontswitch\relax\mathcal{R}}_1) g_2^2(y) dy, \end{align}\] where \(C_{\hat{\boldsymbol{b}}}:= \sup_{0 \le y \le \hat{\boldsymbol{b}}} \left\{ \frac{2\mu(y)}{\sigma^2(y)} + (1-{\@fontswitch\relax\mathcal{R}}_1) |g_1(y) + g_2(y)| \right\}\), for which we recall that \(\hat{\boldsymbol{b}}\) is a uniform upper bound for \(b^*_{{\@fontswitch\relax\mathcal{R}}}\) by Theorem 18. By Grönwall’s inequality, there is some constant \(c>0\) such that \[\begin{align} \sup_{0 \le x \le \hat{\boldsymbol{b}}} |g_1(x) - g_2(x)| \le c e^{C_{\hat{\boldsymbol{b}}}} \left( |(v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}}_1) - (v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}}_2)| + |{\@fontswitch\relax\mathcal{R}}_2 - {\@fontswitch\relax\mathcal{R}}_1| \right). \end{align}\] Since \({\@fontswitch\relax\mathcal{I}}\ni{\@fontswitch\relax\mathcal{R}}\mapsto (v^*)^{\prime}(0;{\@fontswitch\relax\mathcal{R}})\) is continuous (by Step 1), we have \[\begin{align} \label{eq:g95cts} \lim_{\substack{|{\@fontswitch\relax\mathcal{R}}_2 - {\@fontswitch\relax\mathcal{R}}_1| \to 0^+ \\ {\@fontswitch\relax\mathcal{R}}_1,\,{\@fontswitch\relax\mathcal{R}}_2 \in {\@fontswitch\relax\mathcal{I}}}} \sup_{0 \le x \le \hat{\boldsymbol{b}}} |g_1(x) - g_2(x)| = 0. \end{align}\tag{71}\] We thus show that \({\@fontswitch\relax\mathcal{I}}\ni {\@fontswitch\relax\mathcal{R}}\mapsto g_{b^*_{{\@fontswitch\relax\mathcal{R}}}}(x)\) is continuous, uniformly in \(x \in [0, \hat{\boldsymbol{b}}]\).
Recall the definition of \(b^*_{{\@fontswitch\relax\mathcal{R}}}\) in Proposition 37, and the relationship between \(v^*(x;{\@fontswitch\relax\mathcal{R}})\) and \(g_{b^*_{{\@fontswitch\relax\mathcal{R}}}}(x)\) as given in 41 and 43 . Without loss of generality, fix \({\@fontswitch\relax\mathcal{R}}_1<{\@fontswitch\relax\mathcal R}_2\in{\@fontswitch\relax\mathcal{I}}\). We have \[\begin{align} \label{eq:gR95relation} \frac{ \psi^+\big(b^*_{{\@fontswitch\relax\mathcal{R}}_1};{\@fontswitch\relax\mathcal{R}}_1\big) e^{-\int_0^{b^*_{{\@fontswitch\relax\mathcal{R}}_1}} g_1(y)dy} }{\psi^+\big(b^*_{{\@fontswitch\relax\mathcal{R}}_2};{\@fontswitch\relax\mathcal{R}}_2\big) e^{-\int_0^{b^*_{{\@fontswitch\relax\mathcal{R}}_2}} g_2(y)dy}} = \frac{v^*(0;{\@fontswitch\relax\mathcal{R}}_1)}{v^*(0;{\@fontswitch\relax\mathcal{R}}_2)} = 1. \end{align}\tag{72}\] To proceed, we recall several properties established in Section 4. By the definition of \(\psi^+\) in Assumption 15, the mapping \({\@fontswitch\relax\mathcal{R}}\mapsto \psi^+(x;{\@fontswitch\relax\mathcal{R}})\) is continuous and uniformly decreasing for \(x \geq 0\). Moreover, since \(\psi^+(x;{\@fontswitch\relax\mathcal{R}}) - x\) is decreasing for \(x > \underline{b}\), it follows that \((\psi^+)^{\prime}(x;{\@fontswitch\relax\mathcal{R}}) \leq 1\) for \(x > \underline{b}\). In addition, Lemma 40 i. asserts that \(g_i(x) \geq 1 / \psi^+(x;{\@fontswitch\relax\mathcal{R}}_i)\) for \(x > \underline{b}_{{\@fontswitch\relax\mathcal{R}}_i}\).
Taking logarithms on both sides of 72 , we consider the cases separately.
Case 1: \(b^*_{{\@fontswitch\relax\mathcal{R}}_1} < b^*_{{\@fontswitch\relax\mathcal{R}}_2}\). We have \[\begin{align} \ln \psi^+\big(b^*_{{\@fontswitch\relax\mathcal{R}}_1};{\@fontswitch\relax\mathcal{R}}_1\big) - \ln \psi^+\big(b^*_{{\@fontswitch\relax\mathcal{R}}_2};{\@fontswitch\relax\mathcal{R}}_1\big) + \int_{b^*_{{\@fontswitch\relax\mathcal{R}}_1}}^{b^*_{{\@fontswitch\relax\mathcal{R}}_2}} g_1 (y) \, dy &\le \int_0^{b^*_{{\@fontswitch\relax\mathcal{R}}_2}} \big(g_1(y)-g_2(y)\big) \, dy, \\ \int_{b^*_{{\@fontswitch\relax\mathcal{R}}_1}}^{b^*_{{\@fontswitch\relax\mathcal{R}}_2}} \left( g_1 (y) - \frac{(\psi^+)^{\prime}(y;{\@fontswitch\relax\mathcal{R}}_1)}{\psi^+(y;{\@fontswitch\relax\mathcal{R}}_1)} \right) \, dy &\le \int_0^{b^*_{{\@fontswitch\relax\mathcal{R}}_2}} |g_1(y)-g_2(y)| \, dy, \\ 0 \le \int_{b^*_{{\@fontswitch\relax\mathcal{R}}_1}}^{b^*_{{\@fontswitch\relax\mathcal{R}}_2}}\left( g_1 (y) -\frac{1}{\psi^+(y;{\@fontswitch\relax\mathcal{R}}_1)} \right) \, dy &\le \int_0^{b^*_{{\@fontswitch\relax\mathcal{R}}_2}} |g_1(y)-g_2(y)| \, dy, \end{align}\] where the first inequality follows from the monotonicity of \({\@fontswitch\relax\mathcal{R}}\mapsto \psi^+(x;{\@fontswitch\relax\mathcal{R}})\), and in the last line, we use \((\psi^+)^{\prime}(x;{\@fontswitch\relax\mathcal{R}}_1) \le 1\) and \(g_1(x) \ge 1/ \psi^+(x;{\@fontswitch\relax\mathcal{R}}_1)\) for \(x > b^*_{{\@fontswitch\relax\mathcal{R}}_1}>\underline{b}_{{\@fontswitch\relax\mathcal{R}}_1}\).
Case 2: \(b^*_{{\@fontswitch\relax\mathcal{R}}_1} > b^*_{{\@fontswitch\relax\mathcal{R}}_2}\). By analogous reasoning, \[\begin{align} 0 &< \int_{b^*_{{\@fontswitch\relax\mathcal{R}}_2}}^{b^*_{{\@fontswitch\relax\mathcal{R}}_1}} g_2 (y) - \frac{1}{\psi^+(y;{\@fontswitch\relax\mathcal{R}}_1)} \, dy \\ &\le \left| \ln \psi^+\big(b^*_{{\@fontswitch\relax\mathcal{R}}_2};{\@fontswitch\relax\mathcal{R}}_1\big) - \ln \psi^+\big(b^*_{{\@fontswitch\relax\mathcal{R}}_2};{\@fontswitch\relax\mathcal{R}}_2\big) \right| + \int_0^{b^*_{{\@fontswitch\relax\mathcal{R}}_1}} |g_2(y)-g_1(y)| \, dy, \end{align}\] where the strict inequality holds because \(\psi^+(x;{\@fontswitch\relax\mathcal{R}}_2) < \psi^+(x;{\@fontswitch\relax\mathcal{R}}_1)\), which implies \(g_2(x) \ge 1/\psi^+(x;{\@fontswitch\relax\mathcal{R}}_2) > 1/\psi^+(x;{\@fontswitch\relax\mathcal{R}}_1)\) for \(\underline{b}_{{\@fontswitch\relax\mathcal{R}}_2} < b^*_{{\@fontswitch\relax\mathcal{R}}_2} < x\).
For both cases, by 71 and the continuity of \({\@fontswitch\relax\mathcal{R}}\mapsto \psi^+(x;{\@fontswitch\relax\mathcal{R}})\), we can have the desired continuity of \(b^*_{\@fontswitch\relax\mathcal{R}}\). ◻