Publication Cover
Stochastics
An International Journal of Probability and Stochastic Processes
Volume 94, 2022 - Issue 4
895
Views
5
CrossRef citations to date
0
Altmetric
Research Article

Optimal stopping problems for maxima and minima in models with asymmetric information

ORCID Icon &
Pages 602-628 | Received 24 Nov 2020, Accepted 02 Sep 2021, Published online: 23 Sep 2021

Abstract

We derive closed-form solutions to optimal stopping problems related to the pricing of perpetual American withdrawable standard and lookback put and call options in an extension of the Black-Merton-Scholes model with asymmetric information. It is assumed that the contracts are withdrawn by their writers at the last hitting times for the underlying risky asset price of its running maximum or minimum over the infinite time interval which are not stopping times with respect to the observable filtration. We show that the optimal exercise times are the first times at which the asset price process reaches some lower or upper stochastic boundaries depending on the current values of its running maximum or minimum. The proof is based on the reduction of the original necessarily two-dimensional optimal stopping problems to the associated free-boundary problems and their solutions by means of the smooth-fit and normal-reflection conditions. We prove that the optimal exercise boundaries are the maximal and minimal solutions of some first-order nonlinear ordinary differential equations.

1. Introduction

Let us consider a probability space (Ω,F,P) with a standard Brownian motion B=(Bt)t0 and define the process X=(Xt)t0 by: (1) Xt=xexp((rδσ2/2)t+σBt)(1) which solves the stochastic differential equation: (2) dXt=(rδ)Xtdt+σXtdBt(X0=x)(2) where x>0 is fixed, and r>0, δ>0, and σ>0 are some given constants. Assume that the process X describes the price of a risky asset in a financial market, where r is the riskless interest rate, δ is the dividend rate paid to the asset holders, and σ is the volatility rate.

We aim to present closed-form solutions to the discounted optimal stopping problems with the values: (3) V¯i=supτE[er(τθ)Gi(Xτθ,Sτθ)]andU¯i=supζE[er(ζη)Fi(Xζη,Qζη)](3) with G1(x,s)=L1x, G2(x,s)=sL2x, G3(x,s)=sL3 and F1(x,q)=xK1, F2(x,q)=K2xq, F3(x,q)=K3q, for some deterministic constants Li,Ki>0, for i = 1, 2, 3. Suppose that the suprema in (Equation3) are taken over all stopping times τ and ζ with respect to the filtration (Ft)t0. The linear functions Gi(x,s) and Fi(x,q), for i = 1, 2, 3, represent the payoffs of standard and lookback options with floating and fixed strikes, respectively, which are widely used in financial practice. Here, the random times θ and η given by: (4) θ=sup{t0|Xt=St}andη=sup{t0|Xt=Qt}(4) are not stopping times with respect to the natural filtration (Ft)t0 of the process X, but they are honest times in the sense of Nikeghbali and Yor [Citation34]. The processes S=(St)t0 and Q=(Qt)t0 are the running maximum and minimum of X defined by: (5) St=s(max0utXu)andQt=q(min0utXu)(5) for some arbitrary 0<qxs.

Although the random times θ and η are not stopping times of X, we can still reduce the problems of (Equation3) to the associated optimal stopping problems for the two-dimensional (time-homogeneous strong) continuous Markov processes (X,S) and (X,Q). For this purpose, we observe that the expected rewards from (Equation3) admit the representations: (6) E[er(τθ)Gi(Xτθ,Sτθ)]=E[erτGi(Xτ,Sτ)I(τ<θ)+erθGi(Sθ,Sθ)I(θτ)](6) and (7) E[er(ζη)Fi(Xζη,Qζη)]=E[erζFi(Xζ,Qζ)I(ζ<η)+erηFi(Qη,Qη)I(ηζ)](7) for i = 1, 2, 3. Observe that the expressions in (Equation6) and (Equation7) allow to describe the original contracts as standard game (or Israeli) contingent claims introduced by Kifer [Citation27]. Such contracts enable their issuers to exercise their right to withdraw the contracts prematurely, by paying some penalties agreed in advance. Further developments of the Israeli options and the associated zero-sum optimal stopping (Dynkin) games were provided by Kyprianou [Citation29], Kühn and Kyprianou [Citation28], Kallsen and Kühn [Citation26], Baurdoux and Kyprianou [Citation3–5], Ekström and Villeneuve [Citation12], Ekström and Peskir [Citation11], Baurdoux, Kyprianou and Pardo [Citation6], Egami, Leung, and Yamazaki [Citation10], and Leung and Yamazaki [Citation31] among others. In contrast to the concept of game contingent claims mentioned above, in the present paper, we study the withdrawable American standard and lookback options in which the writers can terminate the contracts prematurely, by taking advantage of insider information which is not available to the holders. More precisely, we suppose that the option writers know the last hitting times for the underlying risky asset price process of its running maximum or minimum θ or η over the infinite time interval, respectively. For instance, such situations can occur when the writers issue options on the underlying risky assets which may represent the shares of either their own firms or other companies to the decisions of the board of which they have access and potentially influence them. In this view, the values V¯i and U¯i, for i = 1, 2, 3, in (Equation3) can be interpreted as the rational (or no-arbitrage) prices of the perpetual American withdrawable options in the appropriate extension of the Black-Merton-Scholes model (see, e.g. [Citation45, Chapter VII, Section 3g]). Some extensive overviews of the perpetual American options in diffusion models of financial markets and other related results in the area are provided in Shiryaev [Citation45, Chapter VIII; Section 2a], Peskir and Shiryaev [Citation41, Chapter VII; Section 25], and Detemple [Citation8] among others.

We further study the problems of (Equation3) as the associated optimal stopping problems of (Equation25) and (Equation26) for the two-dimensional continuous Markov processes having the underlying risky asset price X and its running maximum S or minimum Q as their state space components. The resulting problems turn out to be necessarily two-dimensional in the sense that they cannot be reduced to optimal stopping problems for one-dimensional Markov processes. Note that the integrals in the reward functionals of the optimal stopping problems in (Equation25) and (Equation26) contain complicated integrands depending on the asset price as well as its running maximum and minimum processes. This feature initiates further developments of techniques to determine the structure of the associated continuation and stopping regions as well as appropriate modifications of the normal-reflection conditions in the equivalent free-boundary problems. Moreover, we show that the appearance of the withdrawal opportunities for the writers of the contracts at the times θ and η may essentially change the behaviour of the optimal exercise boundaries for the holders of the options. The other problems of perpetual American cancellable or defaultable standard and lookback options in models with last passage times of constant and random levels for the underlying asset prices and zero or linear recoveries were recently considered in Gapeev, Li and Wu [Citation21] and Gapeev and Li [Citation16], respectively.

Discounted optimal stopping problems for the running maxima and minima of the initial continuous (diffusion-type) processes were initiated by Shepp and Shiryaev [Citation44] and further developed by Pedersen [Citation36], Guo and Shepp [Citation24], Gapeev [Citation14], Guo and Zervos [Citation25], Peskir [Citation39,Citation40], Glover, Hulley, and Peskir [Citation22], Gapeev and Rodosthenous [Citation17–19], Rodosthenous and Zervos [Citation43], Gapeev, Kort, and Lavrutich [Citation20], and Gapeev and Al Motairi [Citation15] among others. It was shown, by means of the maximality principle for solutions of optimal stopping stopping problems established by Peskir [Citation37], which is equivalent to the superharmonic characterization of the value functions, that the optimal stopping boundaries are given by the appropriate extremal solutions of certain (systems of) first-order nonlinear ordinary differential equations. More complicated optimal stopping problems in models with spectrally negative Lévy processes and their running maxima were studied by Asmussen, Avram, and Pistorius [Citation1], Avram, Kyprianou, and Pistorius [Citation2], Ott [Citation35], Kyprianou and Ott [Citation30], and Li, Vu and Zhou [Citation32] among others.

The rest of the paper is organized as follows. In Section 2, we embed the original problems of (Equation3) into the optimal stopping problems of (Equation25) and (Equation26) for the two-dimensional continuous Markov processes (X,S) and (X,Q) defined in (Equation1) and (Equation5). It is shown that the optimal stopping times τi and ζi are the first times at which the process X reaches some lower or upper boundaries ai(S) or bi(Q) depending on the current values of the processes S or Q, for i = 1, 2, 3, respectively. In Section 3, we derive closed-form expressions for the associated value functions Vi(x,s) and Ui(x,q) as solutions to the equivalent free-boundary problems and apply the modified normal-reflection conditions at the edges of the two-dimensional state spaces for (X,S) or (X,Q) to characterize the optimal stopping boundaries ai(S) and bi(Q), for i = 1, 2, 3, as the maximal or minimal solutions to the resulting first-order nonlinear ordinary differential equations. In Section 4, by using the change-of-variable formula with local time on surfaces from Peskir [Citation38], we verify that the solutions of the free-boundary problems provide the solutions of the original optimal stopping problems. The main results of the paper are stated in Theorems 2.1 and 4.1.

2. Preliminaries

In this section, we introduce the setting and notation of the two-dimensional optimal stopping problems which are related to the pricing of perpetual American withdrawable standard and lookback put and call options and formulate the equivalent free-boundary problems.

2.1. The optimal stopping problems

In order to compute the expectations in (Equation6) and (Equation7), let us now introduce the conditional survival processes Z=(Zt)t0 and Y=(Yt)t0 defined by Zt=P(θ>t|Ft) and Yt=P(η>t|Ft), for all t0, respectively. Note that the processes Z and Y are called the Azéma supermartingales of the random times θ and η (see, e.g. [Citation33, Section 1.2.1]). By using the fact that the global maximum of a Brownian motion with the drift coefficient (rδ)/σ21/2<0 has an exponential distribution with the mean 1/(2(δr)/σ2+1), while the negative of the global minimum of a Brownian motion with the drift coefficient (rδ)/σ21/2>0 has an exponential distribution with the mean 1/(2(rδ)/σ21) (see, e.g. [Citation42, Chapter II, Exercise 3.12]), we have: (8) Zt={(St/Xt)α,ifα<01,ifα0andYt={(Qt/Xt)α,ifα>01,ifα0(8) for all t0, under s = x and q = x, where we set α=2(rδ)/σ21, respectively. The representations in (Equation8) can also be obtained from applying Doob's maximal equality (see [Citation34, Lemma 2.1 and Proposition 2.2]) to the process Xα=(Xtα)t0, which is a strictly positive continuous local martingale converging to zero at infinity. Then, it follows from a direct application of the tower property for conditional expectations that the first terms in the right-hand sides of the expressions in (Equation6) and (Equation7) have the form: (9) E[erτGi(Xτ,Sτ)I(τ<θ)]=E[erτGi(Xτ,Sτ)(Sτ/Xτ)α](9) when α<0, under s = x, and (10) E[erζFi(Xζ,Qζ)I(ζ<η)]=E[erζFi(Xζ,Qζ)(Qζ/Xζ)α](10) when α>0, under q = x, for any stopping times τ and ζ of the process X, respectively. Moreover, it follows from standard applications of Itô's formula (see, e.g. [Citation42, Chapter IV, Theorem 3.3]) and the properties that the processes S and Q may change their values only when Xt=St and Xt=Qt, for t0, respectively, that the Azéma supermartingales Z and Y from (Equation8) admit the stochastic differentials: (11) dZt=α(StXt)ασdBt+αI(Xt=St)(StXt)αdStSt=α(StXt)ασdBt+αdStSt(11) when α<0, and (12) dYt=α(QtXt)ασdBt+αI(Xt=Qt)(QtXt)αdQtQt=α(QtXt)ασdBt+αdQtQt(12) when α>0, respectively. Hence, it follows from Doob-Meyer decompositions for the processes Z and Y in (Equation11) and (Equation12) and applications of the dual predictable projection property (see, e.g. [Citation34, Corollary 2.4]) that the second terms in the right-hand sides of the expressions in (Equation6) and (Equation7) admit the representations: (13) E[erθGi(Sθ,Sθ)I(θτ)]=E[0τeruGi(Su,Su)α(SuXu)αdSuSu](13) when α<0, under s = x, and (14) E[erηFi(Qη,Qη)I(ηζ)]=E[0ζeruFi(Qu,Qu)α(QuXu)αdQuQu](14) when α>0, under q = x, for any stopping times τ and ζ, and every i=1,2,3, respectively.

Furthermore, by means of standard applications of Itô's formula, taking into account the facts that xxGi(x,s)=0 and xxFi(x,q)=0, we obtain that the processes ertGi(Xt,St)(St/Xt)α and ertFi(Xt,Qt)(Qt/Xt)α admit the representations: (15) ertGi(Xt,St)(St/Xt)α=Gi(x,s)(s/x)α+0teru(xGi(Xu,Su)(rδ)XurGi(Xu,Su))I(XuSu)(SuXu)αdu+0teru(sGi(Xu,Su)Su+αGi(Xu,Su))I(Xu=Su)(SuXu)αdSuSu+Nti,1(15) when α<0, for each 0<xs, and (16) ertFi(Xt,Qt)(Qt/Xt)α=Fi(x,q)(q/x)α+0teru(xFi(Xu,Qu)(rδ)XurFi(Xu,Qu))I(XuQu)(QuXu)αdu+0teru(qFi(Xu,Qu)Qu+αFi(Xu,Qu))I(Xu=Qu)(QuXu)αdQuQu+Nti,2(16) when α>0, for each 0<qx, for every i = 1, 2, 3, and all t0, where we set δ=δ+ασ22rδσ2, that can be considered as a withdrawal adjusted dividend rate. Here, the processes Ni,j=(Nti,j)t0, for i = 1, 2, 3 and j = 1, 2, defined by: (17) Nti,1=0teru(xGi(Xu,Su)XuαGi(Xu,Su))I(XuSu)(SuXu)ασdBu(17) and (18) Nti,2=0teru(xFi(Xu,Qu)XuαFi(Xu,Qu))I(XuQu)(QuXu)ασdBu(18) are continuous uniformly integrable martingales under the probability measure P, when α<0 and α>0, respectively. Note that the processes S and Q may change their values only at the times when Xt=St and Xt=Qt, for t0, respectively, and such times accumulated over the infinite horizon form the sets of the Lebesgue measure zero, so that the indicators in the expressions of (Equation15)–(Equation16) and (Equation17)–(Equation18) can be ignored (see also Proof of Theorem 4.1 below for more explanations and references). Then, inserting τ and ζ in place of t into (Equation15) and (Equation16), respectively, by means of Doob's optional sampling theorem (see, e.g. [Citation42, Chapter II, Theorem 3.2]), we get: (19) E[erτGi(Xτ,Sτ)(Sτ/Xτ)α]=Gi(x,s)(s/x)α+E[0τeru(xGi(Xu,Su)(rδ)XurGi(Xu,Su))(SuXu)αdu+0τeru(sGi(Su,Su)Su+αGi(Su,Su))dSuSu](19) when α<0, and (20) E[erζFi(Xζ,Qζ)(Qζ/Xζ)α]=Fi(x,q)(q/x)α+E[0ζeru(xFi(Xu,Qu)(rδ)XurFi(Xu,Qu))(QuXu)αdu+0τeru(qFi(Qu,Qu)Qu+αFi(Qu,Qu))dQuQu](20) when α>0, for any stopping times τ and ζ, and every i = 1, 2, 3. Hence, getting the expressions in (Equation19) and (Equation20) together with the ones in (Equation13) and (Equation14) above and combining them with the expressions in (Equation6) and (Equation7), we may conclude that the values of (Equation3) are given by: (21) V¯i=Gi(x,x)+supτE[0τeruHi,1(Xu,Su)du+0τerusGi(Su,Su)dSu](21) when α<0, under s = x, and (22) U¯i=Fi(x,x)+supζE[0ζeruHi,2(Xu,Qu)du+0ζeruqFi(Xu,Qu)dQu](22) when α<0, under q = x, and i = 1, 2, 3, where the suprema are taken over all stopping times τ and ζ of the processes (X,S) and (X,Q), respectively. Here, we set: (23) Hi,1(x,s)=(xGi(x,s)(rδ)xrGi(x,s))(s/x)α(23) for all 0<xs, and (24) Hi,2(x,q)=(xFi(x,q)(rδ)xrFi(x,q))(q/x)α(24) for all 0<qx, respectively. In this case, we see that the problems in (Equation21) and (Equation22) can naturally be embedded into the optimal stopping problems for the (time-homogeneous strong) Markov processes (X,S)=(Xt,St)t0 and (X,Q)=(Xt,Qt)t0 with the value functions: (25) Vi(x,s)=supτEx,s[0τeruHi,1(Xu,Su)du+0τerusGi(Su,Su)dSu](25) when α<0, and (26) Ui(x,q)=supζEx,q[0ζeruHi,2(Xu,Qu)du+0ζeruqFi(Qu,Qu)dQu](26) when α>0, for every i = 1, 2, 3, respectively. Here, Ex,s and Ex,q denote the expectations with respect to the probability measures Px,s and Px,q under which the two-dimensional Markov processes (X,S) and (X,Q) defined in (Equation1) and (Equation5) start at (x,s)E1={(x,s)R2|0<xs} and (x,q)E2={(x,q)R2|0<qx}, respectively. We further obtain solutions to the optimal stopping problems in (Equation25) and (Equation26) and verify below that the value functions Vi(x,s) and Ui(x,q), for i = 1, 2, 3, are the solutions of the problems in (Equation21) and (Equation22), and thus, give the solutions of the original problems in (Equation3), under s = x and q = x, respectively.

2.2. The structure of optimal exercise times

Let us now determine the structure of the optimal stopping times at which the holders should exercise the contracts. For this purpose, we formulate the following assertion.

Theorem 2.1

Let the processes (X,S) and (X,Q) be given by (Equation1) and (Equation5), with some r>0, δ>0, and σ>0 fixed, and the inequality δ2rδσ2>0 be satisfied. Suppose that the random times θ and η are defined in (Equation4). Then, the optimal exercise times for the perpetual American withdrawable standard and lookback put and call options with the values in (Equation25) and (Equation26) have the structure: (27) τi=inf{t0|Xtai(St)}andζi=inf{t0|Xtbi(Qt)}(27) under α<0 and α>0, for i = 1, 2, 3, respectively.

The optimal exercise boundaries ai(s) and bi(q) in (Equation27) represent some functions satisfying the inequalities a_i(s)<ai(s)<a¯i(s)s, for s>s_i, and b_i(q)q<bi(q)<b¯i(q), for 0<q<q¯i, as well as the equalities a1(s)=s, a3(s)=0, for all ss_i, and b1(q)=q, b3(q)=, for all qq¯i, for every i = 1, 2, 3. Here, we have a_1(s)=rL1α/(δ(α1)) and a¯1(s)=rL1/δ, a_2(s)=rsα/(δL2(α1)) and a¯2(s)=rs/(δL2), while a_i(s)=0 and a¯3(s)=s, under α<0, with some 0s_1a¯1 as well as s_2=0 and s_3=L3.

We also have b_1(q)=rK1/δ and b¯1(q)=rK1α/(δ(α1)), b_2(q)=rq/(δK2) and b¯2(q)=rqα/(δK2(α1)), under α>1, as well as b¯1(q)= and b¯2(q)=, under 0<α1, while b_3(q)=q and b¯3(q)=, with some q¯1b_1 as well as q¯2= and q¯3=K3. Moreover, the boundary a3(s) is increasing on (L3,L3α/(α+1)), under α<1, and on (L3,), under 1α<0, while the boundary b3(q) is increasing on (K3α/(α+1),K3), under α>0.

Proof.

(i) We first note that, by virtue of properties of the running maximum S and minimum Q from (Equation5) of the geometric Brownian motion X from (Equation1) (see, e.g. [Citation9, Subsection 3.3] for similar arguments applied to the running maxima of the Bessel processes), it is seen that, for any s>0 and q>0 fixed and an infinitesimally small deterministic time interval Δ, we have: (28) SΔ=smax0uΔXu=s(s+ΔX)+o(Δ)asΔ0(28) and (29) QΔ=qmin0uΔXu=q(q+ΔX)+o(Δ)asΔ0(29) where we set ΔX=XΔs and ΔX=XΔq, respectively. Observe that ΔS=o(Δ) when ΔX0, ΔS=ΔX+o(Δ) when ΔX>0, ΔQ=o(Δ) when ΔX0, and ΔQ=ΔX+o(Δ) when ΔX<0, where we set ΔS=SΔs and ΔQ=QΔq, and recall that o(Δ) denotes a random function satisfying o(Δ)/Δ0 as Δ0 (P-a.s.). In this case, using the asymptotic formulas: (30) Es,s[ΔX;ΔX>0]Es,s[ΔXI(ΔX>0)]sσΔ2πasΔ0(30) and (31) Eq,q[ΔX;ΔX<0]Eq,q[ΔXI(ΔX<0)]qσΔ2πasΔ0(31) as well as taking into account the structure of the rewards in (Equation25) and (Equation26), we get: (32) Es,s[erΔHi,1(s,s)Δ+erΔsGi(s,s)ΔS]erΔHi,1(s,s)Δ+erΔsGi(s,s)sσΔ2πasΔ0(32) and (33) Eq,q[erΔHi,2(q,q)Δ+erΔqFi(q,q)ΔQ]erΔHi,2(q,q)ΔerΔqFi(q,q)qσΔ2πasΔ0(33) for each s>0 and q>0 fixed. Since we have sG1(x,s)=0 and sGi(x,s)=1, for i = 2, 3, as well as qF1(x,q)=0 and qFi(x,q)=1, for i = 2, 3, we see that the resulting coefficients by the terms of order Δ in the expressions of (Equation32) and (Equation33) are strictly positive, for all (x,s)E1 as well as (x,q)E2 and every i = 2, 3. Hence, taking into account the facts that the process S is positive and increasing and the process Q is positive and decreasing, we may therefore conclude from the structure of the second integrands in (Equation25) and (Equation26) as well as the heuristic arguments presented in (Equation32) and (Equation33) above that it is not optimal to exercise the withdrawable lookback put options when Xt=St, while it is not optimal to exercise the withdrawable call options when Xt=Qt, for any t0, respectively. In other words, these facts mean that the diagonal d1={(x,s)E1|x=s} belongs to the continuation region Ci,1, for i = 2, 3, which has the form: (34) Ci,1={(x,s)E1|Vi(x,s)>0}(34) while the diagonal d2={(x,q)E2|x=q} belongs to the continuation region Ci,2, for i = 2, 3, which is given by: (35) Ci,2={(x,q)E2|Ui(x,q)>0}(35) for every i = 1, 2, 3 (see, e.g. [Citation41, Chapter I, Subsection 2.2]).

Moreover, it follows from the structure of the first integrands in (Equation25) and (Equation26) that it is not optimal to exercise the perpetual American withdrawable standard or lookback put option when Hi,1(Xt,St)0, while it is not optimal to exercise the appropriate standard or lookback call option when Hi,2(Xt,Qt)0, for any t0, for every i = 1, 2, 3, respectively. In other words, these facts mean that the set {(x,s)E1|Hi,1(x,s)0} belongs to the continuation region Ci,1 in (Equation34), while the set {(x,q)E2|Hi,2(x,q)0} belongs to the continuation region Ci,2 in (Equation35), for every i = 1, 2, 3, respectively. In this respect, if we assume that δ2rδσ20 holds, that obviously implies that α2(rδ)/σ21<0 holds, then we see from the expression in (Equation25) that the equality τ1=0 should hold for the optimal stopping time, so that one should exercise the appropriate perpetual American withdrawable put option instantly. In this view, for simplicity of presentation, we further assume that δ>0 holds, as well as note that the fact that α2(rδ)/σ21>0 holds obviously implies that δ2rδσ2>0 holds. In this case, the inequality H1,1(x,s)=(δxrL1)(s/x)α0 is satisfied if and only if a¯1xs holds with a¯1=rL1/δ, the inequality H2,1(x,s)=(δL2xrs)(s/x)α0 is satisfied if and only if a¯2(s)xs holds with a¯2(s)=rs/(δL2), while the inequality H3,1(x,s)=r(L3s)(s/x)α0 is satisfied if and only if 0<xsL3 holds. Furthermore, the inequality H1,2(x,q)=(rK1δx)(q/x)α0 is satisfied if and only if qxb_1 holds with b_1=rK1/δ, the inequality H2,2(x,q)=(rqδK2x)(q/x)α0 is satisfied if and only if qxb_2(q) holds with b_2(q)=rq/(δK2), while the inequality H3,2(x,q)=r(qK3)(q/x)α0 is satisfied if and only if xqK3 holds (see Figures 1-4 below for the computer drawings of the boundary estimates a¯1, b_1 and a¯2(s), b_2(q)).

(ii) Let us now describe the structure of the continuation regions in (Equation34) and (Equation35). For this purpose, we provide an analysis of the reward functionals of the optimal stopping problems from (Equation25) and (Equation26). On the one hand, we observe that the function H1,1(x,s)=(δxrL1)(s/x)α decreases in x on the interval (0,a_1), and then, it increases in x on the interval (a_1,s) with a_1=rL1α/(δ(α1))<rL1/δ=a¯1, under α<0, for each s>s_1 fixed and some 0s_1a¯1. In this case, the function H1,1(x,s) attains its global minimum at x=a_1, for any s>s_1. According to the comparison results for strong solutions of (one-dimensional) stochastic differential equations (see, e.g. [Citation13, Theorem 1]), this fact means that the process (H1,1(Xt,St))t0 started at the point H1,1(a_1,s) has the smallest sample paths than the one started at any other point H1,1(x,s), for any 0<x<s such that xa_1 and s>s_1. In this respect, we may conclude that the point (a_1,s) belongs to the stopping region Di,1, for i = 1, which has the form: (36) Di,1={(x,s)E1|Vi(x,s)=0}(36) for every i = 1, 2, 3 (see, e.g. [Citation41, Chapter I, Subsection 2.2]), since otherwise, all the points (x,s) such that 0<x<s, for any s>s_1, would belong to the continuation region C1,1 from (Equation34) too. The latter fact contradicts the obvious property that it is better to stop the process (X,S) at time zero than do not stop the process at all during the infinite time interval, under the assumption that α<0. Therefore, taking into account the fact that the function H1,1(x,s) is negative on the interval (0,a_1), we see that all the points (x,s) such that 0<xa_1s, for any s>s_1, belong to the stopping region D1,1 from (Equation36) as well.

Note that similar arguments applied for the function H2,1(x,s)=(δL2xrs)(s/x)α show that all the points (x,s) such that 0<xa_2(s)s, with a_2(s)=rsα/(δL2(α1))<rs/(δL2)=a¯2(s), under α<0, for each s>s_2=0 fixed, belong to the stopping region D2,1 from (Equation36). Moreover, it follows from the property that the function H3,1(x,s)=r(L3s)(s/x)α increases in x on the interval (0,s), under α<0, that, for each s>s_3=L3 fixed, there exists a sufficiently small x>0 such that the point (x,s) belongs to the stopping region D3,1 from (Equation36). According to arguments similar to the ones applied in [Citation9, Subsection 3.3] and [Citation37, Subsection 3.3], the latter properties can be explained by the fact that the costs of waiting until the process X comes from such a small x>0 to the current value of the maximum S may be too high, due to the presence of the discounting factor in the reward functional of (Equation25), one should stop at this x>0 immediately.

On the other hand, we observe that the function H1,2(x,q)=(rK1δx)(q/x)α decreases in x on the interval (q,b¯1), and then, it increases in x on the interval (b¯1,) with b¯1=rK1α/(δ(α1))>rK1/δ=b_1, under α>1, for each q<q¯1 fixed and some q¯1b_1. In this case, the function H1,2(x,q) attains its global minimum at x=b¯1, for any q<q¯1. According to the comparison results for strong solutions of (one-dimensional) stochastic differential equations, this fact means that the process (H1,2(Xt,Qt))t0 started at the point H1,2(b¯1,q) has the smallest sample paths than the one started at any other point H1,2(x,q), for any x>q such that xb¯1 and q<q¯1. In this respect, we may conclude that the point (b¯1,q) belongs to the stopping region Di,2, for i = 1, which has the form: (37) Di,2={(x,q)E2|Ui(x,q)=0}(37) for i = 1, 2, 3, respectively, since otherwise, all the points (x,q) such that x>q, for any q<q¯1, would belong to the continuation region C1,2 from (Equation35) too. The latter fact contradicts the obvious property that it is better to stop the process (X,Q) at time zero than do not stop the process at all during the infinite time interval, under the assumption that α>1. Therefore, taking into account the fact that the function H1,2(x,q) is negative on the interval (b¯1,), we see that all the points (x,q) such that xb¯1q, for any q<q¯1, belong to the stopping region D1,2 from (Equation37) as well.

Note that similar arguments applied for the function H2,2(x,q)=(rqδK2x)(q/x)α show that all the points (x,q) such that xb¯2(q)q, with b¯2(q)=rqα/(δK2(α1))>rq/(δK2)=b_2(q), under α>1, for each q<q¯2= fixed, belong to the stopping region D2,2 from (Equation37). Moreover, it follows from the fact that the function H3,2(x,q)=r(qK3)(q/x)α is strictly increasing in x on the interval (q,), under α>0, that, for each q<q¯3=K3 fixed, there exists a sufficiently large x>0 such that the point (x,q) belongs to the stopping region D3,2 from (Equation37). The same arguments based on the strict increase of the functions Hi,2(x,q), for i = 1, 2, in x on the interval (q,), under 0<α1, for each q<q¯i fixed, for i = 1, 2, with some q¯1>b_1 and q¯2=, show that, there exists a sufficiently large x>0 such that the point (x,q) belongs to the stopping regions Di,2, for i = 1, 2, from (Equation37). The latter properties can be explained by the fact that the costs of waiting until the process X comes from such a large x>0 to the current value of the minimum Q may be too high, due to the presence of the discounting factor in the reward functional of (Equation26), one should stop at this x>0 immediately. In this view, we can set b¯1= and b¯2(q)=, for q>0, under 0<α1 (see Figures  below for the computer drawings of the boundary estimates a_1, b¯1 and a_2(s), b¯2(q)).

Figure 1. A computer drawing of the continuation and stopping regions C1,1 and D1,1 formed by the optimal exercise boundary a1(s) and its estimates a_1 and a¯1.

Figure 1. A computer drawing of the continuation and stopping regions C1,1∗ and D1,1∗ formed by the optimal exercise boundary a1∗(s) and its estimates a_1 and a¯1.

Figure 2. A computer drawing of the continuation and stopping regions C1,2 and D1,2 formed by the optimal exercise boundary b1(s) and its estimates b_1 and b¯1.

Figure 2. A computer drawing of the continuation and stopping regions C1,2∗ and D1,2∗ formed by the optimal exercise boundary b1∗(s) and its estimates b_1 and b¯1.

Figure 3. A computer drawing of the continuation and stopping regions C2,1 and D2,1 formed by the optimal exercise boundary a2(s) and its estimates a_2(s) and a¯2(s).

Figure 3. A computer drawing of the continuation and stopping regions C2,1∗ and D2,1∗ formed by the optimal exercise boundary a2∗(s) and its estimates a_2(s) and a¯2(s).

Figure 4. A computer drawing of the continuation and stopping regions C2,2 and D2,2 formed by the optimal exercise boundary b2(q) and its estimates b_2(q) and b¯2(q).

Figure 4. A computer drawing of the continuation and stopping regions C2,2∗ and D2,2∗ formed by the optimal exercise boundary b2∗(q) and its estimates b_2(q) and b¯2(q).

It is seen from the results of Theorem 4.1 proved below that the value functions Vi(x,s) and Ui(x,q) are continuous, so that the sets Ci,1 and Ci,2 in (Equation34) and (Equation35) are open, while the sets Di,1 and Di,2 in (Equation36) and (Equation37) are closed, for every i = 1, 2, 3 (see Figures for the computer drawings of the continuation and stopping regions Ci,j and Di,j, for i = 1, 2, 3 and j = 1, 2).

Figure 6. A computer drawing of the continuation and stopping regions C3,2 and D3,2 formed by the optimal exercise boundary b3(q) and the points K3 and K3α/(α+1).

Figure 6. A computer drawing of the continuation and stopping regions C3,2∗ and D3,2∗ formed by the optimal exercise boundary b3∗(q) and the points K3 and K3α/(α+1).

(iii) Now, we observe that, if we take some (x,s)Di,1 from (Equation36) such that x>a_i(s) with a_i(s) specified above and use the fact that the process (X,S) started at some (x,s) such that a_i(s)x<x passes through the point (x,s) before hitting the diagonal d1={(x,s)E1|x=s}, then the equality in (Equation25) implies that Vi(x,s)Vi(x,s)=0 holds, so that (x,s)Di,1, for i = 1, 2, 3. Moreover, if we take some (x,q)Di,2 from (Equation37) such that x<b¯i(q) with b¯i(q) specified above and use the fact that the process (X,Q) started at some (x,q) such that b¯i(q)x>x passes through the point (x,q) before hitting the diagonal d2={(x,q)E2|x=q}, then the equality in (Equation26) implies that Ui(x,q)Ui(x,q)=0 holds, so that (x,q)Di,2, for i = 1, 2, 3.

On the other hand, if take some (x,s)Ci,1 from (Equation34) and use the fact that the process (X,S) started at (x,s) passes through some point (x,s) such that x>x before hitting the diagonal d1, then the equality in (Equation25) yields that Vi(x,s)Vi(x,s)>0 holds, so that (x,s)Ci,1, for i = 1, 2, 3. Moreover, if we take some (x,q)Ci,2 from (Equation35) and use the fact that the process (X,Q) started at (x,q) passes through some point (x,q) such that x<x before hitting the diagonal d2, then the equality in (Equation26) yields that Ui(x,q)Ui(x,q)>0 holds, so that (x,q)Ci,2.

Hence, we may conclude that there exist functions ai(s) and bi(q) satisfying the inequalities ai(s)<a¯i(s)s, for all s>s_i, and bi(q)>b_i(q)q, for all q<q¯i, as well as the equalities a1(s)=s, a3(s)=0, for all ss_i, and b1(q)=q, b3(q)=, for all qq¯i, such that the continuation regions Ci,j, for j=1,2, in (Equation34) and (Equation35) have the form: (38) Ci,1={(x,s)E1|ai(s)<xs}andCi,2={(x,q)E2|qx<bi(q)}(38) while the stopping regions Di,j, for j = 1, 2, in (Equation36) and (Equation37) are given by: (39) Di,1={(x,s)E1|xai(s)}andDi,2={(x,q)E2|xbi(q)}(39) for every i = 1, 2, 3, respectively (see Figures for the computer drawings of the optimal stopping boundaries ai(s) and bi(q), for i = 1, 2, 3).

(iv) We finally specify the behaviour of the optimal exercise boundaries a3(s) and b3(q). For the ease of presentation, in the rest of this section, we indicate by (X(x),S(s,x)) and (X(x),Q(q,x)) the dependence of the processes (X,S) and (X,Q) defined in (Equation1) and (Equation5) from their starting points (x,s)E1 and (x,q)E2. Let us first fix some (x,s)C3,1 such that L3=s_3<s<L3α/(α+1), under α<1, so that V3(x,s)>0 holds. Then, consider the optimal stopping time τ3=τ3(x,s) for the problem (Equation25), for i = 3, for this starting point (x,s) of the process (X,S) from (Equation1) and (Equation5). Then, using the property that the function H3,1(x,s)=r(L3s)(s/x)α is decreasing in s on (L3,L3α/(α+1)), under α<1, for each x>0 fixed, and the fact that sG3(s,s)=1, by virtue of the structure of the running maximum S(s,x) of the process X(x), for any other starting point (x,s)E1 such that L3<s<s<L3α/(α+1), we have: (40) V3(x,s)E[0τ3eruH3,1(Xu(x),Su(s,x))du+0τ3erudSu(s,x)]E[0τ3eruH3,1(Xu(x),Su(s,x))du+0τ3erudSu(s,x)]=V3(x,s)>0(40) so that (x,s)C3,1 too. Thus, we may conclude that the left-hand boundary a3(s) is increasing on (L3,L3α/(α+1)), under α<1. Moreover, for any starting point (x,s)C3,1 such that s>s_3=L3, using the fact that the function H3,1(x,s) is decreasing in s on (L3,), under 1α<0, for each x>0 fixed, by means of arguments similar to the ones used above, we may conclude that V3(x,s)V3(x,s)>0 holds, for all s>s>L3, and thus, we have (x,s)C3,1, and the boundary a3(s) is increasing on (L3,), under 1α<0.

Let us now fix some (x,q)C3,2 such that K3α/(α+1)<q<q¯3=K3, under α>0, so that U3(x,q)>0 holds. Then, using the fact that the function H3,2(x,q)=r(qK3)(q/x)α is increasing in q on (K3α/(α+1),K3), under α>0, for each x>0 fixed, by virtue of the structure of the running minimum Q(q,x) of the process X(x), for any other starting point (x,q)E2 such that K3<q<q<K3α/(α+1), we have: (41) U3(x,q)E[0ζ3eruH3,2(Xu(x),Qu(q,x))du0ζ3erudQu(q,x)]E[0ζ3eruH3,2(Xu(x),Qu(q,x))du0ζ3erudQu(q,x)]=U3(x,q)>0(41) so that (x,q)C3,2 too. Thus, we may conclude that the right-hand boundary b3(q) is increasing on (K3α/(α+1),K3), under α>0 (see Figures  and for the computer drawings of locations of the optimal stopping boundaries a3(s) and b3(q) with respect to the points L3, L3α/(α+1) and K3, K3α/(α+1)).

Figure 5. A computer drawing of the continuation and stopping regions C3,1 and D3,1 formed by the optimal exercise boundary a3(s) and the points L3 and L3α/(α+1).

Figure 5. A computer drawing of the continuation and stopping regions C3,1∗ and D3,1∗ formed by the optimal exercise boundary a3∗(s) and the points L3 and L3α/(α+1).

2.3. The free-boundary problems

By means of standard arguments based on the application of Itô's formula, it is shown that the infinitesimal operator L of the process (X,S) or (X,Q) from (Equation2) and (Equation5) has the form: (42) L=(rδ)xx+σ2x22xxin0<x<sor0<q<x(42) (43) s=0at0<x=sorq=0at0<x=q(43) (see, e.g. [Citation37, Subsection 3.1]). In order to find analytic expressions for the unknown value functions Vi(x,s) and Ui(x,q) from (Equation25) and (Equation26) and the unknown boundaries ai(s) and bi(q) from (Equation38) and (Equation39), for every i = 1, 2, 3, we apply the results of general theory for solving optimal stopping problems for Markov processes presented in [Citation41, Chapter IV, Section 8] among others (see also [Citation41, Chapter V, Sections 15-20] for optimal stopping problems for maxima processes and other related references). More precisely, for the original optimal stopping problems in (Equation25) and (Equation26), we formulate the associated free-boundary problems (see, e.g. [Citation41, Chapter IV, Section 8]) and then verify in Theorem 4.1 below that the appropriate candidate solutions of the latter problems coincide with the solutions of the original problems. In other words, we reduce the optimal stopping problems of (Equation25) and (Equation26) to the following equivalent free-boundary problems: (44) (LVirVi)(x,s)=Hi,1(x,s)for(x,s)Ci,1{(x,s)E1|x=s}(44) (45) (LUirUi)(x,q)=Hi,2(x,q)for(x,q)Ci,2{(x,q)E2|x=q}(45) (46) Vi(x,s)|x=ai(s)+=0andUi(x,q)|x=bi(q)=0(46) (47) xVi(x,s)|x=ai(s)+=0andxUi(x,q)|x=bi(q)=0(47) (48) sVi(x,s)|x=s=sGi(s,s)andqUi(x,q)|x=q+=qFi(q,q)(48) (49) Vi(x,s)=0for(x,s)Di,1andUi(x,q)=0for(x,q)Di,2(49) (50) Vi(x,s)>0for(x,s)Ci,1andUi(x,q)>0for(x,q)Ci,2(50) (51) (LVirVi)(x,s)<Hi,1(x,s)for(x,s)Di,1(51) (52) (LUirUi)(x,q)<Hi,2(x,q)for(x,q)Di,2(52) where Ci,j and Di,j are defined as Ci,j and Di,j, for j = 1, 2, in (Equation38) and (Equation39) with the unknown functions ai(s) and bi(q) instead of ai(s) and bi(q), where the functions Hi,1(x,s) and Hi,2(x,q), for every i = 1, 2, 3, are defined in (Equation23) and (Equation24), respectively. Here, the instantaneous-stopping as well as the smooth-fit and normal-reflection conditions of (Equation46)–(Equation48) are satisfied, for all s>s_i and q<q¯i, for i = 1, 2, 3, with some 0s_1a¯1=rL1/δ and q¯1b_1=rK1/δ, as well as s_2=0, s_3=L3 and q¯2=, q¯3=K3. Observe that the superharmonic characterization of the value function (see, e.g. [Citation41, Chapter IV, Section 9]) implies that Vi(x,s) and Ui(x,q) are the smallest functions satisfying (Equation44)–(Equation46) and (Equation49)–(Equation50) with the boundaries ai(s) and bi(q), for every i = 1, 2, 3, respectively. Note that the inequalities in (Equation51) and (Equation52) follow directly from the arguments of parts (ii)–(iii) of Subsection 2.2 above.

3. Solutions to the free-boundary problems

In this section, we obtain solutions to the free-boundary problems in (Equation44)–(Equation52) and derive first-order nonlinear ordinary differential equations for the candidate optimal stopping boundaries.

3.1. The candidate value functions

It is shown that the second-order ordinary differential equations in (Equation44) and (Equation45) have the general solutions: (53) Vi(x,s)=Ci,1(s)xγ1+Ci,2(s)xγ2+Ai,1(s)x1αsα+Ai,2(s)xαsα(53) when α<0, for 0<xs, and (54) Ui(x,q)=Di,1(q)xγ1+Di,2(q)xγ2+Bi,1(q)x1αqα+Bi,2(q)xαqα(54) when α>0, for 0<qx, respectively. Here, Ci,j(s) and Di,j(q), for i=1,2,3 and j = 1, 2, are some arbitrary (continuously differentiable) functions, and γj, for j = 1, 2, are given by: (55) γj=12rδσ2(1)j(12rδσ2)2+2rσ2(55) so that γ2<0<1<γ1 holds. The functions Ai,j(s) and Bi,j(q), for i=1,2,3 and j = 1, 2, are specified by A1,1(s)=1, A1,2(s)=L1, A2,1(s)=L2, A2,2(s)=s, A3,1(s)=0, A3,2(s)=L3s, and B1,1(q)=1, B1,2(q)=K1, B2,1(q)=K2, B2,2(q)=q, B3,1(q)=0, B3,2(q)=qK3. Then, by applying the conditions of (Equation46)–(Equation48) to the functions in (Equation53) and (Equation54), we obtain the equalities: (56) Ci,1(s)aiγ1(s)+Ci,2(s)aiγ2(s)+Ai,1(s)ai1α(s)sα+Ai,2(s)aiα(s)sα=0(56) (57) γ1Ci,1(s)aiγ1(s)+γ2Ci,2(s)aiγ2(s)+Ai,1(s)(1α)ai1α(s)sαAi,2(s)αaiα(s)sα=0(57) (58) Ci,1(s)sγ1+Ci,2(s)sγ2+Ai,1(s)s+Ai,1(s)α+Ai,2(s)+Ai,2(s)α/s=sGi(s,s)(58) for all s>s_i, and (59) Di,1(q)biγ1(q)+Di,2(q)biγ2(q)+Bi,1(q)bi1α(q)qα+Bi,2(q)biα(q)qα=0(59) (60) γ1Di,1(q)biγ1(q)+γ2Di,2(q)biγ2(q)+Bi,1(q)(1α)bi1α(q)qαBi,2(q)αbiα(q)qα=0(60) (61) Di,1(q)qγ1+Di,2(q)qγ2+Bi,1(q)q+Bi,1(q)α+Bi,2(q)+Bi,2(q)α/q=qFi(q,q)(61) for all q<q¯i, respectively. Hence, by solving the systems of equations in (Equation56)–(Equation57) and (Equation59)–(Equation60), we obtain that the candidate value functions admit the representations: (62) Vi(x,s;ai(s))=Ci,1(s;ai(s))xγ1+Ci,2(s;ai(s))xγ2+Ai,1(s)x1αsα+Ai,2(s)xαsα(62) for ai(s)<xs and s>s_i, with (63) Ci,j(s;ai(s))=Ai,1(s)(γ3j+α1)ai(s)+Ai,2(s)(γ3j+α)(γjγ3j)aiγj+α(s)sα(63) for j = 1, 2, and (64) Ui(x,q;bi(q))=Di,1(q;bi(q))xγ1+Di,2(q;bi(q))xγ2+Bi,1(q)x1αqα+Bi,2(q)xαqα(64) for qx<bi(q) and q<q¯i, with (65) Di,j(q;bi(q))=Bi,1(q)(γ3j+α1)bi(q)+Bi,2(q)(γ3j+α)(γjγ3j)biγj+α(q)qα(65) for i = 1, 2, 3 and j = 1, 2, respectively. Moreover, by means of straightforward computations, it can be deduced from the expressions in (Equation62) and (Equation64) that the first- and second-order partial derivatives xVi(x,s;ai(s)) and xxVi(x,s;ai(s)) of the function Vi(x,s;ai(s)) take the form: (66) xVi(x,s;ai(s))=Ci,1(s;ai(s))γ1xγ11+Ci,2(s;ai(s))γ2xγ21+Ai,1(s)(1α)xαsαAi,2(s)αxα1sα(66) and (67) xxVi(x,s;ai(s))=Ci,1(s;ai(s))γ1(γ11)xγ12+Ci,2(s;ai(s))γ2(γ21)xγ22Ai,1(s)(1α)αxα1sα+Ai,2(s)α(α+1)xα2sα(67) on the interval ai(s)<xs, for each s>s_i and every i = 1, 2, 3 fixed, while the first- and second-order partial derivatives xUi(x,q;bi(q)) and xxUi(x,q;bi(q)) of the function Ui(x,q;bi(q)) take the form: (68) xUi(x,q;bi(q))=Di,1(q;bi(q))γ1xγ11+Di,2(q;bi(q))γ2xγ21+Bi,1(q)(1α)xαqαBi,2(q)αxα1qα(68) and (69) xxUi(x,q;bi(q))=Di,1(q;bi(q))γ1(γ11)xγ12+Di,2(q;bi(q))γ2(γ21)xγ22Bi,1(q)(1α)αxα1qα+Bi,2(q)α(α+1)xα2qα(69) on the interval qx<bi(q), for each q<q¯i and every i = 1, 2, 3 fixed.

3.2. The candidate stopping boundaries

By applying the conditions of (Equation58) and (Equation61) to the functions in (Equation63) and (Equation65), we conclude that the candidate boundaries satisfy the first-order nonlinear ordinary differential equations: (70) ai(s)=Ψi,1,1(s,ai(s))sγ1+Ψi,1,2(s,ai(s))sγ2Ξi,1(s)Φi,1,1(s,ai(s))sγ1+Φi,1,2(s,ai(s))sγ2(70) for s>s_i, and (71) bi(q)=Ψi,2,1(q,bi(q))qγ1+Ψi,2,2(q,bi(q))qγ2Ξi,2(q)Φi,2,1(q,bi(q))qγ1+Φi,2,2(q,bi(q))qγ2(71) for q<q¯i, respectively. Here, the functions Φ1,j(s,ai(s)), Ψ1,j(s,ai(s)) and Φ2,j(q,bi(q)), Ψ2,j(q,bi(q)) are defined by: (72) Φi,1,j(s,ai(s))=(γj+α1)(γ3j+α1)Ai,1(s)ai(s)+(γj+α)(γ3j+α)Ai,2(s)(γjγ3j)aiγj+α+1(s)sα(72) (73) Ψi,1,j(s,ai(s))=(Ai,1(s)s+Ai,1(s)α)(γ3j+α1)ai(s)+(Ai,2(s)s+Ai,2(s)α)(γ3j+α)(γjγ3j)aiγj+α(s)s1α(73) (74) Ξi,1(s)=sGi(s,s)+Ai,1(s)s+Ai,1(s)α+Ai,2(s)+Ai,2(s)α/s(74) for s>s_i, and (75) Φi,2,j(q,bi(q))=(γj+α1)(γ3j+α1)Bi,1(q)bi(q)+(γj+α)(γ3j+α)Bi,2(q)(γjγ3j)biγj+α+1(q)qα(75) (76) Ψi,2,j(q,bi(q))=(Bi,1(q)q+Bi,1(q)α)(γ3j+α1)bi(q)+(Bi,2(q)q+Bi,2(q)α)(γ3j+α)(γjγ3j)biγj+α(q)q1α(76) (77) Ξi,2(q)=qFi(q,q)+Bi,1(q)q+Bi,1(q)α+Bi,2(q)+Bi,2(q)α/q(77) for q<q¯i, and every i = 1, 2, 3 and j = 1, 2.

3.3. The maximal and minimal admissible solutions ai(s) and bi(q), i = 1, 2, 3

We further consider the maximal and minimal admissible solutions of first-order nonlinear ordinary differential equations as the largest and smallest possible solutions ai(s) and bi(q) of the equations in (Equation70) and (Equation71) with (Equation72)–(Equation73) and (Equation75)–(Equation76) which satisfy the inequalities ai(s)<sa¯i(s) and bi(q)>qb_i(q), for all s>s_i and q<q¯i, and every i = 1, 2, 3, with some 0s_1a¯1 and q¯1b_1 as well as s_2=0, s_3=L3 and q¯2=, q¯3=K3. Here, we recall that a¯1(s)a¯1=rL1/δ and b_1(q)b_1=rK1/δ as well as a¯2(s)=rs/(δL2) and b_2(q)=rq/(δK2), while a¯3(s)=s and b_3=q, for all s>0 and q>0. By virtue of the classical results on the existence and uniqueness of solutions for first-order nonlinear ordinary differential equations, we may conclude that these equations admit (locally) unique solutions, in view of the facts that the right-hand sides in (Equation70) and (Equation71) with (Equation72)–(Equation74) and (Equation75)–(Equation77) are (locally) continuous in (s,ai(s)) and (q,bi(q)) and (locally) Lipschitz in ai(s) and bi(q), for each s>s_i and q<q¯i fixed, and every i = 1, 2, 3 (see also [Citation37, Subsection 3.9] for similar arguments based on the analysis of other first-order nonlinear ordinary differential equations). Then, it is shown by means of technical arguments based on Picard's method of successive approximations that there exist unique solutions ai(s) and bi(q) to the equations in (Equation70) and (Equation71) with (Equation72)–(Equation73) and (Equation75)–(Equation76), for s>s_i and q<q¯i, started at some points (a¯i(si,0),si,0) and (b_i(qi,0),qi,0), for i=1,2,3, such that si,0>s_i and qi,0<q¯i, for every i=1,2,3 (see also [Citation23, Subsection 3.2] and [Citation37, Example 4.4] for similar arguments based on the analysis of other first-order nonlinear ordinary differential equations).

Hence, in order to construct the appropriate functions ai(s) and bi(q) which satisfy the equations in (Equation70) and (Equation71) and stays strictly above and below the appropriate diagonal, for s>s_i and q<q¯i, and every i = 1, 2, 3, respectively, we can follow the arguments from [Citation40, Subsection 3.5] (among others) which are based on the construction of sequences of the so-called bad-good solutions which intersect the upper or lower bounds or diagonals. For this purpose, for any sequences (si,l)lN and (qi,l)lN such that si,l>s_i and qi,l<q¯i as well as si,l and qi,l0 as l, we can construct the sequence of solutions ai,l(s) and bi,l(q), lN, to the equations (Equation70) and (Equation71), for all s>s_i and q<q¯i such that ai,l(si,l)=a¯i(si,l) and bi,l(qi,l)=b_i(qi,l) holds, for every i = 1, 2, 3 and each lN. It follows from the structure of the equations in (Equation70) and (Equation71) as well as the functions in (Equation72)–(Equation73) and (Equation75)–(Equation76) that the inequalities ai,l(si,l)<a¯i(si,l)1 and bi,l(qi,l)<b_i(qi,l)1 should hold for the derivatives of the appropriate functions, for each lN (see also [Citation36, pages 979-982] for the analysis of solutions of another first-order nonlinear differential equation). Observe that, by virtue of the uniqueness of solutions mentioned above, we know that each two curves sai,l(s) and sai,m(s) as well as qbi,l(q) and qbi,m(q) cannot intersect, for l,mN, lm, and thus, we see that the sequence (ai,l(s))lN is increasing and the sequence (bi,l(q))lN is decreasing, so that the limits ai(s)=limlai,l(s) and bi(q)=limlbi,l(q) exist, for each s>s_i and q<q¯i, and every i = 1, 2, 3, respectively. We may therefore conclude that ai(s) and bi(q) provides the maximal and minimal solutions to the equations in (Equation70) and (Equation71) such that ai(s)<a¯i(s)s and bi(q)>b_i(q)q holds, for all s>s_i and q<q¯i, with some 0s_1a¯1=rL1/δ and q¯1b_1=rK1/δ, as well as s_2=0, s_3=L3 and q¯2=, q¯3=K3.

Moreover, since the right-hand sides of the first-order nonlinear ordinary differential equations in (Equation70) and (Equation71) with (Equation72)–(Equation73) and (Equation75)–(Equation76) are (locally) Lipschitz in s and q, respectively, one can deduce by means of Gronwall's inequality that the functions ai,l(s) and bi,l(q), lN, are continuous, so that the functions ai(s) and bi(q) are continuous too, for every i=1,2,3. The appropriate maximal admissible solutions of first-order nonlinear ordinary differential equations and the associated maximality principle for solutions of optimal stopping problems which is equivalent to the superharmonic characterization of the payoff functions were established in [Citation37] and further developed in [Citation5,Citation14,Citation17–20,Citation22–25,Citation30,Citation35,Citation36,Citation39,Citation40,Citation43] among other subsequent papers (see also [Citation41, Chapter I; Chapter V, Section 17] for other references).

4. Main results and proofs

In this section, based on the expressions computed above, we formulate and prove the main results of the paper.

Theorem 4.1

Let the processes (X,S) and (X,Q) be given by (Equation1) and (Equation5), with some r>0, δ>0, and σ>0, and the inequality δ2rδσ2>0 be satisfied. Suppose that the random times θ and η are defined by (Equation4). Then, the value functions of the perpetual American withdrawable standard and lookback put and call options from (Equation25) and (Equation26) admit the expressions: (78) Vi(x,s)={Vi(x,s;ai(s)),ifai(s)<xsands>s_i0,if0<xai(s)ands>s_i0,if0<xss_i(78) whenever α2(rδ)/σ21<0, and (79) Ui(x,q)={Ui(x,q;bi(q)),ifqx<bi(q)and0<q<q¯i0,ifxbi(q)and0<q<q¯i0,ifxqq¯i(79) whenever α>0.

Here, the function Vi(x,s;ai(s)) is given by (Equation62) with (Equation63), whenever α<0, and the optimal exercise boundary ai(s) provides the maximal solution of the first-order nonlinear ordinary differential equation in (Equation70) with (Equation72)–(Equation74) satisfying the inequalities [a_i(s)<]ai(s)<a¯i(s)s, for all s>s_i and every i = 1, 2, 3, where a_1(s)=rL1α/(δ(α1)) and a¯1(s)=rL1/δ, a_2(s)=rsα/(δL2(α1)) and a¯2(s)=rs/(δL2), while a_i(s)=0 and a¯3(s)=s, under α<0, with some 0s_1a¯1 as well as s_2=0 and s_3=L3.

The function Ui(x,q;bi(q)) is given by (Equation64) with (Equation65), whenever α>0, and the optimal exercise boundary bi(q) provides the minimal solution of the first-order nonlinear ordinary differential equation in (Equation71) with (Equation75)–(Equation77) satisfying the inequalities b_i(q)q<bi(q)[<b¯i(q)], for all q<q¯i and every i = 1, 2, 3, where b_1(q)=rK1/δ and b¯1(q)=rK1α/(δ(α1)), b_2(q)=rq/(δK2) and b¯2(q)=rqα/(δK2(α1)), under α>1, as well as b¯1(q)= and b¯2(q)=, under 0<α1, while b_3(q)=q and b¯3(q)=, with some q¯1b_1 as well as q¯2= and q¯3=K3.

Since both parts of the assertion stated above are proved using similar arguments, we only give a proof for the case of the two-dimensional optimal stopping problem of (Equation26) related to the perpetual American withdrawable standard and lookback call options. Observe that we can put s = x and q = x to obtain the values of the original perpetual American withdrawable standard and lookback put and call option pricing problems of (Equation21) and (Equation22) from the values of the optimal stopping problems of (Equation25) and (Equation26).

Proof.

In order to verify the assertion stated above, it remains for us to show that the function defined in (Equation79) coincides with the value function in (Equation26) and that the stopping time ζi in (Equation27) is optimal with the boundary bi(q) specified above. For this purpose, let bi(q) be any solution of the ordinary differential equation in (Equation71) satisfying the inequality bi(q)>b_i(q)q, for all q<q¯i and every i = 1, 2, 3, where b_1(q)b_1=rK1/δ, b_2(q)=rq/(δK2), and b_3(q)=q, with some q¯1b_1 as well as q¯2= and q¯3=K3. Let us also denote by Uibi(x,q) the right-hand side of the expression in (Equation79) associated with bi(q), for every i = 1, 2, 3. Then, it is shown by means of straightforward calculations from the previous section that the function Uibi(x,q) solves the system of (Equation45) with the right-hand sides of (Equation49)–(Equation50) and (Equation52) and satisfies the right-hand conditions of (Equation46)–(Equation48). Recall that the function Uibi(x,q) is C2,1 on the closure C¯i,2 of Ci,2 and is equal to zero on Di,2, which are defined as C¯i,2, Ci,2 and Di,2 in (Equation38) and (Equation39) with bi(q) instead of bi(q), for i = 1, 2, 3, respectively. Hence, taking into account the assumption that the boundary bi(q) is continuously differentiable, for all q<q¯i, by applying the change-of-variable formula from [Citation38, Theorem 3.1] to the process ertUibi(Xt,Qt) (see also [Citation41, Chapter II, Section 3.5] for a summary of the related results and further references), we obtain the expression: (80) ertUibi(Xt,Qt)=Uibi(x,q)+0teru(LUibirUibi)(Xu,Qu)I(Xubi(Qu),XuQu)du+0teruqUibi(Xu,Qu)I(Xu=Qu)dQu+Mti(80) for all t0, for every i = 1, 2, 3. Here, the process Mi=(Mti)t0 defined by: (81) Mti=0teruxUibi(Xu,Qu)I(XuQu)σXudBu(81) is a continuous local martingale with respect to the probability measure Px,q. Note that, since the time spent by the process (X,Q) at the boundary surface Ci,2={(x,q)E2|x=bi(q)} as well as at the diagonal d2={(x,q)E2|x=q} is of the Lebesgue measure zero (see, e.g. [Citation7, Chapter II, Section 1]), the indicators in the second line of the formula in (Equation80) as well as in the expression of (Equation81) can be ignored. Moreover, since the component Q decreases only when the process (X,Q) is located on the diagonal d2={(x,q)E2|x=q}, the indicator in the third line of (Equation80) can also be set equal to one. Observe that the integral in the third line of (Equation80) will actually be compensated accordingly, due to the fact that the candidate value function Uibi(x,q) satisfies the modified normal-reflection condition of the right-hand part of (Equation48) at the diagonal d2.

It follows from straightforward calculations and the arguments of the previous section that the function Uibi(x,q) satisfies the second-order ordinary differential equation in (Equation45), which together with the right-hand conditions of (Equation46)–(Equation47) and (Equation49) as well as the fact that the inequality in (Equation52) holds imply that the inequality (LUibirUibi)(x,q)Hi,2(x,q) is satisfied with Hi,2(x,q) given by (Equation24), for all 0<q<x such that xbi(q), and i = 1, 2, 3. Moreover, we observe directly from the expressions in (Equation64) and (Equation68) and (Equation69) with (Equation65) that the function Uibi(x,q) is convex and decreases to zero, because its first-order partial derivative xUibi(x,q) is negative and increases to zero, while its second-order partial derivative xxUibi(x,q) is positive, on the interval qx<bi(q), under α>0, for each q<q¯i and every i=1,2,3 fixed. Thus, we may conclude that the right-hand inequality in (Equation50) holds, which together with the right-hand conditions of (Equation46)–(Equation47) and (Equation49) imply that the inequality Uibi(x,q)0 is satisfied, for all (x,q)E2. Let (κi,n)nN be the localizing sequence of stopping times for the process Mi from (Equation81) such that κi,n=inf{t0||Mti|n}, for each nN. It therefore follows from the expression in (Equation80) that the inequalities: (82) 0ζκi,neruHi,2(Xu,Qu)du+0ζκi,neruqFi(Qu,Qu)dQuer(ζκi,n)Ui(Xζκi,n,Qζκi,n)+0ζκi,neruHi,2(Xu,Qu)du+0ζκi,neruqFi(Qu,Qu)dQuUibi(x,q)+Mτκi,ni(82) hold, for any stopping time ζ of the process X and each nN fixed. Then, taking the expectation with respect to Px,q in (Equation82), by means of Doob's optional sampling theorem, we get: (83) Ex,q[0ζκi,neruHi,2(Xu,Qu)du+0ζκi,neruqFi(Qu,Qu)dQu]Ex,q[er(ζκi,n)Uibi(Xζκi,n,Qζκi,n)0ζκi,n+0ζκi,neruHi,2(Xu,Qu)du+0ζκi,neruqFi(Qu,Qu)dQu]Uibi(x,q)+Ex,q[Mζκi,ni]=Uibi(x,q)(83) for all 0<qx and every i = 1, 2, 3, and each nN. Hence, letting n go to infinity and using Fatou's lemma, we obtain from the expressions in (Equation83) that the inequalities: (84) Ex,q[0ζeruHi,2(Xu,Qu)du+0ζeruqFi(Qu,Qu)dQu]Ex,q[erζUibi(Xζ,Qζ)+0ζeruHi,2(Xu,Qu)du+0ζeruqFi(Qu,Qu)dQu]Uibi(x,q)(84) are satisfied, for any stopping time ζ and all 0<qx such that q<q¯i, for i = 1, 2, 3. Thus, taking the supremum over all stopping times ζ and then the infimum over all boundaries b in the expressions of (Equation84), we may therefore conclude that the inequalities: (85) supζEx,q[0ζeruHi,2(Xu,Qu)du+0ζeruqFi(Qu,Qu)dQu]infbiUibi(x,q)=Uibi(x,q)(85) hold, for all 0<qx, where bi(q) is the minimal solution of the ordinary differential equation in (Equation71) as well as satisfying the inequality bi(q)>b_i(q)q, for all q<q¯i and every i = 1, 2, 3. By using the fact that the function Uibi(x,q) is (strictly) increasing in the value bi(q), for each q<q¯i fixed, we see that the infimum in (Equation85) is attained over any sequence of solutions (bi,m(q))mN to (Equation71) satisfying the inequality bi,m(q)>b_i(q)q, for all q<q¯i, for each mN and every i = 1, 2, 3, and such that bi,m(q)bi(q) as m, for each q<q¯i fixed, and every i = 1, 2, 3. It follows from the (local) uniqueness of the solutions to the first-order (nonlinear) ordinary differential equation in (Equation71) that no distinct solutions intersect, so that the sequence (bi,m(q))mN is decreasing and the limit bi(q)=limmbi,m(q) exists, for each q<q¯i fixed. Since the inequalities in (Equation84) hold for bi(q) too, we see that the expression in (Equation85) holds, for bi(q) and (x,q)E2, as well. We also note from the inequality in (Equation83) that the function Uibi(x,q) is superharmonic for the Markov process (X,Q) on E2. Hence, taking into account the facts that Uibi(x,q) is increasing in bi(q)>b_i(q)q, for all q<q¯i and every i = 1, 2, 3, and the inequality Uibi(x,q)0 holds, for all (x,q)E2, we observe that the selection of the minimal solution bi(q) which stays strictly above the diagonal d2={(x,q)E2|x=q} and the curve x=b_i(q), for i = 1, 2, 3, is equivalent to the implementation of the superharmonic characterization of the value function as the smallest superharmonic function dominating the payoff function (cf. [Citation37] or [Citation41, Chapter I and Chapter V, Section 17]).

In order to prove the fact that the boundary bi(q) is optimal, we consider the sequence of stopping times ζi,m, mN, defined as in the right-hand part of (Equation27) with bi,m(q) instead of bi(q), where bi,m(q) is a solution to the first-order ordinary differential equation in (Equation71) and such that bi,m(q)bi(q) as m, for each q<q¯i and every i = 1, 2, 3 fixed. Then, by virtue of the fact that the function Uibi,m(x,q) from the right-hand side of the expression in (Equation79) associated with the boundary bi,m(q) satisfies the equation of (Equation45) and the right-hand condition of (Equation46), and taking into account the structure of ζi in (Equation27), it follows from the expression which is equivalent to the one in (Equation80) that the equalities: (86) 0ζi,mκi,neruHi,2(Xu,Qu)du+0ζi,mκi,neruqFi(Qu,Qu)dQu=er(ζi,mκi,n)Uibi,m(Xζi,mκi,n,Qζi,mκi,n)+0ζi,mκi,neruHi,2(Xu,Qu)du+0ζi,mκi,neruqFi(Qu,Qu)dQu=Uibi,m(x,q)+Mζi,mκi,ni(86) hold, for all 0<qx such that q<q¯i, for each n,mN and every i=1,2,3. Observe that, by virtue of the arguments from [Citation45, Chapter VIII, Section 2a], the property: (87) Ex,q[supt0(0ζiteruHi,2(Xu,Qu)du+0ζiteruqFi(Qu,Qu)dQu)]<(87) holds, for all (x,q)E2. Hence, letting m and n go to infinity and using the condition of (Equation46) as well as the property ζi,mζi (Px,q-a.s.) as m, we can apply the Lebesgue dominated convergence theorem to the appropriate (diagonal) subsequence in the expression of (Equation86) to obtain the equality: (88) Ex,q[0ζieruHi,2(Xu,Qu)du+0ζieruqFi(Qu,Qu)dQu]=Uibi(x,q)(88) for all 0<xq such that q<q¯i and every i = 1, 2, 3, which together with the inequalities in (Equation85) directly implies the desired assertion. We finally recall that the results of part (ii) of the proof of Theorem 2.1 above, which are obtained by standard comparison arguments applied to the value functions of the appropriate optimal stopping problems, show that the inequality bi(q)<b¯i(q), for all 0<q<q¯i and every i = 1, 2, 3, should hold for the optimal exercise boundary, that completes the verification.

Acknowledgements

The authors are grateful to the Editor and two anonymous Referees for their valuable suggestions which helped to improve the presentation of the paper.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • S. Asmussen, F. Avram, and M. Pistorius, Russian and American put options under exponential phase-type Lévy models, Stochastic Process. Appl. 109 (2003), pp. 79–111.
  • F. Avram, A.E. Kyprianou, and M. Pistorius, Exit problems for spectrally negative Lévy processes and applications to (Canadized) Russian options, Ann. Appl. Probab. 14(1) (2004), pp. 215–238.
  • E.J. Baurdoux and A.E. Kyprianou, Further calculations for Israeli options, Stochastics 76 (2004), pp. 549–569.
  • E.J. Baurdoux and A.E. Kyprianou, The McKean stochastic game driven by a spectrally negative Lévy process, Electron. J. Probab. 8 (2008), pp. 173–197.
  • E.J. Baurdoux and A.E. Kyprianou, The Shepp-Shiryaev stochastic game driven by a spectrally negative Lévy process, Theory Probab. Appl. 53 (2009), pp. 481–499.
  • E.J. Baurdoux, A.E. Kyprianou, and J.C. Pardo, The Gapeev-Kühn stochastic game driven by a spectrally positive Lévy process, Stochastic Process. Appl. 121(6) (2011), pp. 1266–1289.
  • A.N. Borodin and P. Salminen, Handbook of Brownian Motion, 2nd ed, Birkhäuser, Basel, 2002.
  • J. Detemple, American-Style Derivatives: Valuation and Computation, Chapman and Hall/CRC, Boca Raton, 2006.
  • L. Dubins, L.A. Shepp, and A.N. Shiryaev, Optimal stopping rules and maximal inequalities for Bessel processes, Theory Probab. Appl. 38(2) (1993), pp. 226–261.
  • M. Egami, T. Leung, and K. Yamazaki, Default swap games driven by spectrally negative Lévy processes, Stochastic Process. Appl. 123(2) (2013), pp. 347–384.
  • E. Ekström and G. Peskir, Optimal stopping games for Markov processes, SIAM J. Control Optim. 47(2) (2008), pp. 684–702.
  • E. Ekström and S. Villeneuve, On the value of optimal stopping games, Ann. Appl. Probab. 16(3) (2006), pp. 1576–1596.
  • G. Ferreyra and P. Sundar, Comparison of solutions of stochastic differential eqations and applications, Stochastic Anal. Appl. 18 (2000), pp. 211–229.
  • P.V. Gapeev, Discounted optimal stopping for maxima of some jump-diffusion processes, J. Appl. Probab. 44 (2007), pp. 713–731.
  • P.V. Gapeev and H. Al Motairi, Discounted optimal stopping problems in first-passage time models with random thresholds, (2021), to appear in J. Appl. Probab., 20 pp.
  • P.V. Gapeev and L. Li, Perpetual American defaultable standard and lookback options in models with incomplete information, (2021), submitted, 30 pp.
  • P.V. Gapeev and N. Rodosthenous, Optimal stopping problems in diffusion-type models with running maxima and drawdowns, J. Appl. Probab. 51(3) (2014), pp. 799–817.
  • P.V. Gapeev and N. Rodosthenous, On the drawdowns and drawups in diffusion-type models with running maxima and minima, J. Math. Anal. Appl. 434(1) (2015), pp. 413–431.
  • P.V. Gapeev and N. Rodosthenous, Perpetual American options in diffusion-type models with running maxima and drawdowns, Stochastic Process. Appl. 126(7) (2016), pp. 2038–2061.
  • P.V. Gapeev, P.M. Kort, and M.N. Lavrutich, Discounted optimal stopping problems for maxima of geometric Brownian motions with switching payoffs, Adv. Appl. Probab. 53(1) (2021), pp. 189–219.
  • P.V. Gapeev, L. Li, and Z. Wu, Perpetual American cancellable standard options in models with last passage times, Algorithms 14(1) (2021), p. 3.
  • K. Glover, H. Hulley, and G. Peskir, Three-dimensional Brownian motion and the golden ratio rule, Ann. Appl. Probab. 23 (2013), pp. 895–922.
  • S.E. Graversen and G. Peskir, Optimal stopping and maximal inequalities for geometric Brownian motion, J. Appl. Probab. 35(4) (1998), pp. 856–872.
  • X. Guo and L.A. Shepp, Some optimal stopping problems with nontrivial boundaries for pricing exotic options, J. Appl. Probab. 38(3) (2001), pp. 647–658.
  • X. Guo and M. Zervos, π options, Stochastic Process. Appl. 120(7) (2010), pp. 1033–1059.
  • J. Kallsen and C. Kühn, Pricing derivatives of American and game type in incomplete markets, Finance Stoch. 8(2) (2004), pp. 261–284.
  • Y. Kifer, Game options, Finance Stoch. 4 (2000), pp. 443–463.
  • C. Kühn and A.E. Kyprianou, Callable puts as composite exotic options, Math. Finance 17(4) (2007), pp. 487–502.
  • A.E. Kyprianou, Some calculations for Israeli options, Finance Stoch. 8(1) (2004), pp. 73–86.
  • A.E. Kyprianou and C. Ott, A capped optimal stopping problem for the maximum process, Acta Appl. Math. 129 (2014), pp. 147–174.
  • T. Leung and K. Yamazaki, American step-up and step-down default swaps under Lévy models, Quant. Finance 13(1) (2013), pp. 137–157.
  • B. Li, N.L. Vu, and X. Zhou, Exit problems for general draw-down times of spectrally negative Lévy processes, J. Appl. Probab. 56(2) (2019), pp. 441–457.
  • R. Mansuy and M. Yor, Random Times and Enlargements of Filtration in a Brownian Setting, Lecture Notes in Mathematics 1873, Springer, Berlin, 2006.
  • A. Nikeghbali and M. Yor, Doob's maximal identity, multiplicative decomposition and enlargement of filtrations, Illinois J. Math. 50(4) (2006), pp. 791–814.
  • C. Ott, Optimal stopping problems for the maximum process with upper and lower caps, Ann. Appl. Probab. 23 (2013), pp. 2327–2356.
  • J.L. Pedersen, Discounted optimal stopping problems for the maximum process, J. Appl. Probab. 37(4) (2000), pp. 972–983.
  • G. Peskir, Optimal stopping of the maximum process: The maximality principle, Ann. Probab. 26(4) (1998), pp. 1614–1640.
  • G. Peskir, A change-of-variable formula with local time on surfaces, in Séminaire de Probabilité XL, Lecture Notes in Mathematics 1899, Springer, Berlin, 2007, pp. 69–96.
  • G. Peskir, Optimal detection of a hidden target: The median rule, Stochastic Process. Appl. 122 (2012), pp. 2249–2263.
  • G. Peskir, Quickest detection of a hidden target and extremal surfaces, Ann. Appl. Probab. 24(6) (2014), pp. 2340–2370.
  • G. Peskir and A.N. Shiryaev, Optimal Stopping and Free-Boundary Problems, Birkhäuser, Basel, 2006.
  • D. Revuz and M. Yor, Continuous Martingales and Brownian Motion, Springer, Berlin, 1999.
  • N. Rodosthenous and M. Zervos, Watermark options, Finance Stoch. 21(1) (2017), pp. 157–186.
  • L.A. Shepp and A.N. Shiryaev, The Russian option: Reduced regret, Ann. Appl. Probab. 3(3) (1993), pp. 631–640.
  • A.N. Shiryaev, Essentials of Stochastic Finance, World Scientific, Singapore, 1999.