Optimal control problems with time delays: Two case studies in biomedicine

  • Received: 30 April 2017 Accepted: 18 March 2018 Published: 01 October 2018
  • MSC : Primary: 49K15, 49M37; Secondary: 92C50

  • There exists an extensive literature on delay differential models in biology and biomedicine, but only a few papers study such models in the framework of optimal control theory. In this paper, we consider optimal control problems with multiple time delays in state and control variables and present two applications in biomedicine. After discussing the necessary optimality conditions for delayed optimal control problems with control-state constraints, we propose discretization methods by which the delayed optimal control problem is transformed into a large-scale nonlinear programming problem. The first case study is concerned with the delay differential model in [21] describing the tumour-immune response to a chemo-immuno-therapy. Assuming L1-type objectives, which are linear in control, we obtain optimal controls of bang-bang type. In the second case study, we introduce a control variable in the delay differential model of Hepatitis B virus infection developed in [7]. For L1-type objectives we obtain extremal controls of bang-bang type.

    Citation: Laurenz Göllmann, Helmut Maurer. Optimal control problems with time delays: Two case studies in biomedicine[J]. Mathematical Biosciences and Engineering, 2018, 15(5): 1137-1154. doi: 10.3934/mbe.2018051

    Related Papers:

    [1] Cristiana J. Silva, Helmut Maurer, Delfim F. M. Torres . Optimal control of a Tuberculosis model with state and control delays. Mathematical Biosciences and Engineering, 2017, 14(1): 321-337. doi: 10.3934/mbe.2017021
    [2] Heping Ma, Hui Jian, Yu Shi . A sufficient maximum principle for backward stochastic systems with mixed delays. Mathematical Biosciences and Engineering, 2023, 20(12): 21211-21228. doi: 10.3934/mbe.2023938
    [3] Jerzy Klamka, Helmut Maurer, Andrzej Swierniak . Local controllability and optimal control for\newline a model of combined anticancer therapy with control delays. Mathematical Biosciences and Engineering, 2017, 14(1): 195-216. doi: 10.3934/mbe.2017013
    [4] Junyoung Jang, Kihoon Jang, Hee-Dae Kwon, Jeehyun Lee . Feedback control of an HBV model based on ensemble kalman filter and differential evolution. Mathematical Biosciences and Engineering, 2018, 15(3): 667-691. doi: 10.3934/mbe.2018030
    [5] Pensiri Yosyingyong, Ratchada Viriyapong . Global dynamics of multiple delays within-host model for a hepatitis B virus infection of hepatocytes with immune response and drug therapy. Mathematical Biosciences and Engineering, 2023, 20(4): 7349-7386. doi: 10.3934/mbe.2023319
    [6] Tingting Xue, Long Zhang, Xiaolin Fan . Dynamic modeling and analysis of Hepatitis B epidemic with general incidence. Mathematical Biosciences and Engineering, 2023, 20(6): 10883-10908. doi: 10.3934/mbe.2023483
    [7] Tahir Khan, Fathalla A. Rihan, Muhammad Ibrahim, Shuo Li, Atif M. Alamri, Salman A. AlQahtani . Modeling different infectious phases of hepatitis B with generalized saturated incidence: An analysis and control. Mathematical Biosciences and Engineering, 2024, 21(4): 5207-5226. doi: 10.3934/mbe.2024230
    [8] Ben Sheller, Domenico D'Alessandro . Analysis of a cancer dormancy model and control of immuno-therapy. Mathematical Biosciences and Engineering, 2015, 12(5): 1037-1053. doi: 10.3934/mbe.2015.12.1037
    [9] Tailei Zhang, Hui Li, Na Xie, Wenhui Fu, Kai Wang, Xiongjie Ding . Mathematical analysis and simulation of a Hepatitis B model with time delay: A case study for Xinjiang, China. Mathematical Biosciences and Engineering, 2020, 17(2): 1757-1775. doi: 10.3934/mbe.2020092
    [10] Amar Nath Chatterjee, Fahad Al Basir, Yasuhiro Takeuchi . Effect of DAA therapy in hepatitis C treatment — an impulsive control approach. Mathematical Biosciences and Engineering, 2021, 18(2): 1450-1464. doi: 10.3934/mbe.2021075
  • There exists an extensive literature on delay differential models in biology and biomedicine, but only a few papers study such models in the framework of optimal control theory. In this paper, we consider optimal control problems with multiple time delays in state and control variables and present two applications in biomedicine. After discussing the necessary optimality conditions for delayed optimal control problems with control-state constraints, we propose discretization methods by which the delayed optimal control problem is transformed into a large-scale nonlinear programming problem. The first case study is concerned with the delay differential model in [21] describing the tumour-immune response to a chemo-immuno-therapy. Assuming L1-type objectives, which are linear in control, we obtain optimal controls of bang-bang type. In the second case study, we introduce a control variable in the delay differential model of Hepatitis B virus infection developed in [7]. For L1-type objectives we obtain extremal controls of bang-bang type.


    1. Introduction

    Differential dynamic systems with time delays play an important role in the modeling of real-life phenomena in various fields of applications. There exists an extensive literature on delay differential models in biology and biomedicine; see e.g., [6,7,16,18,24,26,27]. Though there is a vast literature on optimal control problems with time delays in control and state variables, so far only a few papers have applied the framework of optimal control with delays to biomedicine; cf. the recent papers [15,21,23]. The aim of this paper is to present two case studies in biomedicine which illustrate the application of delayed optimal control problems and demonstrate that there exist efficient numerical techniques to solve such problems.

    In Section 2, we consider optimal control problems with multiple time delays in control and state variables The control process can be subject to mixed control-state constraints. We review the necessary optimality conditions that were derived in [10] in the form of a Pontryagin type Minimum Principle. It is assumed that the so-called commensurability assumption holds which requires that the time delays and the terminal time are integer multiples of a joint stepsize. This assumption also underlies the discretization and nonlinear programming techniques that are briefly reviewed in Section 3. In Section 4, we study the delay differential model for tumour-immune-response with chemo-immunotherapy in Rihan et al. [21]. Aside from the state delay in this model we introduce a time delay in the control variable representing the immune therapy. The delay accounts for the fact that the human immune system needs some time to respond to the immune therapy. In contrast to the L2-type objectives in [21] we consider L1-type objectives characterized by linearly appearing controls which seem to be more appropriate in the biological framework; cf. the remarks in [22]. The computations show that all controls are of bang-bang type. For the non-delayed problem, we can verify the second-order sufficient conditions in [19,17] and thus show that the computed solution provides a strict strong minimum.

    Section 5 considers the delay differential model in [7] describing the spread of Hepatitis B virus (HBV). The dynamical model exhibits a delay in the state variables. We introduce a control variable into this model and formulate an optimal control problem using L1-type objectives. Again, all controls are of bang-bang type. We show that the non-delayed bang-bang control provides a strict strong minimum, whereas the delayed controls are extremal solutions that satisfy the necessary conditions with high accuracy.


    2. Optimal control problems with multiple time-delays in state and control variables


    2.1. Problem statement

    Let x(t)Rn denote the state variable and u(t)Rm the control variable at time t[0,tf] with fixed terminal time tf>0. The time-delays in the state and control variables are given by a constant vector (τ1,,τd)Rd satisfying

    0=:τ0<τ1<<τd.

    Thus τ0 represents the non-delayed variables. In [9,10] we have studied the following optimal control problem with multiple time-delays and mixed control-state constraints (MDOCP): determine a pair of functions (x,u)W1,([0,tf],Rn)×L([0,tf],Rm) that minimize the functional in Mayer form

    J(x,u)=g(x(tf)) (1)

    subject to the delayed (retarded) differential equation, boundary conditions and mixed control-state inequality constraints

    ˙x(t)=f(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd)),a.e.  t[0,tf], (2)
    x(t)=x0(t),   t[τd,0], (3)
    u(t)=u0(t),   t[τd,0), (4)
    ψ(x(T))=0, (5)
    C(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd))0,a.e.  t[0,tf]. (6)

    The functions g:RnR,f:[0,tf]×R(d+1)n×R(d+1)mRn, ψ:RnRq(0qn), and C:[0,tf]×R(d+1)n×R(d+1)mRp are assumed to be continuously differentiable, while the functions x0:[τd,0]Rn,u0:[τd,0]Rm only need to be continuous.

    Without lack of generality we have assumed that the cost functional is given in Mayer form (1). It is well known that an objective in Bolza form,

    J(x,u)=g(x(tf))+tf0L(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd))dt,

    can be reduced to Mayer form by introducing an additional state variable xn+1 defined by

    ˙xn+1(t)=L(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd)),   xn+1(0)=0.

    Then we have to minimize the functional ˜J(x,xn+1,u)=g(x(tf))+xn+1(tf).

    In the following, we shall use the placeholder variables y0,y1,,yd for the delayed state variables and v0,v1,,vd for the delayed control variables. The delayed variables are defined by

    yδ(t)=x(tτδ),   vδ(t)=u(tτδ)  (δ=0,1,,d). (7)

    Note that we do not necessarily assume an equal number of state and control delays. The case of an unequal number of delays in state and control variables is included in this formulation as we admit that

    hyδ=0  or  hvδ=0,  h{f,C,L},  for some  δ{0,,d}.

    2.2. Minimum principle: First-order necessary conditions

    A Pontryagin-type minimum principle for problem (MDOCP) has been derived in [9,10]. The main result requires that all positive time delays τ1,,rd can be expressed as integer multiples of a sufficiently small positive constant (stepsize).

    Assumption 2.1 (Commensurability Condition). Assume that there exist a constant h>0 and integers k1,,kd,N with

    τδ=kδh(δ=1,,d)  and  tf=Nh. (8)

    In view of 0=τ0<τ1<<τd we have 0<k1<<kd. Then in analogy to the non-delayed case we define the Hamiltonian function by

    H(t,y0,,yd,v0,,vd,λ)=λf(t,y0,,yd,v0,,vd),   λRn, (9)

    where the adjoint variable λRn is a row vector. The augmented Hamiltonian function is defined by adjoining the mixed control-state constraint (6) to the Hamiltonian using a multiplier μRp (row vector):

    H(t,y0,,yd,v0,,vd,λ,μ)=H(t,y0,,yd,v0,,vd,λ)+μC(t,y0,,yd,v0,,vd). (10)

    For ease of notation we refrain from denoting an optimal pair

    (x,u)W1,([0,tf],Rn)×L([0,tf],Rm)

    by a hat or a similar symbol. We require the following regularity condition for the active control-state constraints.

    Assumption 2.2 (Regularity Condition). Let (x,u) be a locally optimal pair and let

    J0(t):={j{1,,p}|Cj(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd))=0}

    denote the set of active indices for the inequality constraints (6). Assume that the gradients

    Cj(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd))(v0,,vd),   jJ0(t), (11)

    are linearly independent.

    The following theorem summarizes the first-order necessary conditions for optimality for the control problem (MDOCP) [10].

    Theorem 2.3. (Minimum Principle for Optimal Control Problems with Multiple Time-Delays [10]): Let (x,u) be a locally optimal pair for (MDOCP) with delays 0=τ0<τ1<τd that satisfies the commensurability condition (8) and the regularity condition 2.2. Then there exist an adjoint (costate) function λW1,([0,tf],Rn), a number λ00, a multiplier function μL([0,tf],Rp) and a multiplier νRq, such that the following conditions hold for a.e. t[0,tf]:

    1. Advanced Adjoint Differential Equation:

    ˙λ(t)=dδ=0χ[0,tfτδ](t)Hyδ(t+τδ), (12)

    where Hyδ[t]=Hyδ(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd),λ(t),μ(t)) and χ[0,tfτδ] is the characteristic function of the interval [0,tfτδ].

    2. Transversality Condition:

    λ(tf)=λ0gx(x(tf))+νψx(x(tf)). (13)

    3. Minimum Condition for the Hamiltonian:

    dδ=0χ[0,tfτδ](t)H[t+τδ]H(t,,u,u(tτ1),,u(tτd),λ(t))   +d1δ=1χ[0,tfτδ](t)H(t+τδ,,u(t+τδτδ1),u,u(t+τδτδ+1),)   +χ[0,tfτd](t)H(t+τd,,u(t+τdτ1),,u(t+τdτd1),u,λ(t)) (14)

    for all uRm satisfying

    C(t,x(tτ0),,x(tτd),u(tτ0),,u(tτδ1),u,u(tτδ+1),,u(tτd))0  for  δ=0,,d,

    where H[t]=H(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd),λ(t)).

    4. Local Minimum Condition for the Augmented Hamiltonian Function:

    dδ=0χ[0,tfτδ](t)Hvδ[t+τδ]=0. (15)

    5. Non-negativity of Multiplier and Complementarity Condition: for t[0,tf],

    μ(t)0,μ(t)C(t,x(tτ0),,x(tτd,u(tτ0),,u(tτd))=0. (16)

    3. Numerical discretization methods

    Similar to the case of non-delayed differential equations, we can employ integration methods of Runge-Kutta type or multistep methods, e.g., the Euler method and trapezoidal rule, to discretize the delay differential equation

    ˙x(t)=f(t,x(tτ0),,x(tτd),u(tτ0),,u(tτd)).

    Any integration method based on an equidistant discretization scheme utilizes a uniform step size h>0. Due to the presence of time-delays it is crucial to match the delays τ1,,τd to the grid. This is ensured by the commensurability condition (8) in Assumption 2.1. For this purpose, let h>0 be a step size satisfying (8), i.e.

    τδ=kδh  (δ=0,,d),  tf=Nh,

    with integers 0=k0<k1<<kd and N. Note that this grid can be refined by use of any integer fraction of h, This defines an equidistant discretization mesh with grid points ti=ih for i=0,1,...,N.

    Let xiRn and uiRm denote approximations of x(ti) and u(ti) at the grid points ti for i=0,1,,N. For convenience, we shall use the abbreviations

    fi=f(ti,xi,xik1,,xikδ,ui,uik1,,uikδ).

    The initial value profiles x0() and u0() provide the values

    xi=x0(ih)  (i=0,..,kd),  ui=u0(ih)  (i=1,..,kd). (17)

    Since the focus in this paper is not on discussing various numerical methods, we present only two integration methods that can be easily implemented. The simplest method is the first order method of Euler which is defined by the recursion

    xi+1=xi+hfi,   i=0,1,,N1. (18)

    The trapezoidal rule is an implicit method of second order:

    xi+1=xi+12h(fi+fi+1),   i=0,1,,N1. (19)

    Then for the Euler method and the optimization variable

    z:=(u0,x1,u1,x2,...,uN1,xN)RN(m+n)

    we obtain the following nonlinear programming problem (NLP) with equality and inequality constraints:

    Minimize   J(z)=g(xN) (20)

    subject to

    xi+1=xi+hf(ti,xik0,,xikd,uik0,,uikd),   i=0,,N1, (21)
    C(ti,xik0,,xikd,uik0,,uikd)0,     i=0,,N1, (22)
    ψ(xN)=0, (23)

    and initial values (17). Using the trapezoidal method (19) we simply replace the equations (21) by the equations defined in (19).

    Let λ=(λ0,λ1,,λN1)RnN,λiRn(i=0,,N1), be the Lagrange multipliers for equations (21) and let μ=(μ0,μ1,,μN1)RpN,μiRp (i=0,,N1), be the multipliers for the inequality constraints (22) and νNRq be the multiplier for the boundary condition (23). In [9,10] we have discussed the Karush-Kuhn-Tucker (KKT) necessary optimality conditions for the (NLP) using the Euler scheme (18) and showed that the property of consistency holds. This means that the Lagrange multipliers provide approximations for the adjoint variable λ(t), the multiplier μ(t) and ν according to

    λ(ti)λiRn,   μ(ti)μi/hRp  (i=0,...,N1),   νNν. (24)

    This follows from the fact that the Lagrange multipliers λi satisfy the advanced adjoint equations using the same discretization scheme in a backward mode.

    To solve the optimization problem (NLP) in (20)-(22) numerically, we employ the Applied Modeling Programming Language (AMPL) developed by Fourer, Gay and Kernighan [8] which can be linked to the interior-point optimization solver IPOPT developed by Wächter et al. [28] or to the SQP solver WORHP by Büskens and Gerdts [4]. Every solver provides the Lagrange multipliers and therefore gives access to approximations of adjoint variables and multiplier functions for the control problem (MDOCP) according to (24). Thus we can test whether the numerical solution is an extremal solution which satisfies the necessary optimality conditions in Theorem 2.3.


    4. Optimal control of chemo-immuno-therapy


    4.1. Optimal control problem

    We consider the delay differential model in Rihan et al. [21] that proposes a chemo-immuno-therapy of cancer. The authors introduce a time delay only in the state variable and present a stability analysis of drug free steady states. We shall extend the model by including also a control delay in the control u2 of immune therapy. The delay accounts for the fact that the human immune system takes some time to respond to the immune therapy. The state and variables have the following meaning:

    E: concentration of effector cells (plasma B cells, producing antibodies).

    T: concentration of tumour cells.

    N: concentration of healthy cells.

    U: concentration of cytostatic agent for chemotherapy.

    u1: dose control for chemotherapy,

    u2: dose control for immune therapy of the effector cells.

    Denoting the state delay by τ1 and the control delay by τ2, the dynamical system is given by

    ˙E(t)=σ+(ρη+T(tτ1)μe)E(tτ1)T(tτ1)(δ+a1(1eU(t)))E(t)+u2(tτ2)s1,˙T(t)=(r2(1βT(t))nTE(t)c1N(t)a2(1eU(t)))T(t),˙N(t)=(r3(1β2N(t))c2T(t)a3(1eU(t)))N(t),˙U(t)=u1(t)d1U(t). (25)

    The initial values and initial functions for the delayed state and control variables are as follows:

    E(0)=E0=0.3,E(t)=E0τ1t0,T(0)=T0=300,T(t)=T0τ1t0,N(0)=N0=0.9,u2(t)=0τ2t<0.U(0)=U0=0.0. (26)

    We shall consider the control constraints

    0uk(t)uk,max   t[0,tf]  (k=1,2). (27)

    Let us denote the state and control variables by

    x=(E,T,N,U)R4,   u=(u1,u2)R2.

    For notational convenience, we simplify the notations (7) for the delayed state and control variables. In the context of the dynamical system (25) it is more convenient to consider the delayed state variables y1,y2 and control variable v2 defined by

    y1(t)=x1(tτ1)=E(tτ1),y2(t)=x2(tτ1)=T(tτ1),v2(t)=u2(tτ2). (28)

    With these notations the dynamical system (25) can be written as

    ˙x(t)=f(x(t),y1(t),y2(t),u(t),v2(t)). (29)

    Then the optimal control problem is as follows: determine a control function u=(u1,u2)L([0,tf],R2) that minimizes the objective functional

    Jp(x,u)=tf0(T(t)E(t)+B1(u1(t))p+B2(u2(t))p)dt    (p=1,2) (30)

    subject to the dynamic constraints (25), initial conditions (26) and control constraints (27). The objective functional (30) represents a trade-off between minimizing the tumour cells and the total doses of the cytotoxic and immunologic agents on one hand and maximizing the plasma cells on the other hand. The constants B1>0,B2>0 are appropriate weights which are listed in Table 1 together with the system parameters.

    Table 1. Parameters in the control problem of chemo-immunotherapy [21].
    Parameter Description Value
    tf final time 30 d (days)
    τ1 state delay 1.5 d
    τ2 control delay 3.0 d
    (uk,min,uk,max) control bounds (0,1) for k=1,2
    (a1,a2,a3) cell kill rate response (0.2,0.4,0.1)
    (β,β2) reciprocal carrying capacities of tumour and host cells (0.002,1.0)
    (c1,c2) scaling parameters (3×105,3×108)
    d1 drug decay rate 0.01
    δ immune cell death rate 0.2
    η steepness of immune response 0.3
    μe uninfected effector cell decrease rate 0.003611
    (σ,ρ) immune cell influx and decay rate resp. (0.2,0.2)
    (s1,r2,r3) cell growth rates (0.3,1.03,1.0)
    nT immune effector cell decrease rate 1.0
    (B1,B2) weights (5,10)
     | Show Table
    DownLoad: CSV

    Rihan et al. [21] consider only the L2-type functional J2(x,u) in (30) which is quadratic in the control variable u. L2-type functionals are often used in economics to describe, e.g., production costs, but are mostly not appropriate in a biological framework; cf. the remarks in [22]. The L1 functional J1(x,u) incorporates the total amount of drugs used as a penalty and thus appears to be more realistic. For that reason, we shall mainly focus on the functional J1(x,u) in the sequel.

    Now we apply the necessary optimality conditions in the form of a Minimum Principle as stated in Theorem 2.3. Denoting the adjoint variable by the row vector λ=(λE,λT,λN,λU)R4, the Hamiltonian for the objective J1(x,u) and the control system (29) is given by

    H(x,y1,y2,u,v2,λ)=TE+B1u1+B2u2+λf(x,y1,y2,u,v2). (31)

    According to Theorem 2.3 (1), the advanced adjoint equations are given by

    ˙λE(t)=HE[t]χ[0,tfτ1](t)Hy1[t+τ1],˙λT(t)=HT[t]χ[0,tfτ1](t)Hy2[t+τ1],˙λN(t)=HN[t],˙λU(t)=HU[t]. (32)

    We do not write out the adjoint variables explicitly, since the adjoint variables can be computed as Lagrange multipliers of the discretized control problem as explained in the preceding section. Due to the free terminal state, the transversality condition (13) is

    λ(tf)=(0,0,0,0). (33)

    The optimal control u(t) minimizes the sum of Hamiltonians in (14). Since both controls appear linearly in the Hamiltonian, the minimizing controls are determined by the switching functions

    ϕ1(t)=Hu1[t]=B1+λU(t),ϕ2(t)=Hu2(t)+χ[0,tfτ2](t)Hv2[t+τ2]=B2+χ[0,tfτ2](t)λE(t+τ2)s1, (34)

    according to the control law

    uk(t)={0,ifϕk(t)>0uk,max,ifϕk(t)<0singular,ifϕk(t)=0tIs[0,tf]},   k=1,2. (35)

    Singular controls will not be discussed further, since our computations only yield bang-bang controls. Due to the transversality condition λ(tf)=0 the switching functions satisfy ϕk(tf)=Bk>0 for k=1,2. Hence, the control law (35) shows that uk(t)=0 holds on a terminal interval [tk,tf] for k=1,2. Parameters for the subsequent computations are given in the Table 1.


    4.2. Optimal solution of the non-delayed control problem

    First, we present the solution for the non-delayed control problem with τ1=τ2=0 and the functional J1(x,u). Recall the upper control bounds u1,max=u2,max=1, the terminal time tf=30 (days) and the weights B1=5 and B2=10 from Table 1. Applying AMPL/IPOPT with N=3000 grid points and the trapezoidal rule (19) we find the following bang-bang controls uk(t) with only one switch at tk,

    uk(t)={1for0t<tk0fortkttf}   (k=1,2),   0<t1<t2<tf. (36)

    To obtain a refinement of the solution, we solve the Induced Optimization Problem (IOP) with the switching times t1 and t2 as optimization variables; cf. [17,19]). The arc-parametrization method [17] and the optimal control package NUDOCCCS due to Büskens [2] yield the following numerical results

    J1(x,u)=1399.02,t1=3.93031,t2=9.76562,E(tf)=0.640303,T(tf)=0.180726,N(tf)=0.904968,U(tf)=2.96962.

    The initial values of the adjoint variables are

    λE(0)=770.13,λT(0)=2.9980,λN(0)=0.027548,λU(0)=281.11.

    The non-delayed solution is shown in Figure 1. A common strategy in medical practise is the administration of a pulse therapy or a blockwise application of drugs. Such a strategy is promoted by the controls in Figure 1.

    Figure 1. Optimal solution of the non-delayed control problem with τ1=τ2=0 and weights B1=5,B2=10. Top row: (a) dose control u1(t) of chemotherapy, (b) effector cells E(t), (c) tumour cells T(t). Bottom row: (a) dose control u2(t) of immune therapy, (b) healthy cells N(t), (c) cytostatic agent U(t).

    Now we show that the second-order sufficient conditions in [19], Chapter 7, are satisfied for the bang-bang control (36). For that purpose, we have to check two further conditions. First, notice that the objective J1(x,u) becomes a function J1(t1,t2) of the two switching times t1,t2, if we assume the control structure (36). The Hessian of J1(t1,t2) is computed as the positive definite 2×2 matrix

    D2J1(t1,t2)=(19.16711.12011.12010.887).

    Furthermore, as can be seen in Figure 2, the following strict bang-bang property with respect to the Minimum Principle holds for k=1,2:

    ϕk(t)<0   0t<tk,   ˙ϕk(tk)>0,   ϕk(t)>0   tk<ttf. (37)
    Figure 2. Optimal controls uk(t) and switching functions ϕk(t),(k=1,2) in a neighborhood of the switching times tk illustrating the control-law (35) and the strict bang-bang property (37).

    Hence, the solution shown in Figure 1 provides a strict strong minimum.

    We briefly compare the solutions for the functionals J1(x,u) and J2(x,u). The controls u1 and u2 for the functional J2(x,u) are continuous, since the strict Legendre-Clebsch condition holds and the Hamiltonian has a unique minimum with respect to u1 and u2. Figure 3 displays a comparison of the controls u1 and u2 for both functionals. The state variables for the functional J2(x,u) are very similar to those shown in Figure 1 and thus are not displayed here. The functional value is J2(x,u)=1392.88 versus J1(x,u)=1399.02 and the final state is computed as

    E(tf)=0.615728,T(tf)=0.108124,N(tf)=0.903899,U(tf)=3.20922.
    Figure 3. Optimal controls u1(t) and u2(t) for functionals Jp(p,u),p=1,2, with weights B1=5,B2=10.

    4.3. Numerical solution of the delayed control problem

    We choose the state delay τ1=1.5 and the control delay τ2=3. To obtain a rather precise reference solution, we apply AMPL/IPOPT with N=6000 grid points and tolerance tol=108. As in the non-delayed case we obtain a bang-bang control u(t)=(u1(t),u2(t)), where each uk(t) has only one switch at tk:

    uk(t)={1for0t<tk0fortkttf}  (k=1,2),  0<t1<t2<tf. (38)

    We obtain the numerical results

    J1(x,u)=2126.69,t1=4.692,t2=10.42,E(tf)=0.661258,T(tf)=0.136262,N(tf)=0.902747,U(tf)=3.55546.

    The initial values of the adjoint variables are

    λE(0)=485.41,λT(0)=2.2403,λN(0)=0.022090,λU(0)=248.50.

    Using the Euler method (18) with the same number N=6000 grid points, the numerical results are less accurate by two decimals. The control and state trajectories are shown in Figure 4. Figure 5 displays the controls and the switching functions in a neighborhood of the switching times. The zoom into the controls confirms that the control law (35) is precisely satisfied and that the strict bang-bang property (37) holds as well for the delayed solution. Unfortunately, we can not check any kind of sufficient conditions for the delayed solution, since numerically verifiable sufficient conditions are not available in the literature.

    Figure 4. Optimal solution of the delayed control problem with state delay τ1=1.5, control delay τ2=3.0 and weights B1=5,B2=10. Top row: (a) dose control u1(t) of chemotherapy, (b) effector cells E(t), (c) tumour cells T(t). Bottom row: (a) dose control u2(t) of immune therapy, (b) healthy cells N(t), (c) cytostatic agent U(t).
    Figure 5. Delayed solution with τ1=1.5 and τ2=3.0: controls uk(t) and switching functions ϕk(t),(k=1,2) in a neighborhood of the switching times tk illustrating the control-law (35) and the strict bang-bang property (37).

    Finally, as in the non-delayed case we briefly compare the solutions for the functionals J1(x,u) and J2(x,u). The controls u1 and u2 for the functional J2(x,u) are continuous, since the strict Legendre-Clebsch condition holds and the Hamiltonian has a unique minimum with respect to u1 and u2. Figure 6 displays a comparison of the controls u1 and u2 for both functionals.

    Figure 6. Optimal controls u1(t) and u2(t) for functionals J1(x,u) and J2(x,u) with delays τ1=1.5,τ2=3.0 and weights B1=5,B2=10.

    4.4. Numerical solution of the delayed control problem with mixed control-state constraint U(t)+u2(t)3

    We add the following mixed control-state constraint to the delayed optimal control problem:

    U(t)+u2(t)3   t[0,tf]. (39)

    This constraint means that sum of the cytotoxic agent and the immune dose is bounded from above. Here we consider the augmented Hamiltonian

    H(x,y1,y2,u,v2,λ,μ)=H(x,y1,y2,u,v2,λ)+μ(U+u2), (40)

    where the mixed constraint is adjoined to the Hamiltonian (31) by a multiplier μ0. The local minimum condition (15) yields

    0=Hu2[t]+χ[0,tfτ2](t)Hv2[t+τ2]=ϕ2(t)+μ(t), (41)

    where ϕ2(t)=B2+χ[0,tfτ2](t)λE(t+τ2)s1 is the switching function defined in (34). The multiplier satisfies the complementarity condition μ(t)(U(t)+u2(t)3)=0 for t[0,tf]. Hence, on a boundary arc with U(t)+u2(t)=3 for t[t1,t2] we obtain an explicit formula of the multiplier in view of (41):

    μ(t)=ϕ2(t)=B2χ[0,tfτ2](t)λE(t+τ2)s1   t[t1,t2]. (42)

    Computations show that the control u2(t) is constant on a boundary arc and thus we obtain by differentiation

    0=˙U(t)=u1(t)d1U(t)=u1(t)d1(3u2(t)).
    Figure 7. Optimal solution of the delayed control problem with state delay τ1=1.5, control delay τ2=3.0 and mixed control-state constraint U(t)+u2(t)3. Top row: (a) dose control u1(t) of chemotherapy, (b) function U(t)+u2(t), (c) effector cells E(t). Bottom row: (a) dose control u2(t) of immune therapy, (b) multiplier μ(t) for mixed constraint, (c) tumour cells T(t).

    Since we have u2(t)=1 on a boundary arc, the control u1(t) on the boundary arc is given by

    u1(t)=d1(3u2(t))=0.02   (d1=0.01).

    Using the trapezoidal method (19) with N=3000 grid points we find the control structure

    u1(t)={1for0t<t10.02fort1t<t20fort2ttf},  u2(t)={1for0t<t30fort3ttf} (43)

    with 0<t1<t2<t3<tf and the boundary arc [t1,t2]. We obtain the numerical results:

    J1(x,u)=2236.06,t1=2.045,t2=9.95,t3=10.98,E(tf)=0.725265,T(tf)=0.100546,N(tf)=0.919108,U(tf)=1.63720.

    5. Optimal control of a delay model of Hepatitis B virus infection


    5.1. Optimal control model

    Eikenberry et al. [7] report that currently about two billion people - roughly 30% of the human population - have been infected by Hepatitis B virus (HBV). The disease has attracted considerable attention from mathematical biologists who have developed various models to study the HBV dynamics. Eikenberry et al. [7] present a dynamical model with state variables

    x:  number of healthy cells,p:  number of exposed cells,y:  number of infected cells,v:  free virion load.

    The model (4.1)-(4.4) in [7] does not yet involve a control variable. We choose the control variable u as the effect of treatment which corresponds to the coefficient γ in the dynamic equation (4.4) in [7]. Denoting the time by t[0,tf] with fixed final time tf>0 and the delay in the state variable by τ0, the dynamic system (4.1)-(4.4) in [7] reads as follows:

    ˙x(t)=rx(t)(1T(t)K)dx(t)βv(t)x(t)T(t),˙p(t)=dp(t)+βv(t)x(t)T(t)βedτv(tτ)x(tτ)T(tτ),˙y(t)=βedτv(tτ)x(tτ)T(tτ)ay(t),˙v(t)=k(1u(t))y(t)μy(t). (44)

    The variable T denotes the total number of cells defined by

    T=x+p+y.

    The delay τ appears in all three variables x,p,y. Hence, the initial conditions are given by initial functions for x,p,y and an initial value for v:

    x(t)=x0,p(t)=p0,y(t)=y0  for  τt0,  v(0)=v0. (45)

    We impose the control constraint

    0u(t)1   t[0,tf]. (46)

    Denoting the state vector by X:=(x,p,y,v)R4 and the delayed variable by Y, where Y(t)=X(tτ), the dynamical system can be written as

    ˙X=f(X,Y,u) (47)

    with initial functions and conditions given in (45).

    The optimal control problem then consists in determining a control function uL1([0,tf],R) that minimizes the cost functional

    J(X,u)=tf0(x(t)+Bu(t))dt   (B>0), (48)

    subject to the dynamics (44) with initial conditions (45) and the control constraint (46). The objective functional represents a trade-off between maximizing the number of healthy cells and minimizing the treatment cost.


    5.2. Necessary optimality conditions: Minimum principle

    We briefly discuss the necessary optimality conditions in Theorem 2.1. The Hamiltonian is given by

    H(X,Y,u,λ)=x+Bu+λf(X,Y,u),    λ=(λx,λp,λy,λv)R4. (49)

    We do not explicitly write out the advanced adjoint equation (12):

    ˙λ(t)=HX[t]χ[0,tfτ](t)HY[t+τ]. (50)

    The control variable u appears linearly in the Hamiltonian and does not involve a delay. Hence, defining the switching functions by

    ϕ(t)=Hu[t]=Bλv(t)ky(t), (51)

    the minimizing control is characterized by the control law

    u(t)={0,ifϕ(t)>01,ifϕ(t)<0singular,ifϕ(t)=0tIs[0,tf]}. (52)

    Singular controls will not be discussed further, because we only found bang-bang controls. The following parameters from [7], page 294 below, will be used in our computations:

    a=0.011,d=0.0039,β=4.8105,k=200,K=2,r=1,μ=0.693. (53)

    The state variable X=(x,p,y,v) is scaled by 1011 so that we can choose, e.g., the following initial conditions:

    x(t)=1.4,  p(t)=0.3,  y(t)=0.2    τt0,    v(0)=500. (54)

    The time horizon is tf=500 (days) and the weight parameter in the objective (48) is taken as B=0.05.


    5.3. Comparison of solutions for several delays

    We compare the solutions for the delays τ=0 (non-delayed solution), τ=10 and τ=15. Applying AMPL/IPOPT with N=5000 grid points and using the trapezoidal rule (19), we find a bang-bang control u(t) with only one switch at t1,

    u(t)={1for0t<t10fort1ttf}. (55)

    In the non-delayed case, a refinement of the solution is obtained by solving the Induced Optimization Problem (IOP) with respect to the switching time [17,19]. We get the numerical results:

    τ=0:J(X,u)=893.072,t1=261.70,τ=10:J(X,u)=913.388,t1=293.50,τ=15:J(X,u)=923.032,t1=304.10.

    A comparison of the controls and switching functions for the delays τ=0,10,15 is shown in Figure 8. The bang-bang control for τ=0 provides a strict strong minimum, since second-order sufficient conditions (SSC) in [17,19] are satisfied. The numerical test of SSC proceeds as follows. Since the bang-bang control (55) has only one switch at t1, the objective functional becomes a function J=J(t1) of the scalar optimization variable t1. One verifies numerically that the second derivative is positive: d2J/dt21=0.005028>0. Moreover, the following strict bang-bang property [17,19] for the switching function ϕ(t) holds; cf. Figure 8, left:

    ϕ(t)<0  for0t<t1,  ˙ϕ(t1)>0,  ϕ(t)>0  fort1<ttf=500.
    Figure 8. Controls and switching functions (51) for delays τ=0, τ=10 and τ=15. For all delays the control law (52) is satisfied and the strict bang-bang property holds.

    Note that the strict bang-bang property is also satisfied for the delayed control with delays τ=10 and τ=15. However, as in the preceding section we can not conclude that the delayed controls in Figure 8 provide a strict strong minimum. Figure 9 displays a comparison of the state variables for delays τ=0,τ=10,τ=15.

    Figure 9. Comparison of state variables for delays τ=0,10,15. Top row: (a) healthy cells x, (b) exposed cells p. Bottom row: (a) infected cells y, (b) free virions v.

    6. Conclusion

    We presented two applications of delayed optimal control problems in biomedicine. In the first case study, we extended the delay differential model of tumour-immune-response in Rihan et al. [21] by including a time delay in the control variable u2 which represents the immune therapy. The delay is due to the delayed response of the human immune system to the immune therapy. Rihan et al. [21] considered a L2-type objective which is quadratic in the control variables. From a numerical point of view, the control solution in [21] remained a bit obscure.

    Therefore, we improved the results in this paper in two regards. First, we considered a more realistic L1-type objective which is linear in the two control variables. Secondly, we applied the discretization and nonlinear programming methods [10] (see Section 3) to obtain extremal solutions that satisfy the necessary optimality conditions in Theorem 2.1 with high accuracy. The computations showed that both controls u1 and u2 are of bang-bang type with only one switch from the upper bound uk(t)=uk,max to the zero control uk(t)=0 for k=1,2. Apparently, it is much easier to administer the therapy protocol induced by a bang-bang control then applying a treatment plan resulting from a L2-type objective; cf. Figure 3. In the non-delayed case we could show that the bang-bang controls are indeed optimal, since they satisfy the second-order sufficient conditions in [19,17]. To our knowledge, sufficient conditions for delayed bang-bang controls are not available in the literature. We have also studied the solution under the mixed control-state constraint (39) which combines the cytostatic agent U(t) and the immune control u2(t). The computations gave very accurate extremal solutions.

    The second delay differential model, which describes the spread of Hepatitis B virus, was taken from Eikenberry at al. [7]. We introduced a control variable into the originally uncontrolled model and considered L1-type objectives. For different delays we obtained only bang-bang controls as in the first case study. Sufficient optimality conditions [19,17] could only be verified for the non-delayed bang-bang control.


    [1] [ B. Buonomo,M. Cerasuolo, The effect of time delay in plant-pathogen interactions with host demography, Math. Biosciences and Engineering, 12 (2015): 473-490.
    [2] [ C. Büskens, Optimierungsmethoden und Sensitivitätsanalyse für optimale Steuerprozesse mit Steuer- und Zustands-Beschränkungen, PhD thesis, Institut für Numerische Mathematik, Westfälische Wilhelms-Universität Münster, Germany, 1998.
    [3] [ C. Büskens,H. Maurer, SQP methods for solving optimal control problems with control and state constraints: adjoint variables, sensitivity analysis and real-time control, J. Comput. Appl. Math., 120 (2000): 85-108.
    [4] [ C. Büskens and M. Gerdts, WORHP: Large-Scale Sparse Nonlinear Optimization Solver, http://www.worhp.de.
    [5] [ Q. Chai,R. Loxton,K. L. Teo,C. Yang, A class of optimal state-delay control Problems, Nonlinear Analysis: Real World Applications, 14 (2013): 1536-1550.
    [6] [ R. V. Culshaw,S. Ruan, A delay-differential equation model of HIV infection of CD4+ T-cells, Mathematical Biosciences, 165 (2000): 27-39.
    [7] [ S. Eikenberry,S. Hews,J. D. Nagy,Y. Kuang, The dynamics of a delay model of Hepatitis B virus infection with logistic hepatocyte growth, Mathematical Biosciences, 6 (2009): 283-299.
    [8] [ R. Fourer, D. M. Gay and B. W. Kernighan, AMPL: A Modeling Language for MathematicalProgramming, The Scientific Press, South San Francisco, California, 1993.
    [9] [ L. Göllmann,D. Kern,H. Maurer, Optimal control problems with delays in state and control and mixed control-state constraints, Optimal Control Applications and Methods, 30 (2009): 341-365.
    [10] [ L. Göllmann,H. Maurer, Theory and applications of optimal control problems with multiple time-delays, Journal of Industrial and Management Optimization, 10 (2014): 413-441.
    [11] [ T. Guinn, Reduction of delayed optimal control problems to nondelayed problems, Journal of Optimization Theory and Applications, 18 (1976): 371-377.
    [12] [ R. F. Hartl,S. P. Sethi,R. G. Vickson, A survey of the maximum principles for optimal control problems with state constraints, SIAM Review, 37 (1995): 181-218.
    [13] [ M. R. Hestenes, Calculus of Variations and Optimal Control Theory, John Wiley, New York, 1966.
    [14] [ S. C. Huang, Optimal Control problems with retardations and restricted phase coordinates, Journal of Optimization Theory and Applications, 3 (1969): 316-360.
    [15] [ J. Klamka,H. Maurer,A. Swierniak, Local controllability and optimal control for a model of combined anticancer therapy with control delays, Mathematical Biosciences and Engineering, 14 (2017): 195-216.
    [16] [ Y. Kuang, Delay Differential Equations with Applications in Population Dynamics, Academic Press, San Diego, 1993.
    [17] [ H. Maurer,C. Büskens,J.-H. R. Kim,Y. Kaya, Optimization methods for the verification of second-order sufficient conditions for bang-bang controls, Optimal Control Methods and Applications, 26 (2005): 129-156.
    [18] [ R. M. May, Time-delay versus stability in population models with two and three tropic levels, Ecology, 54 (1973): 315-325.
    [19] [ N. P. Osmolovskii and H. Maurer, Applications to Regular and Bang-Bang Control: Second-Order Necessary and Sufficient Optimality Conditions in Calculus of Variations and Optimal Control, SIAM Advances in Design and Control, Vol. DC 24, SIAM Publications, Philadelphia, 2012.
    [20] [ L. S. Pontryagin, V. G. Boltyanskii, R. V. Gamkrelidze and E. F. Mishchenko, The Mathematical Theory of Optimal Processes, Translation by K. N. Trirogoff, Wiley, New York, 1962.
    [21] [ F. Rihan, D. H. Abdelrahman, F. Al-Maskari, F. Ibrahim and M. A. Abdeen, Delay differential model for tumour-immune-response with chemoimmunotherapy and optimal control. Computational and Mathematical Methods in Medicine, Hindawi Publishing Corporation, Vol. 2014, Article ID 982978, (2014).
    [22] [ H. Schättler,U. Ledzewicz,H. Maurer, Sufficient conditions for strong local optimality in optimal control problems with L2-type objectives and control constraints, Discrete Contin. Dyn. Syst. Ser. B, 19 (2014): 2657-2679.
    [23] [ C. Silva,H. Maurer,D.F.M. Torres, Optimal control of a tuberculosis model with state and control delays, Mathematical Biosciences and Engineering, 14 (2017): 321-337.
    [24] [ C. T. Sreeramareddy,K. V. Panduru,J. Menten,J. V. den Ende, Time delays in diagnosis of pulmonary tuberculosis: A systematic review of literature, BMC Infectious Diseases, 9 (2009): 91-100.
    [25] [ J. Stoer and R. Bulirsch, Introduction ot Numerical Analysis, Third Edition, Texts in Applied Mathematics, Springer-Verlag, Berlin, 1990.
    [26] [ D. G. Storla,S. Yimer,G. A. Bjune, A systematic review in delay in the diagnosis and treatment of tuberculosis, BMC Public Health, 8 (2008): p15.
    [27] [ P. van den Driessche, Some Epidemiological Models with Delays, Report DMS-679-IR, University of Victoria, Department of Mathematics, 1994.
    [28] [ A. Wächter,L. T. Biegler, On the implementation of an interior-point filter line-search algorithm for large-scale nonlinear programming, Mathematical Programming, 106 (2006): 25-57.
    [29] [ H. Yang,J. Wei, Global behaviour of a delayed viral kinetic model with general incidence rate, Discrete Contin. Dyn. Syst. Ser. B, 20 (2015): 1573-1582.
  • This article has been cited by:

    1. F.A. Rihan, S. Lakshmanan, H. Maurer, Optimal control of tumour-immune model with time-delay and immuno-chemotherapy, 2019, 353, 00963003, 147, 10.1016/j.amc.2019.02.002
    2. Cristiana J. Silva, Helmut Maurer, Optimal control of HIV treatment and immunotherapy combination with state and control delays, 2020, 41, 0143-2087, 537, 10.1002/oca.2558
    3. Riccardo Bonalli, Bruno Hérissé, Emmanuel Trélat, Continuity of Pontryagin Extremals with Respect to Delays in Nonlinear Optimal Control, 2019, 57, 0363-0129, 1440, 10.1137/18M119121X
    4. Zohreh Abbasi, Iman Zamani, Amir Hossein Amiri Mehra, Asier Ibeas, Mohsen Shafieirad, Andrei Korobeinikov, Optimal Allocation of Vaccine and Antiviral Drugs for Influenza Containment over Delayed Multiscale Epidemic Model considering Time-Dependent Transmission Rate, 2021, 2021, 1748-6718, 1, 10.1155/2021/4348910
    5. Andrea Cosso, Francesco Russo, Crandall–Lions viscosity solutions for path-dependent PDEs: The case of heat equation, 2022, 28, 1350-7265, 10.3150/21-BEJ1353
    6. Cristiana J. Silva, Stability and optimal control of a delayed HIV/AIDS-PrEP model, 2022, 15, 1937-1632, 639, 10.3934/dcdss.2021156
    7. Aurelio A. de los Reyes, Yangjin Kim, Optimal regulation of tumour-associated neutrophils in cancer progression, 2022, 9, 2054-5703, 10.1098/rsos.210705
    8. Hancheng Guo, Jie Xiong, Jiayu Zheng, Stochastic Maximum Principle for Generalized Mean-Field Delay Control Problem, 2024, 0022-3239, 10.1007/s10957-024-02398-2
    9. Yaxing Ma, Lijuan Wang, Xiuxiang Zhou, Convergence of minimal norm control problems of linear heat equations with time delay, 2024, 0, 2156-8472, 0, 10.3934/mcrf.2024069
    10. Indranil Ghosh, Huey Tyng Cheong, Kok Lay Teo, A novel gradient-based discrete time-delayed optimization algorithm for optimal control problems with Caputo–Fabrizio fractional derivative, 2025, 03770427, 116526, 10.1016/j.cam.2025.116526
  • Reader Comments
  • © 2018 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(5141) PDF downloads(1243) Cited by(10)

Article outline

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog