Research article Special Issues

Optimal control in pharmacokinetic drug administration


  • We consider a two-box model for the administration of a therapeutic substance and discuss two scenarios: First, the substance should have an optimal therapeutic concentration in the central compartment (typically blood) and be degraded in an organ, the peripheral compartment (e.g., the liver). In the other scenario, the concentration in the peripheral compartment should be optimized, with the blood serving only as a means of transport. In either case the corresponding optimal control problem is to determine a dosing schedule, i.e., how to administer the substance as a function u of time to the central compartment so that the concentration of the drug in the central or in the peripheral compartment remains as closely as possible at its optimal therapeutic level. We solve the optimal control problem for the central compartment explicitly by using the calculus of variations and the Laplace transform. We briefly discuss the effect of the approximation of the Dirac delta distribution by a bolus. The optimal control function u for the central compartment satisfies automatically the condition u0. But for the peripheral compartment one has to solve an optimal control problem with the non-linear constraint u0. This problem does not seem to be widely studied in the current literature in the context of pharmacokinetics. We discuss this question and propose two approximate solutions which are easy to compute. Finally we use Pontryagin's Minimum Principle to deduce the exact solution for the peripheral compartment.

    Citation: Norbert Hungerbühler. Optimal control in pharmacokinetic drug administration[J]. Mathematical Biosciences and Engineering, 2022, 19(5): 5312-5328. doi: 10.3934/mbe.2022249

    Related Papers:

    [1] Min-Ku Lee, Jeong-Hoon Kim . Pricing vanilla, barrier, and lookback options under two-scale stochastic volatility driven by two approximate fractional Brownian motions. AIMS Mathematics, 2024, 9(9): 25545-25576. doi: 10.3934/math.20241248
    [2] Shoude Huang, Xin-Jiang He . An analytical approximation formula for European option prices under a liquidity-adjusted non-affine stochastic volatility model. AIMS Mathematics, 2022, 7(6): 10364-10377. doi: 10.3934/math.2022577
    [3] Weiwei Shen, Yan Zhang . Strong convergence of the Euler-Maruyama method for the stochastic volatility jump-diffusion model and financial applications. AIMS Mathematics, 2025, 10(5): 12032-12054. doi: 10.3934/math.2025545
    [4] Shou-de Huang, Xin-Jiang He . Analytical approximation of European option prices under a new two-factor non-affine stochastic volatility model. AIMS Mathematics, 2023, 8(2): 4875-4891. doi: 10.3934/math.2023243
    [5] Shoude Huang, Xinjiang He, Shuqu Qian . An analytical approximation of European option prices under a hybrid GARCH-Vasicek model with double exponential jump in the bid-ask price economy. AIMS Mathematics, 2024, 9(5): 11833-11850. doi: 10.3934/math.2024579
    [6] Yao Fu, Sisi Zhou, Xin Li, Feng Rao . Multi-assets Asian rainbow options pricing with stochastic interest rates obeying the Vasicek model. AIMS Mathematics, 2023, 8(5): 10685-10710. doi: 10.3934/math.2023542
    [7] Sun-Yong Choi, Donghyun Kim, Ji-Hun Yoon . An analytic pricing formula for timer options under constant elasticity of variance with stochastic volatility. AIMS Mathematics, 2024, 9(1): 2454-2472. doi: 10.3934/math.2024121
    [8] Hijaz Ahmad, Muhammad Nawaz Khan, Imtiaz Ahmad, Mohamed Omri, Maged F. Alotaibi . A meshless method for numerical solutions of linear and nonlinear time-fractional Black-Scholes models. AIMS Mathematics, 2023, 8(8): 19677-19698. doi: 10.3934/math.20231003
    [9] Chao Yue, Chuanhe Shen . Fractal barrier option pricing under sub-mixed fractional Brownian motion with jump processes. AIMS Mathematics, 2024, 9(11): 31010-31029. doi: 10.3934/math.20241496
    [10] Jiajia Zhao, Zuoliang Xu . Calibration of time-dependent volatility for European options under the fractional Vasicek model. AIMS Mathematics, 2022, 7(6): 11053-11069. doi: 10.3934/math.2022617
  • We consider a two-box model for the administration of a therapeutic substance and discuss two scenarios: First, the substance should have an optimal therapeutic concentration in the central compartment (typically blood) and be degraded in an organ, the peripheral compartment (e.g., the liver). In the other scenario, the concentration in the peripheral compartment should be optimized, with the blood serving only as a means of transport. In either case the corresponding optimal control problem is to determine a dosing schedule, i.e., how to administer the substance as a function u of time to the central compartment so that the concentration of the drug in the central or in the peripheral compartment remains as closely as possible at its optimal therapeutic level. We solve the optimal control problem for the central compartment explicitly by using the calculus of variations and the Laplace transform. We briefly discuss the effect of the approximation of the Dirac delta distribution by a bolus. The optimal control function u for the central compartment satisfies automatically the condition u0. But for the peripheral compartment one has to solve an optimal control problem with the non-linear constraint u0. This problem does not seem to be widely studied in the current literature in the context of pharmacokinetics. We discuss this question and propose two approximate solutions which are easy to compute. Finally we use Pontryagin's Minimum Principle to deduce the exact solution for the peripheral compartment.



    Following the seminal papers by Black and Scholes [3] and Merton [25] on the pricing formula for European vanilla options, there have been active researches also for path-dependent exotic options based on the geometric Brownian motion framework with constant volatility. The papers by Merton [25] himself and Reiner and Rubinstein [29] for barrier options, Goldman et al. [16] and Conze and Vishwanathan [6] for lookback options are few examples. Refer to Clewlow et al. [5] for a review study of path-dependent options under the Black-Scholes model. Based on the knowledge that the Black-Scholes model with constant volatility does not account for empirically observed phenomena such as volatility smile effect, there have been extensions to local or stochastic volatility model cases. For instance, Davydov and Linetsky [9] obtained closed-form solutions for barrier and lookback options under the constant elasticity of variance (CEV) model of Cox [7] and Cox and Ross [8]. Park and Kim [28] derived an infinite series form of a pricing formula for a lookback option under a general stochastic volatility model. Kato et al. [20] obtained a semi closed-form approximation formula for the price of a barrier option under a certain type of stochastic volatility model covering the stochastic alpha-beta-rho (SABR) model of Hagan et al. [17]. Aquino and Bernard [1] derived semi-analytical pricing formulas for lookback and barrier options under the Heston model. Kim et al. [21] obtained an analytic approximation formula for the price of an external barrier option under a fast mean-reverting stochastic volatility model.

    In this paper, we consider two types of perpetual American options with a path-dependent exotic structure, namely, stop-loss and Russian options. A stop-loss option is a perpetual style option with the structure of both 'knock-in' barrier and lookback options. This option was introduced by Fitt et al. [11] in 1994. If the underlying price reaches a maximum value St at time t and then never goes up beyond St and falls back into a given proportion, say λ, of it at later time t, then the option is knocked in and exercised in such a way that the option holder receives the amount St (=λSt) at that time, where 0<λ<1 is given as a predetermined value. In this case, the exercise time is a stopping time. On the other hand, a Russian option is a perpetual American option with a free boundary that contains a lookback provision. It was first proposed by Shepp and Shiryaev [31] in 1993. At any time t chosen by the option holder, this option pays out the maximum realized asset price St up to that time if the holder wants to claim it. We refer readers to Wilmott et al. [32] for more details on these two options together with the corresponding pricing formulas under the geometric Brownian motion with constant volatility.

    Obtaining analytic pricing formulas for stop-loss and Russian options under a stochastic volatility model is a challenging task because of the complicated exotic nature of these options. The contribution of this work is to derive closed-form formulas explicitly for the approximate prices of these two options under a multiscale stochastic volatility model. To the best of our knowledge, there was no previous report on the formulas in such a multiscale volatility environment. The accuracy of the analytic formulas are verified via Monte-Carlo simulations. The impacts of the multisale stochastic volatility model on the corresponding Black-Scholes prices of those exotic options are disclosed. The performance of the model is compared with that of other models.

    The rest of the paper is organized as follows. In Section 2, we discuss a multiscale stochastic volatility model formulation for the underlying asset prices and the important features of stop-loss and Russian options. Section 3 provides a detailed discussion on how an asymptotic expansion approach can yield partial differential equations (PDEs) for the prices of stop-loss and Russian options and the subsequent ordinary differential equation (ODE) problems. In Section 4, we derive explicitly the closed-form solutions of the PDE problems for the leading-order terms and the first-order corrections. Section 5 verifies that the results given by those analytic formulas match well with those generated by Monte-Carlo simulations and presents the impacts of the multiscale stochastic volatility model on the Black-Scholes option prices and a comparison with other models. Section 6 states a concluding remark. In Appendices A and B, we derive differential equations and their closed-form solutions for the second-order corrections. Appendix C describes the explicit representations of some functions in Appendix A.

    For the price St of a given underlying asset (stock or market index), we consider

    dSt=(rq)Stdt+f(Yt,Zt)StdWst,dYt=(1ϵα(Yt)1ϵβ(Yt)Λ(Yt,Zt))dt+1ϵβ(Yt)dWyt,dZt=(δc(Zt)δg(Zt)Γ(Yt,Zt))dt+δg(Zt)dWzt (2.1)

    under a martingale probability measure Q, where and r and q are risk-free interest and dividend rates, respectively, the function f is smooth and bounded on R2, Λ and Γ represent the market prices of volatility risk, the functions α and β are given in such a way that Yt is an ergodic process that admits a unique invariant distribution, denoted by Φ, and the functions c and g are smooth on R and at most linearly growing infinitely. Wxt, Wyt and Wzt are standard Brownian motions with a correlation structure given by

    dWs,Wyt=ρsydt,dWs,Wzt=ρszdt,dWy,Wzt=ρyzdt.

    Moreover, the constants ϵ and δ are such that 0<δϵδ1. This type of multiscale stochastic volatility model was proposed by Fouque et al. [13] and the extensive study of pricing several types of derivatives under this model can be found in the book of Fouque et al. [14].

    In this paper, we study an evaluation problem of stop-loss options and Russian options under the underlying asset price dynamics given by (2.1). Both options contain no expiration date and they have a lookback provision. We recall that the no-arbitrage price, P(t,s,y,z), of a perpetual American option can be expressed as

    P(t,s,y,z)=suptτEQ[er(τt)h(τ)|St=s,Yt=y,Zt=z],

    where τ is a stopping time and h(τ) denotes the payoff that depends on the 'path' of the underlying price up to time τ. We note that the starting time t does not matter when pricing perpetual options because of an infinite time horizon. Hence, one can write the option price P(t,s,y,z) as P(s,y,z) with the t-dependence. On the other hand, to deal with a lookback type of option, we need to define the maximum value of the underling asset price until arbitrary time t as

    St=sup0utSu,0<t<,

    which becomes another independent variable for the evaluation of options of interest.

    The price of a stop-loss option depends on the underlying asset price, the maximum value of it, the levels of the two volatility driving processes and the pre-determined level λ. It is denoted by Ps/l(s,s,y,z). The payoff h of this option is given by

    h(τ)=Sτ1Sτ=λSτ,

    where '1' stands for the indicator function.

    The price of a Russian option also depends on the underlying asset price, the maximum value of it, the levels of the two volatility driving processes. It is denoted by PR(s,s,y,z). For a Russian option, the payoff h is given by

    h(τ)=Sτ.

    In this work, we use the combined asymptotic expansion and partial differential equation (PDE) approach to evaluate the prices of these two options based on the Feynman-Kac theorem (see Oksendal [26] for example), a link between parabolic PDEs and stochastic processes. From this theorem, the no-arbitrage price P(s,s,y,z) of a perpetual option, which could be a stop-loss option or a Russian option, satisfies

    [1ϵL0+1ϵL1+L2+δϵM3+δM1+δM2]P(s,s,y,z)=0, (2.2)

    where the operators L0, L1, L2, M3, M1 and M2 are given by

    L0=α(y)y+12β2(y)2y2,L1=β(y)(ρsyf(y,z)s2syΛ(y,z)y),L2=12f2(y,z)s22s2+(rq)ssr,M3=ρyzβ(y)g(z)2yz,M1=g(z)(ρszf(y,z)s2szΓ(y,z)z),M2=c(z)z+12g2(z)2z2, (2.3)

    respectively. We are going to employ the asymptotic analysis of Fouque et al. [14] to derive an approximate solution of this singularly perturbed PDE with appropriate boundary conditions imposed for each of the stop-loss and Russian options.

    In the following argument, the following lemma for a Poisson equation is very useful.

    Lemma 2.1. We consider the Poisson equation

    L0X(y)+G(y)=0.

    (a) The existence of a solution for the Poisson equation requires the following condition (called the centering condition):

    G:=RG(y)Φ(y)dy=0,

    where Φ is the invariant distribution of the process Yt.

    (b) If the function G is zero and the solution X does not move as fast as

    Xye2Rα(y)β2(y)dy,y,

    then X is independent of variable y.

    Proof. Refer to Ramm [30] (Fredholm alternative theorem) and Fouque et al. [14] for (a) and (b), respectively.

    In this section, we establish PDE problems for the price Ps/l(s,s,y,z) of a stop-loss option and the price PR(s,s,y,z) of a Russian option, respectively.

    The option price Ps/l(s,s,y,z) satisfies the PDE (2.2) on the interval λs<s<s and boundary conditions given by

    Ps/l(s=λs,s,y,z)=λs,Ps/ls(s=s,s,y,z)=0.

    In addition to these boundary conditions, one might need to use the following linear scaling property:

    Ps/l(νs,νs,y,z)=νPs/l(s,s,y,z).

    For dimensionality reduction, we use the change of independent and dependent variables, sx and Ps/lVs/l, defined by

    x=s/s,Ps/l(s,s,y,z)=sVs/l(x,y,z).

    Then we obtain a PDE problem for Vs/l (instead of Ps/l) as follows:

    LVs/l(x,y,z)=0,λ<x<1,Vs/l(x=λ,y,z)=λ,Vs/l(x=1,y,z)=Vs/lx(1,y,z), (3.1)

    where the operator L is given by

    L=1ϵL0+1ϵL1+L2+δϵM3+δM1+δM2,L1:=β(y)(ρxyf(y,z)x2xyΛ(y,z)y),L2:=12f2(y,z)x22x2+(rq)xxr,M1:=g(z)(ρxzf(y,z)x2xzΓ(y,z)z). (3.2)

    Note that L0, M2 and M3 are the same as in (2.3).

    We are interested in the solution Vs/l(x,y,z) of the expansion form

    Vs/l(x,y,z)=i,j=0ϵi/2δj/2Vs/lij(x,y,z), (3.3)

    where Vs/lij are assumed to satisfy the growth condition stated in Lemma 2.1 (b) so that the solution can reflect the realistic situation in market. In the rest of this section, we are going to derive PDE problems for the terms Vs/lij with (ij)=(0,0), (0,1) and (1,0). First, rearranging the PDE (3.1) by using the series (3.3), we can get

    1ϵL0Vs/l00+δϵL0Vs/l01+δϵL0Vs/l02+1ϵ(L0Vs/l10+L1Vs/l00)+δϵ(L0Vs/l11+L1Vs/l01+M3Vs/l00)+δϵ(L0Vs/l12+L1Vs/l02+M3Vs/l01)+L0Vs/l20+L1Vs/l10+L2Vs/l00+δ(L0Vs/l21+L1Vs/l11+L2Vs/l01+M1Vs/l00+M3Vs/l10)+δ(L0Vs/l22+L1Vs/l12+L2Vs/l02+M1Vs/l01+M2Vs/l00+M3Vs/l11)+ϵ(L0Vs/l30+L1Vs/l20+L2Vs/l10)+ϵδ(L0Vs/l31+L1Vs/l21+L2Vs/l11+M1Vs/l10+M3Vs/l20)+ϵ(L0Vs/l40+L1Vs/l30+L2Vs/l20)+=0. (3.4)

    The following proposition says that the first few terms of the asymptotic expansion (3.3) are independent of variable y.

    Proposition 3.1. The terms Vs/lij with i=0,1 and j=0,1,2 in the asymptotic series (3.3) for the stop-loss option price Vs/l are independent of y; Vs/lij(x,y,z)=Vs/lij(x,z).

    Proof. From the terms of order 1ϵ, δϵ and δϵ in (3.4), we have the Poisson equations L0Vs/l0j=0(j=0,1,2). Then, by Lemma 2.1 (a), Vs/l0j are independent of y for j=0, 1 and 2. Since L1Vs/l0j=0(j=0,1,2) and M3Vs/l0j=0(j=0,1), we have the Poisson equations L0Vs/l1j=0(j=0,1,2) from the terms of order 1ϵ, δϵ and δϵ in (3.4). Thus Vs/l1j are independent of y for j=0, 1 and 2.

    In the following argument, we use an operator, ¯L, defined as

    ¯L:=L2=12σ2(z)x22x2+(rq)xxr,σ(z):=f2(,z). (3.5)

    Also, we use the functions ϕ1, ϕ2, ψ, ξ1 and ξ2 defined by the solutions of

    L0ϕ1(y,z)=f(y,z)f(,z),L0ϕ2(y,z)=f2(y,z)f2(,z),L0ψ(y,z)=Γ(y,z)Γ(,z),L0ξ1(y,z)=β(y)f(y,z)ϕ2(y,z)yβ()f(,z)ϕ2(,z)y,L0ξ2(y,z)=β(y)Λ(y,z)ϕ2(y,z)yβ()Λ(,z)ϕ2(,z)y, (3.6)

    respectively. We note that the function ϕ2(y,z) might need to be assumed to satisfy ϕ2(,z)=0 in order to find a solution for the correction Vs/l20. Refer to Fouque et al. [15] for a detailed discussion on this requirement.

    The next proposition provides the required ODE problems that the leading-order term and the first-order corrections (0i+j1) have to satisfy. The case for the second-order corrections corresponding to i+j=2 is presented in Appendix A.

    Proposition 3.2. The leading-order term and the first-order corrections, Vs/lij(x,z), 0i+j1, in the asymptotic series (3.3) for the stop-loss option price Vs/l satisfy the ODE problems

    {¯LVs/l00(x,z)=0,λ<x<1,Vs/l00(λ,z)=λ,Vs/l00(1,z)=xVs/l00(1,z), (3.7)
    {¯LVs/l10(x,z)=(U3000(z)x33x3+U2000(z)x22x2)Vs/l00(x,z):=B10(x,z),λ<x<1,Vs/l10(λ,z)=0,Vs/l10(1,z)=xVs/l10(1,z), (3.8)

    and

    {¯LVs/l01(x,z)=(U1100(z)xx+U0100(z))zVS/L00(x,z):=B01(x,z),λ<x<1,Vs/l01(λ,z)=0,Vs/l01(1,z)=xVs/l01(1,z), (3.9)

    respectively, where Ukl00(z),(k,l){(3,0),(2,0),(1,1),(0,1)}, are given by

    U3000(z):=12ρxyβ()f(,z)ϕ2y(,z),U2000(z):=ρxyβ()f(,z)ϕ2y(,z)12β()Λ(,z)ϕ2y(,z),U1100(z):=g(z)ρxzf(,z),U0100(z):=g(z)Γ(,z), (3.10)

    respectively.

    Proof. Firstly, by substituting the asymptotic series (3.3) into the boundary conditions in (3.1), we obtain

    i,j=0ϵi/2δj/2Vs/lij(λ,y,z)=λ,i,j=0ϵi/2δj/2(Vs/lij(1,y,z)xVs/lij(1,y,z))=0 (3.11)

    which yields the desired boundary conditions in (3.7)–(3.9) directly.

    Next, by applying Proposition 3.1 to the O(1) terms in Eq (3.4), we obtain the Poisson equation

    L0Vs/l20+L2Vs/l00=0. (3.12)

    Then by Lemma 2.1 ¯LVs/l00=0 holds and thus the ODE in (3.7) is satisfied.

    Similarly, by Proposition 3.1 and Lemma 2.1, the terms of order ϵ in Eq (3.4) lead to

    ¯LVs/l10=L1Vs/l20. (3.13)

    On the other hand, from (3.12) and ¯LVs/l00=0, the term Vs/l20 satisfies

    L0Vs/l20=12(f2(y,z)f2(,z))x22x2Vs/l00. (3.14)

    Then the solution Vs/l20 is given by

    Vs/l20(x,y,z)=12ϕ2(y,z)x22x2Vs/l00(x,z)+Fs/l20(x,z) (3.15)

    for some function Fs/l20(x,z) independent of variable y, where ϕ2 is defined in (3.6). Thus Eq (3.13) becomes

    ¯LVs/l10=(U3000x33x3+U2000x22x2)Vs/l00

    and thus the ODE in (3.8) is satisfied.

    Again, by Proposition 3.1 and Lemma 2.1, the terms of order δ in Eq (3.4) yield ¯LVs/l01=M1Vs/l00. Since M1 is the same as (U1100(z)xx+U0100(z))z, the ODE in (3.9) holds.

    To obtain PDE problems for the price PR(s,s,y,z) of a Russian option, we first note that PR(s,s,y,z) satisfies the PDE (2.2) on the interval sf<s<s and the boundary conditions

    PR(s=sf,s,y,z)=s,PRs(s=s,s,y,z)=0,

    where sf(y,z) stands for a free boundary at which PR and PRs are continuous.

    If we use the change of variables x=s/s and PR(s,s,y,z)=sVR(x,y,z) again for dimensionality reduction, we obtain a PDE problem given by

    LVR(x,x,y,z)=0,xf<x<1,VR(xf,y,z)=1,VRx(xf,y,z)=0,VR(1,y,z)=VRx(1,y,z), (3.16)

    where xf(y,z) is the free boundary corresponding to sf(y,z).

    We are interested in the option price VR and the free boundary xf given in the following form:

    VR(x,y,z)=i,j=0ϵi/2δj/2VRij(x,y,z),xf(y,z)=i,j=0ϵi/2δj/2xfij(y,z). (3.17)

    Proposition 3.3. The terms VRij with i=0,1 and j=0,1,2 in the asymptotic series (3.17) for the Russian option price VR are independent of y; VRij(x,y,z)=VRij(x,z).

    Proof. The proof of this proposition is similar to the proof of Proposition 3.1 for the stop-loss option case since the proof does not depend on the boundary conditions. So, we omit the proof.

    The PDE form for the the Russian option price is the same as the one for the stop-loss option price. The difference between them lies in the boundary conditions and the existence of a free boundary. Thus, in the following proposition about the leading-order term and the first-order corrections (0i+j1), we obtain the same ODEs as in the stop-loss option case but with different boundary conditions and the appearance of a free boundary. The required ODE problems for the second-order corrections corresponding to i+j=2 are given in Appendix B.

    Proposition 3.4. The leading-order term and the first-order corrections, VRij, 0i+j1, in the asymptotic series (3.17) for the Russian option price VR satisfy the ODE problems

    {¯LVR00(x,z)=0,xf00(z)<x<1,VR00(1,z)=xVR00(1,z),VR00(xf00(z),z)=1,xVR00(xf00(z),z)=0, (3.18)
    {¯LVR10(x,z)=(U3000(z)x33x3+U2000(z)x22x2)VR00(x,z),xf00(z)<x<1,VR10(1,z)=xVR10(1,z),VR10(xf00(z),z)=0,xf10(z)=xVR10(xf00(z),z)2x2VR00(xf00(z),z), (3.19)

    and

    {¯LVR01(x,z)=(U1100(z)xx+U0100(z))zVR00(x,z),xf00(z)<x<1,VR01(1,z)=xVR01(1,z),VR01(xf00(z),z)=0,xf01(z)=xVR01(xf00(z),z)2x2VR00(xf00(z),z), (3.20)

    respectively, where Ukl00(z),(k,l){(3,0),(2,0),(1,1),(0,1)}, are given by (3.10).

    Proof. Substituting the series (3.17) into the boundary conditions in (3.16) and applying the Taylor series of the functions VR(x,y,z) and xVR(x,y,z) at xf00(y,z), we have

    i,j=0ϵi/2δj/2(VRij(1,y,z)xVRij(1,y,z))=0,k=01k!(i,j=0ϵi/2δj/2kxkVRij(xf00(y,z),y,z))(i,j=1ϵi/2δj/2xfij(y,z))k=1,k=01k!(i,j=0ϵi/2δj/2k+1xk+1VRij(xf00(y,z),y,z))(i,j=1ϵi/2δj/2xfij(y,z))k=0. (3.21)

    Since VRij is independent of variable y for every (i,j){(0,0),(1,0),(0,1)}, xfij is also independent of y for those (i,j). Rearranging the expansion (3.21) in terms of ϵi/2δj/2, we can obtain the desired boundary conditions as stated in (3.18)–(3.20).

    From Propositions 3.2 and 3.4, we find that the leading-order terms Vs/l00 and VR00 are the Black-Scholes option prices with volatility σ(z) for stop-loss and Russian options, respectively. We obtain the solutions for the first-order corrections in the following section.

    In this section, we solve the problems for Vs/lij and VRij, 0i+j1, derived in Propositions 3.2 and 3.4, respectively. The solutions for the second-order corrections corresponding to i+j=2 are given in Appendix A (stop-loss option) and Appendix B (Russian option).

    Both stop-loss and Russian options share the same PDE structure even if the boundary conditions are different. So, there are identical parts of the PDE solutions for both options. For convenience, we eliminate the superscript 's/l' and 'R' of Vs/l and VR and use notation V for the general PDE solution. In this section, we derive the concrete forms of Vij up to i+j=1.

    First, we have the following proposition.

    Proposition 4.1. The leading-order term and the first-order corrections, Vij(x,z), 0i+j1, can be expressed as

    Vij(x,z)=2k=1(i+2jζ=0Aζij,k(z)(lnx)ζ)xηk(z),(i,j){(0,0),(1,0),(0,1)}, (4.1)

    where A000,k(z) and ηk(z) are defined by

    stop-loss:A000,k(z)=(1ηl(z))λ(1ηk(z))ληl(z)(1ηl(z))ληk(z)(k,l{1,2},kl),Russian:A000,k(z)=ηl(z)(ηl(z)ηk(z))(xf00(z))ηk(z)(k,l{1,2},kl),xf00(z):=(η1(z)(1η2(z))η2(z)(1η1(z)))1η2(z)η1(z),ηk(z):=12rqσ2(z)+(1)k1(12rqσ2(z))2+2rσ2(z)(k{1,2}), (4.2)

    respectively, and Aζij,k(z), ζ=0,,i+2j and i+j=1, are some functions of z to be determined.

    Proof. Since the leading-order term V00 is the option price under the Black-Scholes model with volatility σ(z), the known result in Wilmott et al. [32] says

    V00(x,z)=2k=1A000,k(z)xηk(z), (4.3)

    where A000,k(z) and ηk(z) are given by (4.2). Also, the leading-order term, xf00(z), of the free boundary for a Russian option is given by the expression in (4.1).

    The first-order corrections Vij(x,z), i+j=1, depend on the corresponding inhomogeneous terms Bij(x,z) defined in (3.8) and (3.9). Substituting (4.3) into those Bij(x,z), we can obtain

    B10(x,z)=2k=1B010,k(z)xηk(z),B010,k(z):=(U30002ω=0(ηk(z)ω)+U20001ω=0(ηk(z)ω))A000,k(z),B01(x,z)=2k=1(1ζ=0Bζ01,k(z)(lnx)ζ)xηk(z),B001,k(z):=(U1100(ηk(z)z+ηk(z)z)+U0100z)A000,k(z),B101,k(z):=(U1100ηk(z)+U0100)ηk(z)zA000,k(z), (4.4)

    where Ukl00(z),(k,l){(3,0),(2,0),(1,1),(0,1)}, are given by (3.10). Since ¯L defined by (3.5) is related to the well-known Cauchy-Euler equation of order 2, one can take the following expressions for V10(x,z) and V01(x,z), hinted at by (4.4), respectively:

    V10(x,z)=2k=1(1ζ=0Aζ10,k(z)(lnx)ζ)xηk(z),V01(x,z)=2k=1(2ζ=0Aζ01,k(z)(lnx)ζ)xηk(z) (4.5)

    for some functions Aζ10,k(z), ζ=0,1, and Aζ01,k(z), ζ=0,1,2. So, Proposition 4.1 is proved.

    In the following argument, we obtain the concrete forms of Aζ10,k(z), ζ=0,1, and Aζ01,k(z), ζ=0,1,2. Among those terms, A110,k(z), A101,k(z) and A201,k(z) in (4.1) are commonly shared regardless of stop-loss or Russian. They are first solved in the following proposition.

    Proposition 4.2. The terms A110,k(z) and Aζ01,k(z), ζ=1,2, in (4.1) are given by

    A110,k(z)=B010,k(z)12σ2(z)(2ηk(z)1)+(rq),A101,k(z)=B001,k(z)σ2(z)A201,k(z)12σ2(z)(2ηk(z)1)+(rq),A201,k(z)=B101,k(z)2(12σ2(z)(2ηk(z)1)+(rq)), (4.6)

    respectively, where B010,k(z), B001,k(z) and B101,k(z) are defined in (4.4).

    Proof. Substituting

    Vij(x,z)=2k=1(i+2jζ=1Aζ01,k(z)(lnx)ζ)xηk(z),(i,j){(1,0),(0,1)}

    into the ODEs in (3.8) and (3.9), we can get

    2k=1(12σ2(z)(2ηk(z)1)+(rq))A110,k(z)=2k=1B010,k(x,z),2k=1[σ2(z)A201,k(z)+(12σ2(z)(2ηk(z)1)+(rq))(A101,k(z)+2A201,k(z)lnx)]=2k=1(1ζ=0Bζ01,k(z)(lnx)ζ).

    Then, by simple calculation, we can obtain (4.6).

    Next, since the terms A0ij,k(z), (i,j){(1,0),(0,1)}, in (4.1) depend on the boundary conditions in (3.8) and (3.9) (stop-loss option) or (3.19) and (3.20) (Russian option), we obtain the particular solutions of them as shown below in Propositions 4.3 and 4.4, respectively.

    In this section, we find solutions for A010,k(z) and A001,k(z) which are decided by the boundary conditions in (3.8) and (3.9) corresponding to a stop-loss option.

    Proposition 4.3. The terms A010,k(z) and A001,k(z), k{1,2}, in (4.1) for a stop-loss option are given by

    A010,k(z)=ληl(z)2ω=1A110,ω(z)+(1ηl(z))lnλ2ω=1ληω(z)A110,ω(z)(1ηk(z))ληl(z)(1ηl(z))ληk(z),A001,k(z)=ληl(z)2ω=1A101,ω(z)+(1ηl(z))2ω=1(ληω(z)2ζ=1Aζ01,ω(z)(lnλ)ζ)(1ηk(z))ληl(z)(1ηl(z))ληk(z), (4.7)

    respectively, where k,l1,2 with kl and A110,ω(z) and Aζ01,ω(z)(ζ=1,2) are given in Proposition 4.2.

    Proof. Substituting

    Vij(x,z)=2k=1(i+2jζ=0Aζ01,k(z)(lnx)ζ)xηk(z),(i,j){(1,0),(0,1)}

    into the boundary conditions in (3.8) and (3.9), we obtain the following simultaneous equations:

    {2k=1A0ij,k(z)=2k=1(ηk(z)A0ij,k(z)+A1ij,k(z)),2k=1(ληk(z)i+2jζ=0Aζij,k(z)(lnλ)ζ)=0,(i,j){(1,0),(0,1)}.

    Then, by simple calculation, we can obtain the solutions (4.7).

    In this section, we derive solutions for A010,k(z) and A001,k(z) which are decided by the boundary conditions in (3.19) and (3.20) corresponding to a Russian option.

    Proposition 4.4. The terms A010,k(z) and A001,k(z), k{1,2}, in (4.1) for a Russian option are given by

    A0ij,k(z)=(xf00(z))ηl(z)2ω=1A1ij,ω(z)+(1ηl(z))2ω=1[(xf00(z))ηω(z)i+2jζ=1Aζij,ω(z)(lnxf00(z))ζ](1ηk(z))(xf00(z))ηl(z)(1ηl(z))(xf00(z))ηk(z), (4.8)

    where k,l1,2 with kl and the corresponding free boundaries xf10(z) and xf01(z) are given by

    xfij(z)=2k=1[(xf00(z))ηk(z)1(i+2jζ=1ζAζij,k(z)(lnxf00(z))ζ1+ηk(z)i+2jζ=0Aζij,k(z)(lnxf00(z))ζ)]2k=1[(xf00(z))ηk(z)2ηk(z)(ηk(z)1)A000,k(z)], (4.9)

    where (i,j){(1,0),(0,1)}) and Aζij,k(z), k=1,2,ζ=1,....i+2j, are given in Proposition 4.2.

    Proof. Substituting

    Vij(x,z)=2k=1(i+2jζ=0Aζ01,k(z)(lnx)ζ)xηk(z),(i,j){(1,0),(0,1)},

    into the boundary conditions in (3.19) and (3.20), we obtain the following simultaneous equations:

    {2k=1A0ij,k(z)=2k=1(ηk(z)A0ij,k(z)+A1ij,k(z)),2k=1((xf00(z))ηk(z)i+2jζ=0Aζij,k(z)(lnxf00(z))ζ)=0,(i,j){(1,0),(0,1)}.

    Then we can obtain the solutions (4.8) by calculating this directly.

    Next, we compute the free boundary xfij(z) from the boundary conditions in (3.18)–(3.20). Substituting the solutions V10(x,z) and V01(x,z) given by (4.5) together with (4.6) and (4.8) into the free boundary conditions in (3.18)–(3.20), we can obtain the desired free boundary result (4.9).

    Remark. In the above argument, we have obtained an approximation, ˜V:=V00+ϵV10+δV01, of the option price V (=Vs/l or VR). Using the same analysis as in Fouque et al. [14], one can obtain a theoretical error of the approximation. Instead of repeating the argument here, however, we demonstrate the accuracy of the approximation numerically in the next section.

    In this section, we perform some numerical experiments for stop-loss and Russian options under the multiscale stochastic volatility model (2.1). Since the real market data of these options are not available, we use the parameter values used in Fouque et al. [15] and Fitt et al. [11].

    We approximate the prices, Vs/l and VR, of stop-loss and Russian options by

    Vs/lVs/l00+˜Vs/l1,VRVR00+˜VR1, (5.1)

    respectively, where Vs/l00 and VR00 are the Black-Scholes option prices with volatility σ(z) and ˜Vs/l1 and ˜VR1 are the first order corrections defined by

    ˜Vs/l1:=˜Vsl10+˜Vsl01=ϵVsl10+δVsl01,˜VR1:=˜VR10+˜VR01=ϵVR10+δVR01, (5.2)

    where Vslij and VRij are the corresponding terms in the series (3.3) and (3.17), respectively. We define terms Ukl,ϵ00(z), Ukl,δ00(z), ˜Bζij(z) and ˜Aζij(z) by

    Ukl,ϵ00(z):=ϵUkl00(z),(k,l){(3,0),(2,0)},Ukl,δ00(z):=δσ(z)Ukl00(z),(k,l){(1,1),(1,0)},˜Bζij,k(z):=ϵi/2δj/2Bζij,k(z),˜Aζij,k(z):=ϵi/2δj/2Aζij,k(z), (5.3)

    where the group parameters Ukl00(z), (k,l){(3,0),(2,0),(1,1),(1,0)}, are defined by (3.10) and the functions Bζij,k(z) and Aζij,k(z) are defined by (4.4) and (4.6), respectively. Moreover, for simplicity, we use the unified notations Uϵ00 and Uδ00 defined by Uδ00:=U11,δ00=U10,δ00 and Uϵ00:=U30,ϵ00=U20,ϵ00 when U11,δ00=U10,δ00 and U30,ϵ00=U20,ϵ00.

    For a numerical experiment, we apply the change rule z=σ(z)σ to the functions in (4.4) and use Ukl,ϵ00 and Ukl,δ00 defined in (5.3). Then the functions in (4.4) become

    ˜B010,k(σ):=ϵB010,k(σ)=(U30,ϵ002ω=0(ηk(σ)ω)+U20,ϵ00ηk(σ)(ηk(σ)1))A000,k(σ),˜B001,k(σ):=δB001,k(σ)=(U11,δ00(ηkσ(σ)+ηk(σ)σ)+U01,δ00σ)A000,k(σ),˜B101,k(σ):=δB101,k(σ)=(U11,δ00ηk(σ)ηkσ(σ)+U01,δ00ηkσ(σ))A000,k(σ), (5.4)

    respectively. On the other hand, from ˜Aζij,k(σ) defined in (5.3) and (4.5), we have

    ˜V10:=ϵV10(x,σ)=2k=1(1ζ=0˜Aζ10,k(σ)(lnx)ζ)xηk(σ),˜V01:=δV01(x,σ)=2k=1(2ζ=0˜Aζ01,k(σ)(lnx)ζ)xηk(σ), (5.5)

    respectively. Therefore, the functions ˜Vslij and ˜VRij, (i,j){(1,0),(0,1)}, in (5.5) constitute the correction terms in (5.2) and they are related to ˜Aζij,k which depends on the choice of a stop-loss or Russian option.

    Figures 1 and 2 show the effect of the multiscale stochastic volatility on the Black-Scholes option prices. Depending on each parameter set of the values of Uϵ00 and Uδ00, the option prices affected by the multiscale stochastic volatility (2.1) give a different graphical representation from the Black-Scholes option prices.

    Figure 1.  Stop-loss option prices under the Black-Scholes (BS) model and the multiscale stochastic volatility (SV) model for various levels of Uϵ00 and Uδ00; r=0.08.
    Figure 2.  Russian option prices and free boundaries under the Black-Scholes (BS) model and the multiscale stochastic volatility (SV) model for various levels of Uϵ00 and Uδ00; q=0.02, σ=0.3.

    Figure 1 demonstrates how flexible the multiscale stochastic volatility model is compared to the Black-Scholes model for a stop-loss option. It shows the dependence of the option price on the variable x(=s/s) (the underlying price-the maximum underlying price ratio) and the parameter λ (the rebate level) for different values of the group parameters Uϵ00 and Uδ00, respectively. Depending on the values of the group parameters, the price of a stop-loss option can be over-priced or under-priced. There is no fixed direction of movement. This suggests that the multiscale stochastic volatility model offers a greater degree of flexibility in the way that it can capture the diverse volatile nature of the stock markets.

    In Figure 2, we observe the correction effect on the Black-Scholes price and the free boundary of a Russian option given by the multiscale stochastic volatility model. The multiscale stochastic volatility model offers a greater degree of flexibility in a similar manner to the stop-loss option case. The free boundaries also show a behavior sensitive to the group parameters Uϵ00 and Uδ00.

    Tables 1 and 2 show some comparison results between the Monte-Carlo simulation outcomes Ps/lMC and PRMC and our analytic prices Ps/l and PR, respectively, for each different value of the initial asset price s. For the comparison, we calculate the absolute difference and relative difference values, where the absolute difference is defined by the absolute value of the difference between the option price obtained from the Monte-Carlo simulation and the option price from the approximation formula (5.1), and the relative difference is defined by the ratio of the absolute difference over the price obtained from the Monte-Carlo simulation. Here, the repeated sampling number is set to be 100,000 and the discretized time step dt to be 11500 in Monte-Carlo simulation. The common parameters and functions used in both stop-loss and Russian options are as follows: r=0.1, q=0.05, σ=0.3, s=105, ϵ=0.01, δ=0.001, U30,ϵ00=0.007, U20,ϵ00=0.002, U11,δ00=0.007 and U10,δ00=0.002. We test the multiscale stochastic volatility model by comparing it with Monte-Carlo simulation method because there is no real market data in the case of perpetual American type of options. Here, we brought the values of parameters and the set of functions used in Fouque and Han [12], one of existing research works about the Monte-Carlo simulation for the multiscale stochastic volatility model. They are f(y,z)=ey+z, α(y)=m1y, β(y)=ν12, c(z)=m2z, g(z)=ν22, m1=0.8, m2=0.8, ν1=0.5, ν2=0.8 and Λ=Γ=0.

    Table 1.  Comparison between the approximation formula and the Monte-Carlo simulation for stop-loss option prices against a variety of s; absolute difference =PslPslMC, relative difference =PslPslMCPslMC, y=0.93, z=0.73 and λ=0.5.
    s Ps/l Ps/lMC absolute difference relative difference
    80 52.6759 52.2988 0.3771 0.0072
    85 54.5355 54.2620 0.2735 0.0050
    90 56.7345 56.4162 0.3184 0.0056
    95 59.2284 58.8681 0.3603 0.0061
    100 61.9821 61.5982 0.3839 0.0062
    105 64.9677 64.5273 0.4404 0.0068

     | Show Table
    DownLoad: CSV
    Table 2.  Comparison between the approximation formula and the Monte-Carlo simulation for Russian option prices against a variety of s; absolute difference =PRPRMC, relative difference =PRPRMCPRMC, y=0.745 and z=0.347.
    s PR PRMC absolute difference relative difference
    80 136.9703 137.8832 0.9130 0.0067
    85 145.1347 145.4095 0.2748 0.0019
    90 153.3948 154.1971 0.8023 0.0052
    95 161.7382 162.2437 0.5055 0.0031
    100 170.1538 169.8807 0.2731 0.0016
    105 178.6319 178.4759 0.1560 0.0009

     | Show Table
    DownLoad: CSV

    Table 1 provides a comparison result between the Monte-Carlo simulation and the approximation (5.1) for the stop-loss option prices. For each different value of the initial asset price s, the absolute difference is within '1' and the relative difference is within '0.01'. They are quite similar to each other. Adding higher order terms to (5.1) would bring more similarity between the two results.

    Table 2 shows a comparison result between the Monte-Carlo simulation and the approximation (5.1) for the Russian option prices. In this case, we used the method suggested by Basso and Pianca [2] for obtaining the Monte-Carlo simulation results. The absolute difference is within '1' and the relative difference is within '0.01' again.

    Tables 3 and 4 provide a comparison result between the Monte-Carlo simulation and the first order approximation (5.1) for each different value of rq. The absolute differences are within '2.5' and '2.8' and the relative differences are within '0.07' and "0.03', respectively, for the stop-loss and Russian option prices.

    Table 3.  Comparison between the approximation formula and the Monte-Carlo simulation for stop-loss option prices against a variety of rq; absolute difference =PslPslMC, relative difference =PslPslMCPslMC, y=0.42, z=0.42 and λ=0.5.
    rq Psl PslMC absolute difference relative difference
    0.1 36.5508 38.9692 2.4184 0.0621
    0.12 38.4595 39.6875 1.2281 0.0309
    0.14 40.5777 40.4344 0.1432 0.0035
    0.16 42.8733 41.8477 1.0256 0.0245
    0.18 45.2658 43.3876 1.8782 0.0433
    0.2 47.549 45.649 1.9001 0.0416

     | Show Table
    DownLoad: CSV
    Table 4.  Comparison between the approximation formula and the Monte-Carlo simulation for Russian option prices against a variety of rq; absolute difference =PRPRMC, relative difference =PRPRMCPRMC, y=0.73 and z=0.33.
    rq PR PRMC absolute difference relative difference
    0.1 111.5176 114.0305 2.5129 0.0220
    0.12 113.1129 115.1552 2.0423 0.0177
    0.14 115.0676 116.5339 1.4663 0.0126
    0.16 117.4950 118.0136 0.5187 0.0044
    0.18 120.5682 120.0946 0.4736 0.0039
    0.2 124.5698 121.8557 2.7141 0.0223

     | Show Table
    DownLoad: CSV

    Tables 5 and 6 demonstrate a comparison result for the Monte-Carlo simulation and the approximation (5.1) for each different value of s. The absolute differences are within '0.5' and '2.8' for the stop-loss and Russian option prices, respectively. Moreover, the relative differences are within '0.009' and '0.02' for the stop-loss and Russian option prices, respectively.

    Table 5.  Comparison between the approximation formula and the Monte-Carlo simulation for stop-loss option prices against a variety of s; absolute difference =PslPslMC, relative difference =PslPslMCPslMC, y=0.93, z=0.73 and λ=0.5.
    s Psl PslMC absolute difference relative difference
    85 49.6338 49.4496 0.1842 0.0037
    90 50.0286 49.6868 0.3418 0.0069
    95 50.6724 50.4782 0.1942 0.0038
    100 51.5569 51.1381 0.4188 0.0082
    105 52.6759 52.2889 0.3870 0.0074

     | Show Table
    DownLoad: CSV
    Table 6.  Comparison between the approximation formula and the Monte-Carlo simulation for Russian option prices against a variety of s; absolute difference =PRPRMC, relative difference =PRPRMCPRMC, y=0.74 and z=0.34.
    s PR PRMC absolute difference relative difference
    85 136.1358 135.3692 0.7665 0.0057
    90 136.2416 136.887 0.6454 0.0047
    95 136.4171 137.5976 1.1805 0.0086
    100 136.6607 137.2971 0.6364 0.0046
    105 136.9703 139.7656 2.7953 0.02

     | Show Table
    DownLoad: CSV

    Figure 3 presents a comparison of the numerical values of stop-loss and Russian options under the multiscale stochastic volatility (MSV) model with those under the Black-Scholes (BS) model, the CEV model and the SEV (stochastic elasticity of variance) model of Kim et al. [22]. We could choose the popular Heston and 3/2 models instead of these models but our model is already a generalization of the (rescaled) Heston and 3/2 models. If f(y,z)=y1/2, α(y)=θy, β(y)=ξy1/2 and c(z)=g(z)=0, then our model is reduced to the Heston model. If f(y,z)=y1/2, α(y)=y(θy), β(y)=ξy3/2 and c(z)=g(z)=0, then it becomes the 3/2 model. Here, θ and ξ are constants. So, we compare our model with the CEV and SEV models. The real market data of stop-loss and Russian options do not exist since those options are not traded in exchange. So, we generate 'rough data' by Monte-Carlo simulation. Here, for the security of scientific objectivity, we consider the Black-Scholes model for the underlying asset prices on behalf of market data. We use the pricing results in Lee and Kim [24] under the CEV and SEV models. They are obtained by the first-order approximation formulas in [27] and [23] for the CEV and SEV models, respectively. As shown in Figure 3, we find that the MSV model has quite good performance to fit the data in comparison with the CEV and SEV models that tend to produce almost similar performances to each other in the case of the first-order approximation. The numerical values of the fitting errors between the rough data and the option prices under the four models are shown in Table 7, which demonstrates that the MSV model clearly outperforms the other models.

    Figure 3.  Fitting of the BS, MSV, CEV and SEV models to rough data.
    Table 7.  Numerical error between the BS, MSV, CEV or SEV option prices and rough data; PdataPmodel:=x=80,82,....,100|Pdata(x)P(x)|2.
    The stop-loss option
    PsldataPslBS PsldataPslMSV PsldataPslCEV PsldataPslSEV
    6.5035 5.2410 6.5012 6.5012
    The Russian option
    PRdataPRBS PRdataPRMSV PRdataPRCEV PRdataPRSEV
    10.9108 7.8487 10.9105 10.9105

     | Show Table
    DownLoad: CSV

    In this paper, we have studied the pricing of stop-loss and Russian options, two perpetual American-style type of exotic options with a lookback provision, under a multiscale stochastic volatility framework. We obtain a closed-form expression for the approximate price of each option so that the pricing formula is useful for performance optimization since one can then easily compute the first and second order derivatives. Each formula is given as the Black-Scholes option price plus a correction term. The correction terms produced by the multiscale stochastic volatility provide the flexibility of the option prices and the free boundary (in the case of Russian options) so that the model can capture the real market behavior better. The pricing formulas are tested and verified via Monte-Carlo simulations. A possible direction of future research would be an extension of the current work under a stochastic volatility model with multiple time scales to stochastic volatility models with multiple dimensions such as the hybrid stochastic volatility and CEV model of Choi et al. [4], a stochastic volatility model with stochastic liquidity and regime switching of He and Lin [18] or with stochastic interest rates of He and Lin [19], and the hybrid stochastic elasticity of variance and stochastic volatility model of Escobar-Anel and Fan [10].

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    We thank three anonymous reviewers for providing insightful comments and suggestions to improve this paper.

    The research of J. H. Kim was supported by the National Research Foundation of Korea NRF2021R1A2C1004080.

    The authors declare no conflict of interest.

    Prof. Jeong-Hoon Kim is the Guest Editor of special issue "Multiscale and Multifactor Stochastic Volatility Models and Data Analysis with Applications in Finance" for AIMS Mathematics. Prof. Jeong-Hoon Kim was not involved in the editorial review and the decision to publish this article.

    In this section, we derive the ODE problems and their solutions for the second-order corrections Vs/lij with i+j=2 for a stop-loss option. We first note that the term Vs/l20 in (3.3) depends on y, whereas the other terms Vs/lij, 0i+j2 and (i,i)(2,0), are independent of y. To determine the value of Vs/l20(x,y,z), we need to assume that Vs/l20(λ,,z)=0 and xVs/l20(1,,z)=Vs/l20(1,,z) hold based on the stochastic terminal layer analysis of Fouque et al. [15].

    Proposition 6.1. Under the assumption ϕ2(,z)=0, the second-order corrections Vs/lij, i+j=2, in the asymptotic series (3.3) for a stop-loss option satisfy the following ODE problems, respectively. First, Vs/l20(x,y,z) is given by (3.15), i.e.,

    Vs/l20(x,y,z)=12ϕ2(y,z)x22x2Vs/l00(x,z)+Fs/l20(x,z),

    where Fs/l20(x,z) satisfies

    {¯LFs/l20(x,z)=(˜U4000(z)x44x4+˜W3000(z)x33x3+˜W2000(z)x22x2)Vs/l00(x,z)+(U3000(z)x33x3+U2000(z)x22x2)Vs/l10(x,z):=B20(x,z),λ<x<1,Fs/l20(λ,z)=0,Fs/l20(1,z)=xFs/l20(1,z). (6.1)

    Second, Vs/l11(x,z) and Vs/l02(x,z) satisfy

    {¯LVs/l11(x,z)=(˜U2100(z)x22x2+W1100(z)xx+W0100(z))Vs/l00(x,z)+(U1100(z)xx+U0100(z))zVs/l10(x,z)+(U3000(z)x33x3+U2000(z)x22x2)Vs/l01(x,z):=B11(x,z),λ<x<1,Vs/l11(λ,z)=0,Vs/l11(1,z)=xVs/l11(1,z), (6.2)

    and

    {¯LVs/l02(x,z)=(U1100(z)xx+U0100(z))zVs/l01(x,z)(c(z)z+12g2(z)2z2)Vs/l00(x,z):=B02(x,z),λ<x<1,Vs/l02(λ,z)=0,Vs/l02(1,z)=xVs/l02(1,z), (6.3)

    respectively, where

    U4000(z):=12ρ2xyβfξ1y,˜U4000(z):=U4000(z)+14ϕ2f2,W3000(z):=12ρxy(5ρxyβfξ1yβfξ2yβΛξ1y),˜W3000(z):=W3000(z)+ϕ2f2,W2000(z):=(2ρ2xyβfξ1yρxyβfξ2yρxyβΛξ1y+12βΛξ2y),˜W2000(z):=W2000(z)+12ϕ2f2,U2100(z):=g(z)ρxyρxzβfϕ1y,˜U2100(z):=U2100(z)+12g(z)ρyzβϕ2y,W1100(z):=g(z)(ρxyρxzβfϕ1yρxzβΛϕ1yρxyβfψy),W0100(z):=g(z)βΛψy.

    Proof. In the middle of the proof of Proposition 3.2, we have shown that Vs/l20 is given by (3.15) in which the function Fs/l20 has not been determined. We now derive an ODE problem for Fs/l20 as follows. If Lemma 2.1 is applied to the terms of order ϵ in (3.4), then we have L1Vs/l20+¯LVs/l00=0. The terms of order ϵ in (3.4) give L0Vs/l30+L1Vs/l20+L2Vs/l10=0. So, we obtain L0Vs/l30=(L1Vs/l20L1Vs/l20)(L2¯L)Vs/l10 whose general solution is given by

    Vs/l30(x,y,z)=(12ρxyξ1(y,z)x33x3+(ρxyξ1(y,z)12ξ2(y,z))x22x2)Vs/l00(x,z)12ϕ2(y,z)x22x2Vs/l10(x,z)+Fs/l30(x,z) (6.4)

    for some function Fs/l30 independent of y, where the functions ξ1 and ξ2 are defined in (3.6). From the terms of order ϵ in (3.4), we have

    L2Vs/l20=L1Vs/l30 (6.5)

    by Lemma 2.1. With the help of (3.15), the left side of Eq (6.5) becomes

    L2Vs/l20=12ϕ2x22x2(L2Vs/l00¯LVs/l00)+¯LFs/l20=14ϕ2f2(x44x4+4x33x3+2x22x2)Vs/l00+¯LFs/l20. (6.6)

    With the help of (6.4), the right side of Eq (6.5) becomes

    L1Vs/l30=(U4000x44x4+W3000x34x3+W2000x22x2)Vs/l00+(U3000x33x3+U2000x22x2)Vs/l10. (6.7)

    Putting (6.5)–(6.7) together, we obtain the ODE for Fs/l20 as in (6.1).

    Next, by applying Proposition 3.1 and Lemma 2.1 to the terms of order δ in (3.4), we can obtain L0Vs/l21+L2Vs/l01+M1Vs/l00=0 and ¯LVs/l01+M1Vs/l00=0. These two equations yield L0Vs/l21=((L2¯L)Vs/l01+(M1M1)Vs/l00) whose solution Vs/l21 is given by

    Vs/l21(x,y,z)=12ϕ2(y,z)x22x2Vs/l01(x,z)g(z)(ρxzϕ1(y,z)xxψ(y,z))zVs/l00(x,z)+Fs/l21(x,z) (6.8)

    for some function Fs/l21 independent of variable y. Then, by applying Lemma 2.1 to the terms of order ϵδ in (3.4) and using the result (6.8), we can obtain the ODE for Vs/l11 in (6.2).

    Moreover, the ODE ¯LVs/l02=(M1Vs/l01+M2Vs/l00) for Vs/l02 can be obtained from the terms of order δ in (3.4). Then (6.3) follows from the definitions of M2 in (2.3) and M1 in (3.2).

    On the other hand, the terms of order ϵi/2δj/2, (i,j){(1,1),(0,2)}, in (3.11) provide the boundary conditions in (6.2) and (6.3) for Vs/l11 and Vs/l02, respectively. For the boundary conditions of Fs/l(x,z), if we apply the assumptions Vs/l20(λ,,z)=0 and xVs/l20(1,,z)=Vs/l20(1,,z) to the terms of order ϵ in (3.11), then the boundary conditions in (6.1) are obtained.

    Proposition 6.2. The solutions of the ODE problems in Proposition 6.1 are given by

    Vs/l20(x,y,z)=12ϕ2(y,z)2k=1ηk(z)(ηk(z)1)A000,k(z)xηk(z)+2k=1(2ζ=0Aζ20,k(z)(lnx)ζ)xηk(z),Vs/lij(x,z)=2k=1(i+2jζ=0Aζij,k(z)(lnx)ζ)xηk(z),(i,j){(1,1),(0,2)}, (6.9)

    respectively, where the terms Aζij,k, (i,j){(1,0),(0,1),(2,0),(1,1),(0,2)} and ζ=1,,i+2j, are recursively given by

    Aζij,k(z)=Bζ1ij,k(z)ζ(ζ+1)2σ2(z)Aζ+1ij,k(z)ζ(12σ2(z)(2ηk(z)1)+(qr)),ζ=1,,i+2j1,Aζij,k(z)=Bζ1ij,k(z)ζ(12σ2(z)(2ηk(z)1)+(qr)),ζ=i+2j (6.10)

    and the terms A0ij,k, (i,j){(1,0),(0,1),(2,0),(1,1),(0,2)}, are given by

    A0ij,k(z)=ληl(z)2ω=1A1ij,ω(z)+(1ηl(z))2ω=1(ληω(z)i+2jζ=1Aζij,ω(z)(lnλ)ζ)(1ηk(z))ληl(z)(1ηl(z))ληk(z),k,l{1,2},kl. (6.11)

    Proof. Substituting (4.3) and (4.5) into Bij(x,z), (i,j){(2,0),(1,1),(0,2)}, defined in Proposition 6.1, we first have

    Bij(x,z)=2k=1(i+2j1ζ=0Bζij,k(z)(lnx)ζ)xηk(z), (6.12)

    where the explicit (long) representations of Bζij,k(z), ζ=0,1,,i+2j1, are given below in Appendix C.

    To solve the ODE problems (6.1)–(6.3), we consider the functions F20, V11 and V02 defined by

    F20(x,z)=2k=1(2ζ=0Aζ20,k(z)(lnx)ζ)xηk(z),Vij(x,z)=2k=1(i+2jζ=0Aζij,k(z)(lnx)ζ)xηk(z),(i,j)=(1,1),(0,2), (6.13)

    where, for legibility, the simplified notation without the superscript 's/l' has been used. Substituting (6.13) into the first equations in (6.1)–(6.3), we obtain Aζij,k(z), ζ=1,,i+2j, as follows:

    A1ij,k(z)=B0ij,k(z)σ2(z)A2ij,k(z)12σ2(z)(2ηk(z)1)+(qr),(i,j)=(2,0),(1,1),(0,2),A220,k(z)=B120,k(z)2(12σ2(z)(2ηk(z)1)+(qr)),A2ij,k(z)=B111,k(z)3σ2(z)A3ij,k(z)2(12σ2(z)(2ηk(z)1)+(qr)),(i,j)=(1,1),(0,2),A311,k(z)=B211,k(z)3(12σ2(z)(2ηk(z)1)+(qr)),A302,k(z)=B202,k(z)6σ2(z)A402,k(z)3(12σ2(z)(2ηk(z)1)+(qr)),A402,k(z)=B302,k(z)4(12σ2(z)(2ηk(z)1)+(qr)). (6.14)

    Then (6.10) follows from (4.6) and (6.14).

    Next, substituting (6.13) into the boundary conditions in (6.1)–(6.3), we obtain A0ij,k in (6.13) as follows:

    A0ij,k(z)=ληl(z)2ω=1A1ij,ω(z)+(1ηl(z))2ω=1(ληω(z)i+2jζ=1Aζij,ω(z)(lnλ)ζ)(1ηk(z))ληl(z)(1ηl(z))ληk(z),(i,j)=(2,0),(1,1),(0,2). (6.15)

    Then (6.11) follows from (4.7) and (6.15).

    In this section, we derive ODE problems and their solutions for the second-order corrections VRij, i+j=2 (Russian option). We first note that the terms of order ϵ in the first and third equations in (3.21) given by

    {VR20(1,y,z)xVR20(1,y,z),VR20(xf00(z),y,z)+xVR00(xf00(z),z)xf20(y,z)+xVR10(xf00(z),z)xf10(z)+122x2VR00(xf00(z),z)(xf10(z))2

    become

    {VR20(1,y,z)xVR20(1,y,z),VR20(xf00(z),y,z)122x2VR00(xf00(z),z)(xf10(z))2, (6.16)

    respectively, since xVR00(xf00(z),z)=0 from (3.18) and xVR10(xf00(z),z)=2x2VR00(xf00(z),z)xf10(z) from (3.19). In the following proposition, we assume that VR20(1,,z)=xVR20(1,,z) and VR20((xf00(z),,z)=122x2VR00(xf00(z),z)(xf10(z))2 hold.

    Proposition 6.3. Under the assumption ϕ2(,z)=0, the second-order corrections, VRij, i+j=2, in the asymptotic series (3.17) for a Russian option satisfy the following ODE problems. First, VR20 is given by

    VR20(x,y,z)=12ϕ2(y,z)x22x2VR00(x,z)+FR20(x,z), (6.17)

    where FR20 solves

    {¯LFR20(x,z)=[˜U4000(z)x44x4+˜W3000(z)x33x3+˜W2000(z)x22x2]VR00(x,z)+(U3000(z)x33x3+U2000(z)x22x2)VR10(x,z),xf00(z)<x<1,FR20(1,z)=xFR20(1,z),FR20(xf00(z),z)=122x2VR00(xf00(z),z)(xf10(z))2,xf20(y,z)=xVR20(xf00(z),y,z)+2x2VR10(xf00(z),z)xf10(z)+123x3VR00(xf00(z),z)(xf10(z))22x2VR00(xf00(z),z). (6.18)

    Second, VR11 and VR02 satisfy

    {¯LVR11(x,z)=(˜U2100(z)x22x2+W1100(z)xx+W0100(z))VR00(x,z)+(U1100(z)xx+U0100(z))zVR10(x,z)+(U3000(z)x33x3+U2000(z)x22x2)VR01(x,z),xf00(z)<x<1,VR11(1,z)=xVR11(1,z),VR11(xf00(z),z)=2x2VR00(xf00(z),z)xf10(z)xf01(z),xf11(z)=xVR11(xf00(z),z)+2x2VR10(xf00(z),z)xf01(z)+2x2VR01(xf00(z),z)xf10(z)+3x3VR00(xf00(z),z)xf10(z)xf01(z)2x2VR00(xf00(z),z) (6.19)

    and

    {¯LVR02(x,z)=(U1100(z)xx+U0100(z))zVR01(x,z)(c(z)z+12g2(z)2z2)VR00(x,z),xf00(z)<x<1,VR02(1,z)=xVR02(1,z),VR02(xf00(z),z)=122x2VR(xf00(z),z)(xf01(z))2,xf02(z)=xVR02(xf00(z),z)+2x2VR01(xf00(z),z)xf01(z)+123x3VR00(xf00(z),z)(xf01(z))22x2VR00(xf00(z),z), (6.20)

    respectively.

    Proof. In the cases of VR11 and VR02, one can show the y-independence of xf11 and xf02 and obtain the results (6.19) and (6.20) including the boundary conditions for VRij and the free boundary conditions for xfij by using the same argument as in the proof of Proposition 4.4. So, we don't repeat the proof here.

    Next, in the case of VR20, as shown in the process of obtaining the solution Vs/l20(x,y,z) of (3.15), it can be expressed as (6.17) for some function FR20(x,z) which is independent of y. As shown in Proposition 6.1 for a stop-loss option, FR20 satisfies the ODE in (6.18).

    It remains to find the boundary conditions for FR20 and the free boundary condition for xf20. Substituting (6.17) into the assumptions VR20(1,y,z)=xVR20(1,y,z) and VR20((xf00(z),y,z)=122x2VR00(xf00(z),z)(xf10(z))2, we obtain the boundary conditions FR20(1,z)=xFR20(1,z) and FR20(xf00(z),z)=122x2VR00(xf00(z),z)(xf10(z))2. Lastly, from the terms of order ϵ in the third equation of (3.21), one can get directly the free boundary condition for xf20(y,z) as in (6.18).

    Proposition 6.4. The solutions of the ODE problems in Proposition 6.3 are given by

    VR20(x,y,z)=12ϕ2(y,z)2k=1ηk(z)(ηk(z)1)A000,k(z)xηk(z)+2k=1(2ζ=0Aζ20,k(z)(lnx)ζ)xηk(z),VRij(x,z)=2k=1(i+2jζ=0Aζij,k(z)(lnx)ζ)xηk(z),(i,j){(1,1),(0,2},

    respectively, and the corresponding free boundaries xfij, (i,j){(2,0),(1,1),(0,2)}), are given by

    xfij(z)=2k=1[i+2jζ=1ζAζij,k(z)(lnxf00(z))ζ1+ηk(z)i+2jζ=0Aζij,k(z)(lnxf00(z))ζmax{0,i1}2ϕ2(y,z)η2k(z)(ηk(z)1)A000,k(z)](xf00(z))ηk(z)1+imax{i,j}(xf10(z))1{ij}(xf01(z))1{i=j}×2k=1[(2ηk(z)1)A110,k(z)+ηk(z)(ηk(z)1)1ζ=0Aζ10,k(z)(lnxf00(z))ζ](xf00(z))ηk(z)2+jmax{i,j}(xf10(z))1{i=j}(xf01(z))1{ij}×2k=1[2A201,k(z)+(2ηk(z)1)2ζ=1ζAζ01,k(z)(lnxf00(z))ζ1+ηk(z)(ηk(z)1)2ζ=0Aζ01,k(z)(lnxf00(z))ζ](xf00(z))ηk(z)2+1max{i,j}(xf10(z))i(xf01(z))j2k=12ω=0(ηk(z)ω)A000,k(z)(xf00(z))ηk(z)32k=1ηk(z)(ηk(z)1)A000,k(z)(xf00(z))ηk(z)2,

    respectively, where the (general) solutions Aζij,k, (i,j){(1,0),(0,1),(2,0),(1,1),(0,2)} and ζ=1,,i+2j, are the same as (6.10) and the (particular) solutions A0ij,k, (i,j){(2,0),(1,1),(0,2)} are given by

    A0ij,k(z)=(xf00(z))ηl(z)2ω=1A1ij,ω(z)+(1ηl(z))[2ω=1((xf00(z))ηω(z)i+2jζ=1Aζij,ω(z)(lnxf00(z))ζ)1max(i,j)(xf10(z))i(xf01(z))j×2ω=1((xf00(z))ηω(z)2ηω(z)(ηω(z)1)A000,ω(z))](1ηk(z))(xf00(z))ηl(z)(1ηl(z))(xf00(z))ηk(z).

    Proof. The derivation of the solutions Aζij,k, (i,j){(1,0),(0,1),(2,0),(1,1),(0,2)} and ζ=1,,i+2j, and the solutions A0ij,k, (i,j){(2,0),(1,1),(0,2)} are similar to the case of a stop-loss option in 9. Moreover, from the terms of order ϵi/2δj/2, (i,j){(2,0),(1,1),(0,2)}, in the third equation of (3.21), one can calculate directly the free boundary xfij, (i,j){(2,0),(1,1),(0,2)} to get the above result. So, we omit the detailed derivation here.

    In this section, we present the detailed expressions of the functions Bζij,k(z) in (6.12) for (i,j)=(2,0) with ζ=0,1, (i,j)=(1,1) with ζ=0,1,2 and (i,j)=(0,2) with ζ=0,1,2,3.

    (1) Bζ20,k(z) for ζ=0,1

    B020,k(z)=(˜U40003ω=0(ηk(z)ω)+˜W30002ω=0(ηk(z)ω)+˜W2000ηk(z)(ηk(z)1))A000,k(z)+U3000(2ω=0(ηk(z)ω)A010,k(z)+(3η2k(z)6ηk(z)+2)A110,k(z))+U2000(ηk(z)(ηk(z)1)A010,k(z)+(2ηk(z)1)A110,k(z)),B120,k(z)=(U30002ω=0(ηk(z)ω)+U2000ηk(z)(ηk(z)1))A110,k(z).

    (2) Bζ11,k(z) for ζ=0,1,2

    B011,k(z)=U3000(2ω=0(ηk(z)ω)A001,k(z)+(3η2k(z)6ηk(z)+2)A101,k(z)+6(ηk(z)1)A201,k(z))+U2000(ηk(z)(ηk(z)1)A001,k(z)+(2ηk(z)1)A101,k(z)+2A201,k(z))+˜U2100((2ηk(z)1)ηk(z)z+ηk(z)(ηk(z)1)z)A000,k(z)+W1100(ηk(z)z+ηk(z)z)A000,k(z)+W0100zA000,k(z)+U1100((ηk(z)z+ηk(z)z)A010,k(z)+zA110,k(z))+U0100zA010,k(z),B111,k(z)=U3000(2ω=0(ηk(z)ω)A101,k(z)+2(3η2k(z)6ηk(z)+2)A201,k(z))+U2000(ηk(z)(ηk(z)ω)A101,k(z)+2(2ηk(z)1)A201,k(z))+˜U2100ηk(z)(ηk(z)1)ηk(z)zA000,k(z)+W1100ηk(z)ηk(z)zA000,k(z)+W0100ηk(z)zA000,k(z)+U1100(ηk(z)ηk(z)zA010,k(z)+(2ηk(z)z+ηk(z)z)A110,k(z))+U0100(ηk(z)z+z)A010,k(z),B211,k(z)=(U30002ω=0(ηk(z)ω)+U2000ηk(z)(ηk(z)1))A201,k(z))+(U1100ηk(z)ηk(z)z+U0100ηk(z)z)A110,k(z).

    (3) Bζ02,k(z) for ζ=0,1,2,3

    B002,k(z)=U1100((ηk(z)z+ηk(z)z)A001,k(z)+zA101,k(z))+U0100zA001,k(z)(c(z)z+12g2(z)2z2)A000,k(z),B102,k(z)=U1100(ηk(z)ηk(z)zA001,k(z)+(ηk(z)z+2ηk(z)z)A101,k(z)+2zA201,k(z))+U0100(ηk(z)zA001,k(z)+zA101,k(z))(c(z)ηk(z)z+12g2(z)(2ηk(z)z2+2ηk(z)zz))A000,k(z),B202,k(z)=U1110(ηk(z)ηk(z)zA101,k(z)+(ηk(z)z+3ηk(z)z)A201,k(z))+U0110(ηk(z)zA101,k(z)+zA201,k(z))12g2(z)(ηk(z)z)2A000,k(z),B302,k(z)=(U1110ηk(z)ηk(z)z+U0110ηk(z)z)A201,k(z).


    [1] Y. Cherruault, M. Guerret, Parameters identification and optimal control in pharmacokinetics, Acta Appl. Math., 1 (1983), 105–120. https://doi.org/10.1007/BF00046831 doi: 10.1007/BF00046831
    [2] B. Some, Y. Cherruault, Optimization and optimal control of pharmacokinetics, Biomedicine & pharmacotherapy, 40 (1986), 183–187.
    [3] F. Angaroni, A. Graudenzi, M. Rossignolo, D. Maspero, T. Calarco, R. Piazza, et al., An optimal control framework for the automated design of personalized cancer treatments, Front. Bioeng. Biotechnol., 2020. https://doi.org/10.3389/fbioe.2020.00523 doi: 10.3389/fbioe.2020.00523
    [4] M. Leszczyński, U. Ledzewicz, H. Schättler, Optimal control for a mathematical model for chemotherapy with pharmacometrics, Math. Model. Nat. Phenom., 15 (2020), 23. https://doi.org/10.1051/mmnp/2020008 doi: 10.1051/mmnp/2020008
    [5] I. Irurzun-Arana, A. Janda, S. Ardanza-Trevijano, I. F. Trocóniz, Optimal dynamic control approach in a multi-objective therapeutic scenario: Application to drug delivery in the treatment of prostate cancer, PLoS Comput. Biol., 14 (2018), e1006087. https://doi.org/10.1371/journal.pcbi.1006087 doi: 10.1371/journal.pcbi.1006087
    [6] F. Bachmann, G. Koch, M. Pfister, G. Szinnai, J. Schropp, Optidose: Computing the individualized optimal drug dosing regimen using optimal control, J. Optim. Theory Appl., 189 (2021), 46–65. https://doi.org/10.1007/s10957-021-01819-w doi: 10.1007/s10957-021-01819-w
    [7] P. Sopasakis, P. Patrinos, H. Sarimveis, Robust model predictive control for optimal continuous drug administration, Comput. Methods Programs Biomed., 116 (2014), 193–204. https://doi.org/10.1016/j.cmpb.2014.06.003 doi: 10.1016/j.cmpb.2014.06.003
    [8] S. Lucia, M. Schliemann, R. Findeisen, E. Bullinger, A set-based optimal control approach for pharmacokinetic/pharmacodynamic drug dosage design, IFAC-PapersOnLine, 49 (2016), 797–802. https://doi.org/10.1016/j.ifacol.2016.07.286 doi: 10.1016/j.ifacol.2016.07.286
    [9] L. T. Aschepkov, D. V. Dolgy, T. Kim, R. P. Agarwal, Optimal Control, Springer, Cham, 2016.
    [10] H. P. Geering, Optimal Control with Engineering Applications, Springer, Berlin, 2007.
    [11] S. Lenhart, J. T. Workman, Optimal Control Applied to Biological Models, New York, 2007. https://doi.org/10.1201/9781420011418
    [12] S. E. Rosenbaum, Basic Pharmacokinetics and Pharmacodynamics: An Integrated Textbook and Computer Simulations, Wiley, 2016.
    [13] F. H. Clarke, Inequality constraints in the calculus of variations, Can. J. Math., 29 (1977), 528–540. https://doi.org/10.4153/CJM-1977-055-6 doi: 10.4153/CJM-1977-055-6
    [14] F. A. Valentine, The problem of lagrange with differential inequalities as added side conditions, in Traces and Emergence of Nonlinear Programming, (2014), 313–357. https://doi.org/10.1007/978-3-0348-0439-4_16
  • This article has been cited by:

    1. T. A. Averina, K. A. Rybakov, Rosenbrock-Type Methods for Solving Stochastic Differential Equations, 2024, 17, 1995-4239, 99, 10.1134/S1995423924020010
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2710) PDF downloads(135) Cited by(0)

Figures and Tables

Figures(9)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog