Processing math: 100%
Research article

Mathematical analysis of a phase-field model of brain cancers with chemotherapy and antiangiogenic therapy effects

  • Received: 22 October 2021 Accepted: 27 October 2021 Published: 28 October 2021
  • MSC : 35B50, 35D30, 35Q92, 92C50

  • Our aim in this paper is to study a mathematical model for brain cancers with chemotherapy and antiangiogenic therapy effects. We prove the existence and uniqueness of biologically relevant (nonnegative) solutions. We then address the important question of optimal treatment. More precisely, we study the problem of finding the controls that provide the optimal cytotoxic and antiangiogenic effects to treat the cancer.

    Citation: Monica Conti, Stefania Gatti, Alain Miranville. Mathematical analysis of a phase-field model of brain cancers with chemotherapy and antiangiogenic therapy effects[J]. AIMS Mathematics, 2022, 7(1): 1536-1561. doi: 10.3934/math.2022090

    Related Papers:

    [1] Kaushik Dehingia, Hemanta Kumar Sarmah, Kamyar Hosseini, Khadijeh Sadri, Soheil Salahshour, Choonkil Park . An optimal control problem of immuno-chemotherapy in presence of gene therapy. AIMS Mathematics, 2021, 6(10): 11530-11549. doi: 10.3934/math.2021669
    [2] Hussein Raad, Cyrille Allery, Laurence Cherfils, Carole Guillevin, Alain Miranville, Thomas Sookiew, Luc Pellerin, Rémy Guillevin . Simulation of tumor density evolution upon chemotherapy alone or combined with a treatment to reduce lactate levels. AIMS Mathematics, 2024, 9(3): 5250-5268. doi: 10.3934/math.2024254
    [3] Hsiu-Chuan Wei . Mathematical modeling of ER-positive breast cancer treatment with AZD9496 and palbociclib. AIMS Mathematics, 2020, 5(4): 3446-3455. doi: 10.3934/math.2020223
    [4] Zholaman Bektemessov, Laurence Cherfils, Cyrille Allery, Julien Berger, Elisa Serafini, Eleonora Dondossola, Stefano Casarin . On a data-driven mathematical model for prostate cancer bone metastasis. AIMS Mathematics, 2024, 9(12): 34785-34805. doi: 10.3934/math.20241656
    [5] Ahmed J. Abougarair, Mohsen Bakouri, Abdulrahman Alduraywish, Omar G. Mrehel, Abdulrahman Alqahtani, Tariq Alqahtani, Yousef Alharbi, Md Samsuzzaman . Optimizing cancer treatment using optimal control theory. AIMS Mathematics, 2024, 9(11): 31740-31769. doi: 10.3934/math.20241526
    [6] Xingxiao Wu, Lidong Huang, Shan Zhang, Wenjie Qin . Dynamics analysis and optimal control of a fractional-order lung cancer model. AIMS Mathematics, 2024, 9(12): 35759-35799. doi: 10.3934/math.20241697
    [7] Laura Hatchondo, Carole Guillevin, Mathieu Naudin, Laurence Cherfils, Alain Miranville, Rémy Guillevin . Mathematical modeling of brain metabolites variations in the circadian rhythm. AIMS Mathematics, 2020, 5(1): 216-225. doi: 10.3934/math.2020013
    [8] Noufe H. Aljahdaly, Nouf A. Almushaity . A diffusive cancer model with virotherapy: Studying the immune response and its analytical simulation. AIMS Mathematics, 2023, 8(5): 10905-10928. doi: 10.3934/math.2023553
    [9] Marcello Delitala, Mario Ferraro . Is the Allee effect relevant in cancer evolution and therapy?. AIMS Mathematics, 2020, 5(6): 7649-7660. doi: 10.3934/math.2020489
    [10] Karim Gasmi, Ahmed Kharrat, Lassaad Ben Ammar, Ibtihel Ben Ltaifa, Moez Krichen, Manel Mrabet, Hamoud Alshammari, Samia Yahyaoui, Kais Khaldi, Olfa Hrizi . Classification of MRI brain tumors based on registration preprocessing and deep belief networks. AIMS Mathematics, 2024, 9(2): 4604-4631. doi: 10.3934/math.2024222
  • Our aim in this paper is to study a mathematical model for brain cancers with chemotherapy and antiangiogenic therapy effects. We prove the existence and uniqueness of biologically relevant (nonnegative) solutions. We then address the important question of optimal treatment. More precisely, we study the problem of finding the controls that provide the optimal cytotoxic and antiangiogenic effects to treat the cancer.



    There has been recently a strong interest in the development, mathematical study and numerical analysis of phase-field models for tumor growth. Such models describe the evolution of a tumor surrounded by healthy tissues and take into account mechanisms such as proliferation of cells via nutrient consumption, therapies, clustering effects in brain tumors, etc. We refer the interested reader to, e.g., [1,3,4,5,8,9,10,11,12,18,21,22,23] for more details.

    In this paper we address a phase-field model of tumor growth driven by a vital nutrient and subject to medical treatment. The model takes into account both the effects of a cytotoxic drug inhibiting the tumor proliferation rate, and those of an antiangiogenic therapy which reduces the nutrient supply. Indeed, since one of the hallmarks of cancer is microvascularization and it is more pronounced in certain tumors (such as gliomas), chemotherapy is often supported by an attendant antiangiogenic drug (usually bevacizumab, see e.g. [7]). This interesting approach has been introduced in the mathematical literature only recently by Colli et al. in [3,4] in the context of prostatic cancer.

    Therein, the authors propose a model of PCa growth and chemotherapy based on [26], and then perform a complete mathematical analysis of the resulting system. This is based on three parabolic equations, each ruling the evolution of a characteristic variable. The first is a phase-field variable u that identifies the spatial location and geometry of the tumor. In particular, the healthy tissue corresponds to tumor absence, that is u=0, while u=1 in the tumoral case. The second variable in play is the concentration of a vital nutrient denoted by σ, while the third variable p identifies the prostate-specific antigen released by cancerous prostatic cells. In the present paper, aiming to model different tumors, we will not consider the last equation. Instead, we will focus on the equation for σ and we will modify it in a form that seems more suitable to describe the evolution of oxygen - one of the main nutrients for brain cancers like glioma (see [13]).

    Here is a description of the model. We assume that the evolution of u is governed by a nonconserved phase-field equation, as justified by Xu, Vilanova and Gomez in [26] by using the concept of tumor free energy and gradient dynamics. It reads as follows

    tuλΔu=Ψu, (1.1)

    where λ is the diffusion coefficient of tumor cells, and Ψ=Ψ(u,σ,c) is a double-well potential having local minima at u=0 and u=1. Note that the so-called chemical free energy Ψ here also depends on the administered cytotoxic drug c. We will detail later the choice of the functional Ψ, explaining in particular how nutrient and chemotherapy influence tumor proliferation rate.

    Concerning the nutrient dynamics, the starting point is the reaction diffusion equation proposed in [3,4]

    tσΔσ=Sh(1u)+(Scs)u(γh(1u)+γcu)σ,

    where γc and γh are the nutrient uptake rate in cancerous and healthy tissue, respectively. Analogously Sc is the nutrient supply rate in the cancerous tissue, while Sh refers to nutrient supply in the healthy tissue. Besides, the model incorporates the action of an antiangiogenic treatment, via the control function s providing a reduction of the intratumoral nutrient supply rate.

    Indeed, since in this paper we take oxygen as nutrient, we modify the equation by considering a nonlinear term of the form

    g(σ)=σ1+σ

    that accounts for the oxygen uptake by cells, assuming Michaelis–Menten kinetics (and setting some biological constants equal to 1), see e.g. [13].

    Accordingly, we propose the following equation for the nutrient dynamics

    tσΔσ=Sh(1u)+(Scs)u(γh(1u)+γcu)g(σ),

    i.e.,

    tσΔσ+γhg(σ)+γchg(σ)u=Sh(1u)+(Scs)u,

    having set γch=γcγh.

    We stress that the nonlinear term g(σ) in the equation of the nutrient is in particular relevant in brain cancers and, more precisely, in gliomas (see [13]). Note that another important nutrient for brain cancers development is lactate (see [2,17]; see also [19,20]); in that case, one also has Michaelis–Menten kinetics, leading to a similar equation for lactate. We will address this model in a forthcoming paper.

    Our aim in this work is twofold.

    First task: Well posedness of the model. Assuming that the controls (c,s) are given functions as above, in the first part of the paper we shall provide an existence and uniqueness result for the proposed model, see (1.2) below, after setting the proper mathematical framework.

    This will be addressed in Sections 2-4; Section 1 is devoted to the precise mathematical setting.

    Second task: Optimal control. We study the problem of finding the controls (c,s) that provide the optimal cytotoxic and antiangiogenic effects to treat a certain glioma described by (1.2).

    From the mathematical viewpoint, the problem consists in

    minimizing a certain cost function J(c,s) subject to system (1.2), in a prescribed class of admissible controls (c,s)Uad.

    This will be accomplished in Sections 5-6, exploiting classical tools of control theory suitably formulated in Appendix 7.

    Let Ω be a bounded and regular domain of RN with N=1,2,3 being the spatial dimension. Let T>0 be a finite time. By defining the space/time sets QT:=Ω×(0,T) and ΣT=Ω×(0,T) the model can be formulated as

    {tuΔu=F(u)+[m(σ)c]h(u)in QT,tσΔσ+γhg(σ)+γchg(σ)u=Sh(1u)+(Scs)uin QT,nu=nσ=0on ΣT,u(0)=u0,σ(0)=σ0in Ω, (1.2)

    where u0 and σ0 are sufficiently smooth given functions defined in Ω. We recall that γh,Sh,Sc are positive constants related with the biological mechanisms in the glioma, while γch=γcγ. Besides,

    g(σ)=σ1+σ.

    The given functions (c,s) correspond respectively to a cytotoxic drug administered as chemotherapy and to an antiangiogenic therapy; they are supposed to be positive and bounded. We further suppose that

    Scs.

    Let us now detail the choice of the nonlinearities F,m,h appearing in the first equation.

    As derived in [26], the phenomenological equation for tumor phase-field relies on the choice of the tumor chemical free energy functional. This is defined as

    Ψ(u,σ)=F(u)m(σ)h(u),

    where F(u)=u2(1u)2 is the prototype double-well potential, the wells being u=0 and u=1. It is perturbed by the term m(σ)h(u), where

    h(u)=u2(32u)

    is an interpolation function with the property h(0)=0, h(1)=1 and h(0)=h(1)=0, while the tilting function m accounts for the effects of hypoxia. Indeed, a possible choice of m is

    m(σ)=2π3.01arctan(σσhv),

    where σhv is the threshold oxygen concentration for hypoxia. According to [26], one should assume that |m(σ)|<1/3 so that the perturbation does not destroy the double-well structure of F. As a consequence, the free energy functional Ψ has a local maximum in (0,1) and preserves two local minima in u=0 and u=1 for any oxygen concentration σ>0.

    To understand how hypoxia influences proliferation, observe that, when σ is below the threshold σhv, the function m(σ) is negative. Thus, since Ψ(0,σ)=0<Ψ(1,σ)=m(σ), the preferable energy level corresponds to healthy tissue. In turn, without hypoxia, σ>σhv, we have Ψ(0,σ)=0>Ψ(1,σ)=m(σ), so the phase-field equation will favor tumor growth.

    At this point, aiming at incorporating in the equation the tumor-inhibiting effect of a cytotoxic drug c, we define

    Ψ(u,σ,c)=F(u)[m(σ)c]h(u).

    Accordingly, the phase-field equation (1.1) reads

    tuλΔu=F(u)+[m(σ)c]h(u),

    that can also be written as

    tuλΔu4u2(1u)=2u(1u)[3(m(σ)c)1].

    We refer the reader to [6,13] for alternative models of hypoxia effects on tumor growth.

    Remark 1.1. For further use, we observe that, when σ0, the Michaelis-Menten nonlinearity g satisfies 0g(σ)<1 and 0<g(σ)<1. Besides,

    F,hC(R).

    Furthermore, as tilting term m we consider any function satisfying

    mC(R):m,mLipschitz continuous withm,m,m"∈L(R). (1.3)

    Without loss of generality, we set λ=1.

    In view of the optimal control problem, we consider a cost function that is based on prescribed target functions for the tumor volume and the oxygen, respectively, on QT and on Ω at the final time T of the pharmacological treatment. Accordingly, for assigned uQ,σQL2(QT), uΩ,σΩL2(Ω), we define

    J(c,s)=k12QT[u(x,t)uQ]2dxdt+k22Ω[u(x,T)uΩ]2dx+k3Ωu(x,T)dx+k42QT[σ(x,t)σQ]2dxdt+k52Ω[σ(x,T)σΩ]2dx+k62QTc2(x,t)dxdt+k72QTs2(x,t)dxdt,

    where (u,σ) is the (unique) solution to (1.2) originated by any observed initial state (u0,σ0) of the system. Besides, the set of admissible controls will be

    Uad={(c,s)L2(QT)×L2(QT):0cUmax,0sSmaxa.e. inQT},

    where the given quantities Umax>0 and 0<SmaxSc are two threshold positive values. We shall prove the existence of an optimal control and we shall provide a necessary condition for a control to be optimal that, in particular, allows its identification via numerical simulations.

    We will use the classical Lebesgue spaces Lp(Ω) (p1), denoting their norms by Lp, and the Sobolev spaces Hk(Ω) of functions in L2(Ω) with distributional derivative of order less than or equal to k in L2(Ω). As customary, we set H=L2(Ω) with inner product denoted by (,) and corresponding norm . We also set V=H1(Ω) equipped with the norm

    f2V=f2+f2,

    and by V its dual space, the symbol , standing for the corresponding duality pairing.

    Finally, we set

    WH2N(Ω)={uH2(Ω):nu=0 on Ω}V.

    We will also make use of spaces of functions that depend on time with values in a Banach space. Hence, given a generic Banach space B with norm B and an interval I[0,), Lp(I;B) is the set of measurable functions f:IB such that tf(t)B belongs to Lp(I). Recall that L2(I;H) is isomorphic to L2(Ω×I). With the symbol W1,p(I;B) we will denote functions f:IB such that both f and its (weak) derivative tf belong to Lp(I;B). The family of continuous functions f:IB is denoted by C(I,B).

    Throughout the paper, by C>0 we shall denote a constant that may change from line to line, depending on the problem parameters, the final time T, the norms of the initial data, and possibly on the norms of c and s.

    For a fixed T>0, we set QT=Ω×(0,T) and we introduce the phase space

    XW1,2(0,T;V)L2(0,T;V)C([0,T],H).

    Definition 2.1. Let cL(QT), sL(QT) be given and take (u0,σ0)H×H. A (weak) solution on [0,T] to the initial value problem (1.2) endowed with Neumann boundary conditions is a pair (u,σ) with

    uXandσX

    satisfying

    tu(t),v+(u(t),v)=F(u)+[m(σ)c]h(u),v,vV,tσ(t),w+(σ,w)+γh(g(σ),w)+γchg(σ)u,w=Sh(1u,w)+(Scs)(u,w),wV,

    for almost every t(0,T). Moreover, nu=nσ=0 almost everywhere on ΣT and (u(0),σ(0))=(u0,σ0) almost everywhere in Ω.

    Remark 2.2. By a classical result (see, e.g., [24]), the regularity fL2(0,T;V), tfL2(0,T;V) ensures that fC([0,T],H). Besides, tu(t)2 is absolutely continuous and ddtf2=tf,f.

    Theorem 2.3. Let cL(QT), sL(QT) with sSc be given, and (u0,σ0)H×H be such that

    0u01andσ00a.e. in Ω.

    Then, system (1.2) has a unique weak solution (u,σ)X×X such that

    0u1andσ0a.e. (x,t) in QT.

    Besides, the following uniform estimate holds:

    u2X+σ2XC(u02+σ02+1).

    Furthermore, if σ0L(Ω), then σL(QT) and

    σLC(σ0L+1). (2.1)

    In particular, our theorem tells that the biologically relevant region

    S={(u,σ)H×H:0u1,σ0},

    is invariant for the differential system (1.2), namely: if we consider any biologically meaningful initial datum z0=(u0,σ0)S, then any weak solution of (1.2) departing from z0 remains in S for every time.

    The next Section 3 and Section 4 are devoted to the proof of Theorem 2.3 via a number of steps. The first consists in the introduction of an auxiliary problem where we suitably modify some of the nonlinearities involved in system (1.2).

    In this and the next section, according to the assumptions of Theorem 2.3, we assume (c,s)L(QT)×L(QT), with sSc, fixed. We introduce the cut–off function

    k(r)={2r(1r)r[0,1],0r[0,1].

    Note that k is globally bounded and Lipschitz on R. Then, defining

    ˜f(σ,c)=[13(m(σ)c)],

    we consider the auxiliary system

    {tuΔu4u2(1u)=˜f(σ,c)k(u)in QT,tσΔσ+γhσ1+|σ|+γchσu1+|σ|=Sh(1u)+(Scs)uin QT,nu=nσ=0on ΣT,u(0)=u0,σ(0)=σ0in Ω. (3.1)

    Due to the special form of the nonlinearities, it is easy to show that any solution of system (3.1) originated from an initial data (u0,σ0)S belongs to S a.e. in QT. This is done in the sext section.

    Let (u(t),σ(t))H×H be any weak solution on [0,T] to the auxiliary Cauchy problem (3.1).

    Lemma 3.1. If 0u01 a.e. in Ω, then 0u(t)1 a.e. in QT.

    Proof. Testing the first equation of (3.1) by uV, where u=max(0,u), we obtain

    12ddtu2+u2+4u4L4+4u3L3=Ω˜f(σ,c)k(u)u.

    Indeed, the rhs is identically zero since k(u)=0 whenever u0. As a result,

    ddtu20,

    and the Gronwall lemma yields

    u(t)2u(0)2=0.

    This means that u0 a.e. in QT. Now we consider w=u1. Note that w solves the equation

    twΔw+4u2w=˜f(σ,c)k(u),

    hence, testing by w+ we get

    12ddtw+2+w+2+4Ωu2(w+)2=Ω˜f(σ,c)k(u)w+.

    Again the rhs is identically zero since by construction k(u)=0 whenever u1, namely, on the support of w+. Reasoning as above, we reach the conclusion

    w+(t)2w+(0)2.

    Since w(0)=u010, then w+(0)=0: as a consequence w+(t)=0 in Ω×[0,T], meaning that u1 a.e., as claimed.

    Lemma 3.2. If 0u01 and σ00 a.e. in Ω, then σ(t)0 a.e. in QT.

    Proof. Testing the second equation of (3.1) by σV, and taking into account that 0u1 a.e., we get

    12ddtσ2+σ2+γhΩ|σ|21+|σ|=γchΩu|σ|21+|σ|ShΩ(1u)σΩ(Scs)uσ|γch|Ω|σ|2.

    Integrating over [0,t] the final differential inequality ddtσ22|γch|σ2, we obtain

    σ(t)2e2|γch|tσ(0)2

    for all times. Since by assumption σ00, it turns out that σ(0)=0, yielding the thesis.

    Let (u0,σ0)H×H be arbitrarily given. In this section we prove that the Cauchy problem (3.1) admits (at least) a local solution, which is defined in a maximal time interval [0,τ), for some τ>0.

    Let {ej}j=1 be a smooth orthonormal basis in H which is also orthogonal in V. Then define Vn=Span{e1,,en} and denote by Pn the corresponding projection. Now, for any fixed nN, we consider the following finite dimensional problem: Find tn>0 and functions aj,bjC1([0,tn)) such that

    un(t)=nj=1aj(t)ejandσn(t)=nj=1bj(t)ejC1([0,tn),Vn)

    satisfy, for almost every t(0,tn),

    tun(t),v+(un,v)=(4u2n(1un)+˜f(σn,c)k(un),v),

    and

    tσn(t),w+(σn,w)=(γhσn1+|σn|γchσnun1+|σn|+Sh(1un)+(Scs)un,w),

    for every test function vVn and wVn, along with the initial conditions

    un(0)=Pnu0,σn(0)=Pnσ0,a.e. inΩ.

    Indeed, choosing v=w=ej for any j{1,,n}, everything boils down to a system of 2n nonlinear ordinary differential equations with locally Lipschitz nonlinearities. Hence, by classical results in ODE's theory, the local existence (and uniqueness) of a solution (un,σn) is guaranteed on a certain maximal interval [0,tn). Besides, the solution satisfies

    un,σnC1([0,tn),V).

    We now wish to find estimates that are independent of n.

    Along the proof, C>0 will stand for a generic constant independent of n. Test the first equation by un and the second one by σn to find

    12ddt(un2+σn2)+un2+σn2+4un4L4+γhΩσ2n1+|σn|=4un3L3+Ω˜f(σn,c)k(un)unγchΩunσ2n1+|σn|+ShΩ(1un)σn+Ω(Scs)unσn.

    We now estimate the rhs. Recalling that |k(r)|c(1+r2) and that ˜f(σn,c)LC by construction, we have

    Ω˜f(σn,c)k(un)unC(1+un3L3).

    Besides, we easily obtain

    γchΩunσ2n1+|σn||γch|Ω|un||σn|C(un2+σn2)

    and

    ShΩ(1un)σn+Ω(Scs)unσnC(un2+σn2)+C.

    It is then apparent that we end up with the inequality

    12ddt(un2+σn2)+4un4L4+un2+σn2Cun3L3+C(un2+σn2)+C.

    There we can control the L3-norm of un via the Young inequality with exponents 43,4 as follows

    Cun3L3Cun3L43un4L4+C.

    Thus we arrive at

    ddtΛ+un4L4+un2+σn2CΛ+C (3.2)

    having set

    Λ(t)=un(t)2+σn(t)2.

    Therefore, by Gronwall's lemma,

    Λ(t)Ct[0,tn],

    so that there exists τ>0 independent on n such that

    (un(t),σn(t))C,t[0,τ].

    Since there exists C>0 such that

    un4L4+un212un4L4+un2VC

    then, going back to (3.2) and integrating in time over [0,τ], we further learn that

    (un(t),σn(t))[L2(0,τ;V)]2C,

    and

    unL4(Qτ)C.

    Hence, by comparison,

    (tun,tσn)[L2(0,τ;V)]2C.

    Due to uniform bounds above, there exists (u,σ) such that

    unu weak star in L(0,τ;H) and weakly inL2(0,τ;V)L4(Qτ),σnσ weak star in L(0,τ;H) and weakly inL2(0,τ;V).

    By the uniform control on tun and tσn, we also learn that

    unuandσnσ strongly in L2(0,τ;H),

    so, in particular,

    (un,σn)(u,σ)a.e. (x,t)Qτ.

    This allows to pass to the limit in the weak formulation to prove that (u,σ) is a weak solution of (3.1) on [0,τ]. All the convergences are straightforward but those involving the nonlinear terms. We start by proving that for any wV, for any φC0(0,t) with tτ,

    t0u3nu3,wφ(y)dy0.

    This is easily seen by noticing that fn=u3nu30 a.e. in Qτ and fnL4/3(Qτ)C, hence fn0 weakly in L4/3(Qτ) (Lebesgue convergence, weak form). Let us now prove that

    t0[˜f(σn,c)k(un)˜f(σ,c)k(u)],wφ(y)dy0.

    To this end, observe that, since ˜f is globally Lipschitz and k is a bounded function,

    t0[˜f(σn,c)˜f(σ,c)]k(un),wφ(y)dyCφt0Ω|σnσ||w|dyCφt0σn(y)σ(y)wdy0

    by the strong convergence σnσ in L2(0,τ;H). Analogously, exploting the fact that ˜f is bounded and k globally Lipschitz

    t0˜f(σ,c)[k(un)k(u)],wφ(y)dyCφt0un(y)u(y)wdy0

    by the strong convergence unu in L2(0,τ;H).

    In the second equation, we have to show that

    t0unσn1+|σn|uσ1+|σ|,wφ(y)dy0asn+.

    Indeed, we rewrite the difference as

    unσn1+|σn|uσ1+|σ|=(unu)σn1+|σn|+u[σn1+|σn|σ1+|σ|]

    and, noticing that |s|1+|s|1, we obtain

    t0(unu)σn1+|σn|,wφ(y)dyCφwVt0unudy

    where

    t0unudyτ(t0unu2dy)1/20.

    On account of the Lipschitz continuity of s1+|s|, we find

    |σn1+|σn|σ1+|σ|||σnσ|.

    Hence, the last term in the second equation can be handled as

    t0(σn1+|σn|σ1+|σ|),wφ(y)dyφt0σn1+|σn|σ1+|σ|wdyφwVt0σnσdyτφwV(t0σnσ2dy)1/20.

    In this section we show that any solution to the auxiliary initial value problem (3.1) originated from (u0,σ0)S is defined for all positive times.

    Theorem 3.3. Let T>0 be given and z0=(u0,σ0)S. Then, any weak solution (u,σ) to (3.1) departing from z0 is global in time on [0,T].

    Proof. Let us define

    ¯t=sup{t0: there exists a weak solution in S on [0,t) departing from z0}.

    We know by the previous section that there exists a solution (u,σ)S defined on [0,τ], hence ¯tτ>0 and

    u(t)2|Ω|t[0,¯t).

    Besides, inequality (3.2) holds for (u,σ) in place of (un,σn) since all the involved constants are independent of n. Then Gronwall's lemma yields

    σ(t)2cect,t[0,¯t), (3.3)

    for some c>0 independent of ¯t. This implies that ¯t=T: indeed, the uniform bounds of the H-norms tell that limt¯tu(t) and limt¯tσ(t) exist in H (at least for a subsequence). Now we can consider a solution to the Cauchy problem with initial datum (u(¯t),σ(¯t)), which is defined on an interval [¯t,¯t+δ]), for some δ>0 (see also the extension theorem [14, Lemma 3.1, p. 13]). In this way we contradict the definition of ¯t.

    Let T>0 and take an initial datum (u0,σ0)S. In light of the above analysis, let (u,σ) be any global weak solution to the auxiliary system (3.1) originated by (u0,σ0). We actually proved in Section 3.1 that (u,σ)S a.e. on QT, implying that

    k(u)=2u(1u)andσ1+|σ|=σ1+σ=g(σ).

    It turns out that the pair (u,σ) actually solves the original problem (1.2) on [0,T]. This proves the first part of Theorem 2.3, namely the global existence of solutions to the original model and the invariance of the set S. Let us now prove uniform energy estimates and a continuous dependence result, where we highlight the role of the controls for further use.

    Theorem 4.1. Let (u,σ) be a solution to (1.2) originated from (u0,σ0)S. Then, the following uniform estimate holds

    u2X+σ2XC(u02+σ02+c2L2(0,T;H)+s2L2(0,T;H)+1).

    Proof. Along the line of Section 3.2.2, we multiply the first equation of (1.2) by u and the second one by σ to find

    12ddt(u2+σ2)+u2+σ2+4u4L4+γhΩσ21+σ=4u3L3+Ω˜f(σ,c)k(u)uγchΩuσ21+σ+ShΩ(1u)σ+Ω(Scs)uσ.

    At this point we estimate the rhs, recalling that 0u1, but now paying attention to the role of c and s. Since

    ˜f(σ,c)=[13(m(σ)c)]

    and m is bounded, we have

    Ω˜f(σ,c)k(u)uC(c2+1).

    Besides, we easily get

    γchΩuσ21+σ|γch|Ω|u||σ|C(u2+σ2),

    and

    ShΩ(1u)σ+Ω(Scs)uσC(σ2+s2+1).

    Hence, calling

    Λ(t)=u(t)2+σ(t)2,

    we obtain the differential inequality

    ddtΛ+ω(u4L4+u2V+σ2)CΛ+C(c2+s2+1),

    for some ω>0. An application of the Gronwall lemma on [0,T] yields

    Λ(t)Λ(0)eC+CeCT0(c(y)2+s(y)2+1)dy,

    for every t[0,T], saying that

    u(t)2+σ(t)2C(u02+σ02+c2L2(0,T;H)+s2L2(0,T;H)+1),

    for every t. Going back to the differential inequality and integrating in time over [0,T], we obtain the desired control for u and σ in L2(0,T;V) and uL4(Qτ). Finally, by comparison in the system we get an analogous estimate for (tu,tσ)[L2(0,τ;V)]2C, completing the proof.

    Theorem 4.2. Let (ui,σi) be two solutions to to (1.2) corresponding to controls (ci,si)L(QT)×L(QT), with siSc, and initial data zi=(u0,i,σ0,i)S, i=1,2. Then, the following continuous dependence estimate holds:

    u1(t)u2(t)2+σ1(t)σ2(t)2+u1u22L2(0,T;V)+σ1σ22L2(0,T;V)C(z1z22+c1c22L2(0,T;H)+s1s22L2(0,T;H))

    for all t[0,T].

    Proof. Observe that, by Section 3, we know that 0ui1 and σi0 a.e. in QT, for i=1,2. Let us denote by (u,σ)=(u1u2,σ1σ2) and (c,s)=(c1c2,s1s2). Then

    {tuΔu+4u(u21+u1u2+u22)=4u(u1+u2)+˜f(σ1,c1)k(u1)˜f(σ2,c2)k(u2),tσΔσ+γh[g(σ1)g(σ2)]+γch[g(σ1)u1g(σ2)u2]=(ScSh)us1u1+s2u2.

    Recalling that ˜f(σ,c)=[13(m(σ)c)], we rewrite

    ˜f(σ1,c1)k(u1)˜f(σ2,c2)k(u2)=[˜f(σ1,c1)˜f(σ2,c2)]k(u1)+˜f(σ2,c2)[k(u1)k(u2)]=3[m(σ2)m(σ1)+c]k(u1)˜f(σ2,c2)[k(u2)k(u1)].

    Then, multiplying the first equation by u, taking into account that u1+u22, we get

    12ddtu2+u28u2+3Ω[m(σ2)m(σ1)+c]k(u1)uΩ˜f(σ2,c2)[k(u2)k(u1)]u.

    Since m is a globally Lipschitz function and k(u) is bounded, we immediately get

    3Ω[m(σ2)m(σ1)+c]k(u1)uCΩ|u|(|σ|+|c|)cu(σ+c).

    Besides, exploiting the global Lipschitz continuity of k and the boundedness of ˜f,

    Ω˜f(σ2,c2)[k(u2)k(u1)]uCu2.

    We thus end up with

    12ddtu2+u2C(u2+σ2+c2).

    As a second step we consider the second equation in the differential system solved by (u,σ). We observe that

    g(σ1)u1g(σ2)u2=[g(σ1)g(σ2)]u1+g(σ2)u,

    hence a multiplication by σ yields

    12ddtσ2+σ2=Ω(γh+γchu1)[g(σ1)g(σ2)]σγchΩg(σ2)uσ+Ω(ScShs2)uσΩsu1σ.

    Since |g(σ1)g(σ2)||σ| and 0u11, the first term on the rhs is easily estimated as

    Ω(γh+γchu1)[g(σ1)g(σ2)]σCσ2.

    We proceed noticing that, since 0g(σ)<1,

    γchΩg(σ2)uσCΩ|σ||u|Cuσ.

    Finally,

    Ω(ScShs2)uσΩsu1σC(u+s)σ.

    Collecting everything, we end up with the differential inequality

    ddt(u2+σ2)+ω(u2V+σ2V)C(u2+σ2)+C(c2+s2),

    for some ω>0. Let now T>0 be fixed. An application of the Gronwall lemma on [0,T] yields

    u(t)2+σ(t)2eC(u(0)2+σ(0)2)+CeCT0(c(y)2+s(y)2)dy,

    where u(0)=u0,1u0,2 and σ(0)=σ0,1σ0,2, proving the claimed continuous dependence estimate.

    As an immediate consequence of Theorem 4.1, we see that the (global) weak solution to (1.2) departing from any (u0,σ0)S is unique. Indeed, let us denote by (ui,σi), i=1,2 two solutions, corresponding to a fixed pair of controls (c,s)L(QT)×L(QT) with sSc, departing from the same initial datum z0=(u0,σ0)S. Then, setting c1=c2=c, s1=s2=s and z1=z2=z0 in the continuous dependence estimate, we get

    u(t)2+σ(t)20,t[0,T],

    saying that (u,σ)(0,0) hence (u1,σ1)(u2,σ2) in [0,T].

    As a last step, we are left to prove that, if the initial datum σ0 is bounded, then the solution remains bounded for all times. To this aim, we rewrite the equation for σ as

    tσΔσ=γhσ1+σγchσu1+σ+Sh(1u)+(Scs)u,

    noticing that the right-hand side belongs to L(0,T;H). If we consider σ0L(Ω), then, by a classical result in the theory of linear parabolic PDEs (see e.g. Theorem 7.1 in [16]), we immediately find the desired conclusion σL(QT), together with the uniform estimate (2.1). The proof of Theorem 2.3 is now completed.

    From now on, let

    (u0,σ0)Swithσ0L(Ω)

    be fixed. In light of the existence result Theorem 2.3, we can define the control-to-state mapping as

    G:U={(c,s)L(QT)×L(QT):sSc}H×H
    (c,s)(u,σ),

    where (u,σ) is the unique weak solution to (1.2) corresponding to (c,s) with initial datum (u0,σ0) as above. Here we set

    H=C([0,T],H)L2(0,T;V).

    Besides, by Theorem 2.3 we also know that

    0u1,σ0 a.e. QTand σL(QT)C. (5.1)

    Observe that the mapping G is Lipschitz continuous (having endowed L(QT) with the L2- topology); indeed, owing to Theorem 4.2 we have

    u1u22H+σ1σ22HC(c1c22L2(QT)+s1s22L2(QT)), (5.2)

    for all (ci,si)U and associated states (ui,σi)=G(ci,si).

    Let us now show that G possesses certain directional derivatives at any point in U. To this aim, let (c,s)U be fixed and denote by (u,σ)=G(c,s) the corresponding state. Then, for (c,s)U, we introduce the linearized system at (u,σ), defined as

    {YtΔY+AYBZ=ch(u)in QT,ZtΔZ+CZ+DY=suin QT,nY=nZ=0on ΣT,Y(0)=Z(0)=0in Ω, (5.3)

    where the coefficients are defined as follows

    A=F"(u)m(σ)h"(u)+ch"(u),B=m(σ)h(u),C=(γh+γchu)g(σ),D=sSch+γchg(σ),

    and Sch=ScSh. Notice that the four coefficients, as well as the source terms ch(u) and su, are in L(QT), due to the assumptions on the nonlinearities and the fact that 0u1 and σ0 a.e. in QT. By the theory of linear parabolic equations (see the subsequent Theorem 7.1), there exists a unique strong solution to (5.3) with

    Y2C([0,T],V)L2(0,T;W)+Z2C([0,T],V)L2(0,T;W)C(c2L2(0,T;H)+s2L2(0,T;H)). (5.4)

    At this point, we consider any (ˉc,ˉs)U and notice that

    (cμ,sμ)=(c+μ(ˉcc),s+μ(ˉss))U

    for any μ(0,1). Therefore, we can consider the corresponding state (uμ,σμ)=G(cμ,sμ) satisfying all the results proven in the previous sections. Note that, letting μ0, by construction cμc and sμs in L2(QT) As a consequence, since G is Lipschitz continuous by (5.2), we have

    uμuandσμσinH. (5.5)

    Lemma 5.1. In the limit μ0+ we have

    (uμuμ,σμσμ)(Y,Z)inH×H,

    where (Y,Z) is the solution to the linearized system (5.3) with (c,s)=(ˉcc,ˉss).

    Proof. We set

    Yμ=uμuμY,Zμ=σμσμZ.

    Accordingly, we have to prove that Yμ0 and Zμ0 in H. The first step consists in writing in a suitable form a differential system for (Yμ,Zμ). After some computations, it is not difficult to check that the following holds:

    {YμtΔYμ+A1Yμ+A2Y+A3Zμ+A4Z=c[h(uμ)h(u)],ZμtΔZμ+B1Yμ+B2Y+B3Zμ+B4Z=s[uμu], (5.6)

    having defined

    A1=F"(xμ)+[m(σ)c]h"(xμ),A2=[F"(xμ)F"(u)][m(σ)c][h"(xμ)h"(u)],A3=m(sμ)h(uμ),A4=m(σ)h(u)m(sμ)h(uμ),

    and

    B1=γchg(σμ)Sch+s,B2=γch[g(σμ)g(σ)],B3=γhg(sμ)+γchug(sμ),B4=γh[g(sμ)g(σ)]+γchu[g(sμ)g(σ)].

    Here, xμ, xμ and sμ, sμ are measurable functions arising from the application of an extension of Lagrange Theorem (see [4, Appendix]) as follows:

    F(uμ)F(u)=(uμu)F"(xμ),h(uμ)h(u)=(uμu)h"(xμ),m(σμ)m(σ)=(σμσ)m(sμ),g(σμ)g(σ)=(σμσ)g(sμ).

    We recall that xμ and xμ attain intermediate values between the ones of uμ and u, while sμ and sμ are in between σμ and σ.

    As a second step, we test system (5.6) with the pair (Yμ,Zμ), so obtaining the differential equality

    12ddt(Yμ2+Zμ2)+Yμ2+Zμ2=ΩA1|Yμ|2ΩB3|Zμ|2Ω(A3+B1)YμZμΩ(A2YYμ+A4ZYμ+B2YZμ+B4ZZμ)Ωc[h(uμ)h(u)]YμΩs[uμu]Zμ.

    We proceed by estimating the rhs. Since Ai,BjL(QT) in light of (5.1) and thanks to the regularity of the involved nonlinearities, the first three terms in the rhs are easily controlled by

    C(Yμ2+Zμ2).

    Besides, the last two terms can be estimated exploiting the Lipschitz continuity of h as

    Cuμu2+C(Yμ2+Zμ2).

    Finally, exploiting the fact that Y,ZL(0,T;V) by (5.4),

    Ω(A2YYμ+A4ZYμ+B2YZμ+B4ZZμ)C(A2+A4)YμV+C(B2+B4)ZμV12(Yμ2+Zμ2)+C(Yμ2+Zμ2)+C(A22+A42+B22+B42).

    Integrating on [0,t], observing that Yμ(0)2+Zμ(0)2=0, we get

    Yμ(t)2+Zμ(t)2+t0(Yμ2+Zμ2)dyCt0(Yμ2+Zμ2)dy+Rμ

    having set

    Rμ=C(A22L2(QT)+A42L2(QT)+B22L2(QT)+B42L2(QT))+Cuμu2L2(QT).

    We claim that

    limμ0Rμ=0.

    Indeed, we have the convergences (5.5), implying in turn that

    xμ,xμuandsμ,sμσinC([0,T],H).

    Now the conclusion follows by invoking the Lipschitz continuity of F",h,h" and of m,g,g. As a final step we apply the Gronwall lemma on [0,T] that yields

    Yμ(t)2+Zμ(t)2+t0(Yμ2+Zμ2)dyRμ,t[0,T].

    Letting μ0 the proof is done.

    Our optimal control problem consists in finding the control functions c and s (if any) that provide the optimal cytotoxic and antiangiogenic effects to treat a certain glioma whose evolution is modeled by (1.2).

    In order to state the problem, we first fix the desired targets for the tumor phase and for the oxygen in QT and in Ω at the final time T, respectively given by

    uQ,σQL2(QT)anduΩ,σΩL2(Ω). (6.1)

    Then, for any (u,σ)[C([0,T],H)]2 and any (c,s)[L2(0,T;H)]2, we introduce the functional

    J(u,σ,c,s)=k12QT[u(x,t)uQ]2dxdt+k22Ω[u(x,T)uΩ]2dx+k3Ωu(x,T)dx+k42QT[σ(x,t)σQ]2dxdt+k52Ω[σ(x,T)σΩ]2dx+k62QTc2(x,t)dxdt+k72QTs2(x,t)dxdt,

    where ki are given nonnegative constants, with at least one strictly positive.

    Next, we define the set of all the admissible controls (c,s). Given two positive thresholds Umax>0 and 0<SmaxSc, we define

    K1={cL2(QT):0cUmaxa.e. in QT},K2={sL2(QT):0sSmaxa.e. in QT},

    and we set

    Uad={(c,s)L2(QT)×L2(QT):cK1,sK2}.

    Finally, given the initial state

    (u0,σ0)Swithσ0L(Ω),

    we consider the control-to-state map defined in Section 5. Accordingly, for any pair (c,s)Uad we set (u,σ)=G(c,s) as the weak solution to (1.2) corresponding to (c,s) with initial datum (u0,σ0), and we define the (reduced) cost functional

    J(c,s)=J(G(c,s),c,s).

    Our control problem can be stated as follows: find, if possible, an optimal control (c,s)Uad such that

    J(c,s)=min(c,s)UadJ(c,s). (6.2)

    We start the analysis by proving an existence result.

    Theorem 6.1. Under assumption (6.1), for any fixed (u0,σ0)S with σ0L(Ω) there exists at least a solution (c,s)Uad to (6.2) with corresponding optimal state (u,σ).

    Proof. Indeed, since J0, it is immediate to see that inf(c,s)UadJ(c,s)=δ0. We can consider then a minimizing sequence (cn,sn)Uad such that

    δJ(cn,sn)δ+1n,nN,

    and, according to Theorem 2.3, the corresponding state (un,σn): this, in particular is uniformly bounded in X×X with 0un1 a.e. in QT and σn0 a.e. in QT satisfies (2.1). By the boundedness of Uad and Theorem 4.1, we can select subsequences (that we still denote as) (cn,sn) and (un,σn) such that

    (cn,σn)(c,s)weak star inL(QT),(un,σn)(u,σ)weakly inH1(0,T;V)L2(0,T;V)andweak star inL(QT).

    It is worth noticing that so far no relation connects (c,s) and (u,σ). Our aim will be to prove that (u,σ)=G(c,s), namely, that (u,σ) is the state corresponding to the control, and that J(c,s)=δ. First of all, by compactness,

    (un,σn)(u,σ)strongly inL2(0,T;H), (6.3)

    and, owing to the Ascoli-Arzelá Theorem,

    (un(t),σn(t))(u(t),σ(t))strongly inV×V,uniformly int[0,T]. (6.4)

    Therefore, it follows that (u(0),σ(0))=(u0,σ0). Besides, (u,σ)S and σ satisfies (2.1). Furthermore, due to the boundedness and Lipschitz continuity of all the involved nonlinear functions, these convergences allow to pass to the limit in the problem solved by (un,σn), proving that (u,σ) solves the initial boundary value problem correspondig to (c,s), that is, (u,σ)=G(c,s).

    To accomplish our second task, we decompose the functional J in three parts, namely,

    J=J1+J2+J3,

    where

    J1(c,s)=k12QT[u(x,t)uQ]2dxdt+k42QT[σ(x,t)σQ]2dxdt,J2(c,s)=k3Ωu(x,T)dx,J3(c,s)=k22Ω[u(x,T)uΩ]2dx+k52Ω[σ(x,T)σΩ]2dx+k62QTc2(x,t)dxdt+k72QTs2(x,t)dxdt.

    Now, convergence (6.3) immediately gives

    limnJ1(cn,sn)=J1(c,s).

    By the uniform boundedness of un(T) following from Theorem 4.1, we infer that, up to a subsequence,

    un(T)u(T)weakly inH×H

    so that

    limnJ2(cn,sn)=J2(c,s).

    The last functional J3 is weakly lower semicontinuous thus

    J3(c,s)lim infnJ3(cn,sn).

    Collecting all our computations, we conclude

    δJ(c,s)lim infnJ(cn,sn)=δ,

    showing that indeed J realizes its minimum value at (c,s).

    Once the existence of an optimal control is established, the next goal is devising a necessary condition for a control to be optimal that, in particular, allows its identification by numerical simulations.

    Let (c,s)Uad be an optimal control and denote by (u,σ) the corresponding optimal state. Then, for any (ˉc,ˉs)Uad, we notice that

    (cμ,sμ)=(c+μ(ˉcc),s+μ(ˉss))Uad

    for any μ(0,1). Therefore, we can consider the corresponding state (uμ,σμ) and observe that

    J(cμ,sμ)J(c,s)μ0,μ(0,1). (6.5)

    Now, owing to Lemma 5.1, we can pass to the limit as μ0+ in (6.5), saying that the derivative of J at (c,s) in the direction of (ˉcc,ˉss) is nonnegative. Invoking Lemma 5.1 once again, we easily obtain

    {k1QT(uuQ)Ydxdt+k2Ω(u(T)uΩ)Y(T)dx+k3ΩY(x,T)dx+k4QT(σ(x,t)σQ)Zdxdt+k5Ω[σ(T)σΩ]Z(T)dx}+k6QTc(ˉcc)dxdt+k7QTs(ˉss)dxdt0, (6.6)

    where (Y,Z) solves the linearized problem (5.3). The above inequality is the so-called first order optimality condition, although from its expression it is really difficult to identify the optimal control even by numerical simulations.

    Therefore, as it is done in the classical control theory (see the subsequent Theorem 7.1), we introduce the so-called adjoint problem, here defined as

    {wtΔw+Aw+Dz=k1(uuQ)in QT,ztΔz+CzBw=k4(σσQ)in QT,nw=nz=0on ΣT,w(T)=k2[u(T)uΩ]+k3,z(T)=k5[σ(T)σΩ]in Ω, (6.7)

    where ki, i=1,,5, and uQ,σQ,uΩ,σΩ are exactly the constants and the target functions appearing in the cost functional.

    By Theorem 7.1, we learn that there exists a unique weak solution (w,z)X×X to (6.7) and that the analogous of (7.1) holds true, namely,

    k1QT(uuQ)Ydxdt+k2Ω(u(T)uΩ)Y(T)dx+k3ΩY(x,T)dx+k4QT(σ(x,t)σQ)Zdxdt+k5Ω[σ(T)σΩ]Z(T)dx=QT[(ˉcc)h(u)w(ˉss)uz]dxdt.

    As a consequence, inequality (6.6) turns into the much simpler form

    QT[(ˉcc)h(u)w(ˉss)uz]dxdt+k6QTc(ˉcc)dxdt+k7QTs(ˉss)dxdt0,

    that we write as

    (h(u)wk6c,cˉc)L2(QT)+(uzk7s,sˉs)L2(QT)0,(ˉc,ˉs)Uad,

    Notice that this is equivalent to

    (h(u)wk6c,cˉc)L2(QT)0,ˉcK1,(uzk7s,sˉs)L2(QT)0,ˉsK2.

    The geometric meaning of these inequalities is clear: indeed, leaning on the elementary theory of projections in Hilbert spaces (see e.g. Remark 4.6 in [4]), we have obtained the first order optimality conditions

    Theorem 6.2 (First order optimality conditions). Let (c,s)Uad be an optimal control, with corresponding state (u,σ). Then

    c=ProjK1(1k6h(u)w)ands=ProjK2(1k7uz), (6.8)

    where (w,z)X×X is the solution to the adjoint system (6.7).

    Let us conclude our analysis by expressing the optimal control in the easiest possible form, recalling that the projection of vL2(QT) into

    K={xL2(QT):0xb a.e. in QT}

    is

    ProjK(v)={0 if v<0,v if 0vb,b if v>b.

    As a result, the optimal control is characterized by the following two formulas

    c={0 if h(u)w<0,1k6h(u)w if 01k6h(u)wUmax,Umax if 1k6h(u)w>Umax,

    and

    s={0 if uz<0,1k7uz if 01k7uzSmax,Smax if 1k7uz>Smax.

    In classical control theory (see e.g. [25]), the more feasible expression of the first order optimality condition relies on the solutions to a suitable linear problem and its adjoint. We briefly describe the main tools in a suitable form to treat the model under study in this paper. We consider the two linear systems

    (L){ytΔy+c0y=bv,ny=0,y(0)=0,(L){ptΔp+cT0p=aQ,np=0,p(T)=aΩ,

    where the unknowns are the vectors y=(Y,Z)T and p=(w,z)T so that, in particular, Δy=(ΔY,ΔZ)T. Besides, c0 is a 2×2 matrix whose transpose is cT0, while all the other given quantities b,v,aQ,aΩ are vector functions. The second system (L) is called the adjoint to system (L). The link between the two systems, that turns out to be quite useful in order to identify the optimal control, is expressed by (7.1) in the next result.

    Theorem 7.1. Provided that the entries of the matrix c0 and of the vectors b and v belong to L(QT), if aQ[L2(QT)]2 and aΩH×H, then there exists a unique strong solution y[C([0,T],V)L2(0,T;W)]2 to problem (L) such that

    y[C([0,T],V)L2(0,T;W)]2C(v[L2(QT)]2+1).

    Moreover, there exists a unique weak solution pX×X to the adjoint problem (L) such that

    pX×XC.

    Finally, the following equality holds true

    aΩ,y(T)+T0aQ,ydt=T0bv,pdt. (7.1)

    Proof. First of all, we see that problem (L) is well posed: indeed, under our assumptions, the coefficients belong to L(QT), the source terms to L2(QT) and the null initial data are in particular in V×V, hence by classical results on parabolic systems (see, e.g., [15, Theorem 1.1]), there exists a unique strong solution y[C([0,T],V)L2(0,T;W)]2 satisfying (7.1). Reversing time by the change of variable tTt, problem (L) turns into a forward system with L(QT) coefficients, L2(QT) source terms and initial data in H×H. Thus, the aforementioned theorem on linear parabolic system applies, yielding the existence of a unique weak solution pX×X to (L) satisfying (7.1). Since y is a strong solution to problem (L) then it is also a weak solution and, by definition,

    T0yt,ϕdt+T0Δy+c0y,ϕdt=T0bv,ϕdt

    for every ϕX×X. Choosing ϕ=p as weak solution to problem (L) and integrating by parts, we obtain

    T0yt,pdt=y(T),p(T)+T0pt,ydt.

    Besides,

    Δy+c0y,p=Δp+cT0p,y.

    We thus find

    y(T),p(T)+T0pt,ydt+T0Δp+cT0p,y=T0bv,pdt.

    Collecting our computations, we obtain (7.1).

    All authors declare no conflicts of interest in this paper.



    [1] A. C. Aristotelous, O. A. Karakashian, S. M. Wise, Adaptive, second-order in time, primitive-variable discontinuous Galerkin schemes for a Cahn-Hilliard equation with a mass source, IMA J. Numer. Anal., 35 (2015), 1167–1198. doi: 10.1093/imanum/dru035.
    [2] A. Aubert, R. Costalat, Interaction between astrocytes and neurons studied using a math- ematical model of compartmentalized energy metabolism, J. Cereb. Blood Flow Metab., 25 (2005), 1476–1490. doi: 10.1038/sj.jcbfm.9600144.
    [3] P. Colli, H. Gomez, G. Lorenzo, G. Marinoschi, A. Reali, E. Rocca, Mathematical analysis and simulation study of a phase-field model of prostate cancer growth with chemotherapy and antiangiogenic therapy effects, Math. Models Methods Appl. Sci., 30 (2020), 1253–1295. doi: 10.1142/S0218202520500220. doi: 10.1142/S0218202520500220
    [4] P. Colli, H. Gomez, G. Lorenzo, G. Marinoschi, A. Reali, E. Rocca, Optimal control of cytotoxic and antiangiogenic therapies on prostate cancer growth, preprint.
    [5] P. Colli, A. Signori, J. Sprekels, Optimal control of a phase field system modelling tumor growth with chemotaxis and singular potentials, Appl. Math. Optim., 83 (2021), 2017–2049. doi: 10.1007/s00245-019-09618-6.
    [6] M. Conti, S. Gatti, A. Miranville, Mathematical analysis of a model for proliterative-to-invasive transition of hypoxic glioma cells, Nonlinear Analysis, 189 (2019), 111572. doi: 10.1016/j.na.2019.111572. doi: 10.1016/j.na.2019.111572
    [7] H. S. Friedman, M. D. Prados, P. Y. Wen, T. Mikkelsen, D. Schiff, L. E. Abrey, et al., Bevacizumab alone and in combination with irinotecan in recurrent glioblastoma, J. Clin. Oncol., 27 (2009), 4733–4740. doi: 10.1200/JCO.2008.19.8721.
    [8] H. Garcke, K. F. Lam, Well-posedness of a Cahn-Hilliard system modelling tumour growth with chemotaxis and active transport, Europ. J. Appl. Math., 28 (2016), 284–316. doi: 10.1017/S0956792516000292.
    [9] H. Garcke, K. F. Lam, R. Nurnberg, E. Sitka, A multiphase Cahn-Hilliard-Darcy model for tumour growth with necrosis, Math. Models Methods Appl. Sci., 28 (2018), 525–577. doi: 10.1142/S0218202518500148.
    [10] H. Garcke, K. F. Lam, E. Rocca, Optimal control of treatment time in a diffuse interface model of tumor growth, Appl. Math. Optim., 78 (2017), 495–544. doi: 10.1007/s00245-017-9414-4.
    [11] H. Garcke, K. F. Lam, A. Signori, Sparse optimal control of a phase field tumour model with mechanical effects, SIAM J. Control Optim., 59 (2021), 1555–1580. doi: 10.1137/20M1372093.
    [12] H. Garcke, K. F. Lam, E. Sitka, V. Styles, A Cahn-Hilliard-Darcy model for tumour growth with chemotaxis and active transport, Math. Models Methods Appl. Sci., 26 (2016), 1095–1148. doi: 10.1142/S0218202516500263.
    [13] H. Gomez, Quantitative analysis of the proliferative-to-invasive transition of hypoxic glioma cells, Integr. Biol., 9 (2017), 257–262. doi: 10.1039/c6ib00208k.
    [14] P. Hartman, Ordinary differential equation, Corrected reprint of the second (1982) edition, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2002.
    [15] J. L. Lions, Équations Différentielles Opérationnelles at Problémes aux Limites non Linéares, Springer-Verlag, Berlin/Göttingen/Heidelberg, Germany, 1961.
    [16] O. A. Ladyzenskaja, V. A. Solonnikov, N. N. Uralceva, Linear and Quasi-linear Equations of Parabolic Type, Vol. 23, Translations of Mathematical Monographs, American Mathematical Society, 1968.
    [17] M. Lahutte-Auboin, Modélisation biomathématique du métabolisme énergétique cérébral : réduction de modèle et approche multi-échelle, application à l'aide à la décision pour la pathologie des gliomes, PhD thesis, Université Pierre et Marie Curie, 2015.
    [18] L. Li, A. Miranville, R. Guillevin, A Cahn-Hilliard model with a proliferation term for the proliferative-to-invasive transition of hypoxic glioma cells, Comm. Math. Sci., to appear.
    [19] A. Miranville, A singular reaction-diffusion equation associated with brain lactate kinetics, Math. Methods Appl. Sci., 40 (2017), 2454–2465. doi: 10.1002/mma.4150.
    [20] A. Miranville, Mathematical analysis of a parabolic-elliptic model for brain lactate kinetics, Solvability, regularity, and optimal control of boundary value problems for PDEs, 379–403, Springer INdAM Ser. 22, Springer, Cham, 2017. doi: 10.1007/978-3-319-64489-9_15.
    [21] A. Miranville, The Cahn-Hilliard equation: recent advances and applications, CBMS-NSF Regional Conference Series in Applied Mathematics 95, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2019. doi: 10.1137/1.9781611975925.
    [22] A. Miranville, The Cahn-Hilliard equation with a nonlinear mass source, J. Diff. Equ., 294 (2021), 88–117. doi: 10.1016/j.jde.2021.05.045.
    [23] A. Miranville, E. Rocca, G. Schimperna, On the long time behavior of a tumor growth model, J. Diff. Equ., 267 (2019), 2616–2642. doi: 10.1016/j.jde.2019.03.028.
    [24] R. Temam, Infinite-dimensional dynamical systems in mechanics and physics, Second edition, Applied Mathematical Sciences, Vol. 68, Springer-Verlag New York, 1997.
    [25] F. Troltzsch, Optimal Control of Partial Differential Equations, Graduate Studies in Mathematics, 112. American Mathematical Society, Providence, RI, 2010. doi: 10.1090/gsm/112.
    [26] J. Xu, G. Vilanova, H. Gomez, A mathematical model coupling tumor growth and angiogenesis, PLoS ONE, 11 (2016), e0149422. doi: 10.1371/journal.pone.0149422.
  • This article has been cited by:

    1. Hussein Raad, Laurence Cherfils, Cyrille Allery, Rémy Guillevin, Optimal control of a model for brain lactate kinetics, 2023, 18758576, 1, 10.3233/ASY-221823
    2. Laurence Cherfils, Stefania Gatti, Carole Guillevin, Alain Miranville, Rémy Guillevin, On a tumor growth model with brain lactate kinetics, 2022, 39, 1477-8599, 382, 10.1093/imammb/dqac010
    3. Niusha Narimani, Mehdi Dehghan, Vahid Mohammadi, A weighted combination of reproducing kernel particle shape functions with cardinal functions of scalable polyharmonic spline radial kernel utilized in Galerkin weak form of a mathematical model related to anti-angiogenic therapy, 2024, 135, 10075704, 108059, 10.1016/j.cnsns.2024.108059
    4. Hardik Joshi, Mehmet Yavuz, Chaotic dynamics of a cancer model with singular and non-singular kernel, 2025, 0, 1937-1632, 0, 10.3934/dcdss.2025016
  • 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(2376) PDF downloads(174) Cited by(4)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog