
Citation: Verónica Anaya, Mostafa Bendahmane, Mauricio Sepúlveda. Mathematical and numerical analysis for Predator-prey system in a polluted environment[J]. Networks and Heterogeneous Media, 2010, 5(4): 813-847. doi: 10.3934/nhm.2010.5.813
[1] | Vincent Renault, Michèle Thieullen, Emmanuel Trélat . Optimal control of infinite-dimensional piecewise deterministic Markov processes and application to the control of neuronal dynamics via Optogenetics. Networks and Heterogeneous Media, 2017, 12(3): 417-459. doi: 10.3934/nhm.2017019 |
[2] | Simone Göttlich, Stephan Martin, Thorsten Sickenberger . Time-continuous production networks with random breakdowns. Networks and Heterogeneous Media, 2011, 6(4): 695-714. doi: 10.3934/nhm.2011.6.695 |
[3] | Yilun Shang . Group pinning consensus under fixed and randomly switching topologies with acyclic partition. Networks and Heterogeneous Media, 2014, 9(3): 553-573. doi: 10.3934/nhm.2014.9.553 |
[4] | Ciro D'Apice, Olha P. Kupenko, Rosanna Manzo . On boundary optimal control problem for an arterial system: First-order optimality conditions. Networks and Heterogeneous Media, 2018, 13(4): 585-607. doi: 10.3934/nhm.2018027 |
[5] | Xianmin Geng, Shengli Zhou, Jiashan Tang, Cong Yang . A sufficient condition for classified networks to possess complex network features. Networks and Heterogeneous Media, 2012, 7(1): 59-69. doi: 10.3934/nhm.2012.7.59 |
[6] | Charles Bordenave, David R. McDonald, Alexandre Proutière . A particle system in interaction with a rapidly varying environment: Mean field limits and applications. Networks and Heterogeneous Media, 2010, 5(1): 31-62. doi: 10.3934/nhm.2010.5.31 |
[7] | Michael Herty, Lorenzo Pareschi, Sonja Steffensen . Mean--field control and Riccati equations. Networks and Heterogeneous Media, 2015, 10(3): 699-715. doi: 10.3934/nhm.2015.10.699 |
[8] | Guy Barles, Emmanuel Chasseigne . (Almost) Everything you always wanted to know about deterministic control problems in stratified domains. Networks and Heterogeneous Media, 2015, 10(4): 809-836. doi: 10.3934/nhm.2015.10.809 |
[9] | Jésus Ildefonso Díaz, Tommaso Mingazzini, Ángel Manuel Ramos . On the optimal control for a semilinear equation with cost depending on the free boundary. Networks and Heterogeneous Media, 2012, 7(4): 605-615. doi: 10.3934/nhm.2012.7.605 |
[10] | Karoline Disser, Matthias Liero . On gradient structures for Markov chains and the passage to Wasserstein gradient flows. Networks and Heterogeneous Media, 2015, 10(2): 233-253. doi: 10.3934/nhm.2015.10.233 |
Optogenetics is a recent and innovative technique which allows to induce or prevent electric shocks in living tissues, by means of light stimulation. Successfully demonstrated in mammalian neurons in 2005 ([8]), the technique relies on the genetic modification of cells to make them express particular ionic channels, called rhodopsins, whose opening and closing are directly triggered by light stimulation. One of these rhodopsins comes from an unicellular flagellate algae, Chlamydomonas reinhardtii, and has been baptized Channelrodhopsins-2 (ChR2). It is a cation channel that opens when illuminated with blue light.
Since the field of Optogenetics is young, the mathematical modeling of the phenomenon is quite scarce. Some models have been proposed, based on the study of the photocycles initiated by the absorption of a photon. In 2009, Nikolic and al. [33] proposed two models for the ChR2 that are able to reproduce the photocurrents generated by the light stimulation of the channel. Those models are constituted of several states that can be either conductive (the channel is open) or non-conductive (the channel is closed). Transitions between those states are spontaneous, depend on the membrane potential or are triggered by the absorption of a photon. For example, the four-states model of Nikolic and al. [33] has two open states (
The purpose of this paper is to extend to infinite dimension the optimal control of Piecewise Deterministic Markov Processes (PDMPs) and to define an infinite-dimensional controlled Hodgkin-Huxley model, containing ChR2 channels, as an infinite-dimensional controlled PDMP and prove existence of optimal ordinary controls. We now give the definition of the model.
We consider an axon, described as a 1-dimensional cable and we set
The deterministic Hodgkin-Huxley model was introduced in [30]. A stochastic infinite-dimensional model was studied in [4], [10], [27] and [39]. The Sodium (
For a given scale
Definition 1.1. Stochastic controlled infinite-dimensional Hodgkin-HuxleyChR2 model. Let
• A state space
• A control space
• A set of uncontrolled PDEs: For every
{v′(t)=1CmΔv(t)+fd(v(t)),v(0)=v0∈V,v0(x)∈[V−,V+]∀x∈I,v(t,0)=v(t,1)=0,∀t>0, | (1) |
with
D(Δ)=V,fd(v):=1N∑i∈IN(gK1{di=n4}(VK−v(iN))+gNa1{di=m3h1}(VNa−v(iN))+gL(VL−v(iN))+gChR2(1{di=O1}+ρ1{di=O2})(VChR2−v(iN)))δiN, | (2) |
with
• A jump rate function
λd(v,u)=∑i∈IN∑x∈D∑y∈D,y≠xσx,y(v(iN),u)1{di=x}, | (3) |
with
| |
| |
| |
• A discrete transition measure
Q({di:y}|v,d)=σdi,y(v(iN),u)1{di≠y}λd(v,u), | (4) |
where
This model can be physically understood as follows. Between jumps of the stochastic component, the membrane potential evolves according to the Hodgkin-Huxley dynamics with fixed conductances given by the state of the piecewise constant stochastic component (equation (2)). The propagation of the potential along the membrane is governed by the Laplacian term
Remark 1. If the PDEs in Definition 1.1 do not depend on the control variable, the theory developed in this papers addresses optimal control problems where the function
The optimal control problem consists in mimicking an output signal that encodes a given biological behavior, while minimizing the intensity of the light applied to the neuron. For example, it can be a time-constant signal and in this case, we want to change the resting potential of the neuron to study its role on its general behavior. We can also think of pathological behaviors that would be fixed in this way. The minimization of light intensity is crucial because the range of intensity experimentally reachable is quite small and is always a matter of preoccupation for experimenters. These considerations lead us to formulate the following mathematical optimal control problem.
Suppose we are given a reference signal
Jz(α)=Eαz[∫T0(κ||Xαt(ϕ)−Vref||2V+α(Xαt))dt],z∈Υ, | (5) |
where
We will prove the following result.
Theorem 1.2. Under the assumptions of Section 2.1, there exists an optimal control strategy
Jz(α∗)=infα∈AEαz[∫T0(κ||Xαt(ϕ)−Vref||2V+α(Xαt))dt], |
and the value function
Piecewise Deterministic Markov Processes constitute a large class of Markov processes suited to describe a tremendous variety of phenomena such as the behavior of excitable cells ([4], [10], [36]), the evolution of stocks in financial markets ([11]) or the congestion of communication networks ([22]), among many others. PDMPs can basically describe any non diffusive Markovian system. The general theory of PDMPs, and the tools to study them, were introduced by Davis ([18]) in 1984, at a time when the theory of diffusion was already amply developed. Since then, they have been widely investigated in terms of asymptotic behavior, control, limit theorems and CLT, numerical methods, among others (see for instance [9], [14], [15], [17] and references therein). PDMPs are jump processes coupled with a deterministic evolution between the jumps. They are fully described by three local characteristics: the deterministic flow
Optimal control of such processes have been introduced by Vermes ([40]) in finite dimension. In [40], the class of piecewise open-loop controls is introduced as the proper class to consider to obtain strongly Markovian processes. A Hamilton-Jabobi-Bellman equation is formulated and necessary and sufficient conditions are given for the existence of optimal controls. The standard broader class of so-called relaxed controls is considered and it plays a crucial role in getting the existence of optimal controls when no convexity assumption is imposed. This class of controls has been studied, in the finite-dimensional case, by Gamkrelidze ([26]), Warga ([42] and [41]) and Young ([45]). Relaxed controls provide a compact class that is adequate for studying optimization problems. Still in finite dimension, many control problems have been formulated and studied such as optimal control ([25]), optimal stopping ([16]) or controllability ([28]). In infinite dimension, relaxed controls were introduced by Ahmed ([1], [2], [3]). They were also studied by Papageorgiou in [37] where the author shows the strong continuity of relaxed trajectories with respect to the relaxed control. This continuity result will be of great interest in this paper.
A formal infinite-dimensional PDMP was defined in [10] for the first time, the set of ODEs being replaced by a special set of Partial Differential Equations (PDE). The extended generator and its domain are provided and the model is used to define a stochastic spatial Hodgkin-Huxley model of neuron dynamics. The optimal control problem we have in mind here regards those Hodgkin-Huxley type models. Seminal work on an uncontrolled infinite-dimensional Hodgkin-Huxley model was conducted in [4] where the trajectory of the infinite-dimensional stochastic system is shown to converge to the deterministic one, in probability. This type of model has then been studied in [39] in terms of limit theorems and in [27] in terms of averaging. The extension to infinite dimension heavily relies on the fact that semilinear parabolic equations can be interpreted as ODEs in Hilbert spaces.
To give a sense to Definition 1.1 and to Theorem 1.2, we will define a controlled infinite-dimensional PDMP for which the control acts on the three local characteristics. We consider controlled semilinear parabolic PDEs, jump rates
˙x(t)=Lx(t)+f(x(t),u(t)), |
where
minuE∫T0c(x(t),u(t))dt, |
where
To address this optimal control problem, we use the fairly widespread approach that consists in studying the imbedded discrete-time Markov chain composed of the times and the locations of the jumps. Since the evolution between jumps is deterministic, there exists a one-to-one correspondence between the PDMP and a pure jump process that enable to define the imbedded Markov chain. The discrete-time Markov chain belongs to the class of Markov Decision Processes (MDPs). This kind of approach has been used in [25] and [12] (see also the book [31] for a self-contained presentation of MDPs). In these articles, the authors apply dynamic programming to the MDP derived from a PDMP, to prove the existence of optimal relaxed strategies. Some sufficient conditions are also given to get non-relaxed, also called ordinary, optimal strategies. However, in both articles, the PDMP is finite dimensional. To the best of our knowledge, the optimal control of infinite-dimensional PDMPs has not yet been treated and this is one of our main objectives here, along with its motivation, derived from the Optogenetics, to formulate and study infinite-dimensional controlled neuron models.
The paper is structured as follows. In Section 2 we adapt the definition of a standard infinite-dimensional PDMP given in [10] in order to address control problems of such processes. To obtain a strongly Markovian process, we enlarge the state space and we prove an extension to controlled PDMPs of [10,Theorem 4]. We also define in this section the MDP associated to our controlled PDMP and that we study later on. In Section 3 we use the results of [37] to define relaxed controlled PDMPs and relaxed MDPs in infinite dimension. Section 4 gathers the main results of the paper. We show that the optimal control problems of PDMPs and of MDPs are equivalent. We build up a general framework in which the MDP is contracting. The value function is then shown to be continuous and existence of optimal relaxed control strategies is proved. We finally give in this section, some convexity assumptions under which an ordinary optimal control strategy can be retrieved.
The final Section 5 is devoted to showing that the previous theoretical results apply to the model of Optogenetics previously introduced. Several variants of the model are discussed, the scope of the theoretical results being much larger than the model of Definition 1.1.
In the present section we define the infinite-dimensional controlled PDMPs that we consider in this paper in a way that enables us to formulate control problems in which the three characteristics of the PDMP depend on an additional variable that we call the control parameter. In particular we introduce the enlarged process which enable us to address optimization problems in the subsequent sections.
Let
Let
A:={a:(0,T)→U measurable}, |
where
a→∫T0e−tw(t,a(t))dt |
is measurable for all bounded and measurable functions
Between two consecutive jumps of the discrete component, the continuous component
{˙vt=−Lvt+fd(vt,a(t)),v0=v,v∈V. | (6) |
For
P[Tn+1−Tn>Δt|Tn,vTn,dTn]=exp(−∫Tn+ΔtTnλ(ϕas(vTn,dTn),ds,a(s))ds). | (7) |
for
P[dt=d|dt−≠dt]=Q({d}|dt−,vt−,a(t)). | (8) |
The triple (
We will make the following assumptions on the local characteristics of the PDMP.
(H(
1. There exist
δ≤λd(x,z)≤Mλ,∀(x,z)∈H×Z. |
2.
3.
|λd(x,z)−λd(y,z)|≤lλ(K)||x−y||H∀(x,y,z)∈K2×Z. |
(H(
(H(
1.
2.
3.
4.
(H(
1.
||fd(x,z)−fd(y,z)||H≤lf||x−y||H∀(x,z)∈H×Z,lf>0. |
2.
Let us make some comments on the assumptions above. Assumption (H(
(H(f))': For every
1.
2.
In particular, assumption (H(
Finally, assumptions (H(
Theorem 2.1. (see [24,Theorem 4.29])
1. For a strongly continuous semigroup
(a)
(b)
2. Let
We define
ϕat(v,d)=S(t)v+∫t0S(t−s)fd(ϕas(v,d),a(s))ds, | (9) |
with
Definition 2.2. We define the sets
Definition 2.3. Control strategies. Enlarged controlled PDMP. Survival function.
a) The set
A:={α:Υ→Uad([0,T];U) measurable}. |
b) On
•
•
•
•
c) Let
ddtχat(z)=−χat(z)λd(ϕat(z),a(t)), χa0(z)=1, |
and its immediate extension
P[T1>t]=χαt(z). |
The notation
ϕat(z):=S(t)v+∫t0S(t−s)fd(ϕas(z),a(s))ds. |
and
Remark 2. ⅰ) Thanks to [46,Lemma 3], the set of admissible control strategies is in bijection with
{α:Υ×[0,T]→U measurable}, |
and thus can be seen as a set of measurable feedback controls acting on
ⅱ) In view of Definition 2.3, given
{˙vt=−Lvt+fd(vt,α(v,d,s)(τt)), vs=v∈E,˙dt=0, ds=d∈D,˙τt=1, τs=0,˙ht=0, hs=s∈[0,T],˙νt=0, νs=vs=v, | (10) |
with
ⅲ) If the relation
ⅳ) Because of the special definition of the enlarged process, for every control strategy in
Definition 2.4. Space of coherent initial points.
Take
ϕαt(x):=S(t)v0+∫t0S(t−s)fd0(ϕαs(x),α(v0,d0,h0)(τs))ds |
The set
Ξα:={(v,d,τ,h,ν)∈Ξ∣v=ϕατ(ν,d,0,h,ν)}. | (11) |
Then we have for all
ϕαt(x):=S(t)v0+∫t0S(t−s)fd0(ϕαs(x),α(ν0,d0,h0)(τs))ds | (12) |
Note that
Proposition 1. The flow property.
Take
Based on equation (12) and the definition of
Notation. Let
Up to now, thanks to Definition 2.3, we can formally associate the PDMP
Theorem 2.5. Assume that assumptions (H(λ)), (H(
a) There exists a filtered probability space satisfying the usual conditions such that for every control strategy
b) For every compact set
supt∈[0,T]||vαt||H≤cK. |
The proof of Theorem 2.5 is given in Appendix B. In the next section, we introduce the MDP that will allow us to prove the existence of optimal strategies.
Because of the particular definition of the state space
Q′(B×C×E|z,a)=∫T−h0ρtdt, | (13) |
for any
ρt:=λd(ϕat(z),a(t))χat(z)1E(h+t)1B(ϕat(z))Q(C|ϕat(z),d,a(t)), |
with
Relaxed controls are constructed by enlarging the set of ordinary ones, in order to convexify the original system, and in such a way that it is possible to approximate relaxed strategies by ordinary ones. The difficulty in doing so is twofold. First, the set of relaxed trajectories should not be much larger than the original one. Second, the topology considered on the set of relaxed controls should make it a compact set and, at the same time, make the flow of the associated PDE continuous. Compactness and continuity are two notions in conflict so being able to achieve such a construction is crucial. Intuitively a relaxed control strategy on the action space
Notation and reminder.
Let
{t→γ(t,C) is measurable for all C∈B(Z),γ(t,⋅)∈M1+(Z) for all t∈[0,T]. |
We will denote by
Recall that we consider the PDE (6):
˙vt=Lvt+fd(vt,a(t)), v0=v, v∈V, a∈Uad([0,T],U). | (14) |
The relaxed PDE is then of the form
˙vt=Lvt+∫Zfd(vt,u)γ(t)(du), v0=v, v∈V, γ∈R([0,T],U), | (15) |
where
γ→∫T0∫Zf(t,z)γ(t)(dz)dt∈R, |
for every Carathodory integrand
{t→f(t,z) is measurable for all z∈Z,z→f(t,z) is continuous a.e., |f(t,z)|≤b(t) a.e., with b∈L1((0,T),R). |
This topology is called the weak topology on
Finally, by Alaoglu's Theorem,
For the same reasons why (14) admits a unique solution, by setting
Theorem 3.1. If assumptions (H(
a) the space of relaxed trajectories (i.e. solutions of 15) is a convex, compact set of
b) The mapping that maps a relaxed control to the solution of (15) is continuous from
First of all, note that since the control acts on all three characteristics of the PDMP, convexity assumptions on the fields
Definition 3.2. Relaxed control strategies, relaxed local characteristics.
a) The set
AR:={μ:Υ→R([0,T];U) measurable}. |
Given a relaxed control strategy
b) For
{λd(v,γ):=∫Zλd(v,u)γ(du),Q(C|v,d,γ):=(λd(v,γ))−1∫Zλd(v,u)Q(C|v,d,u)γ(du), | (16) |
the expression for the enlarged process being straightforward. This allows us to give the relaxed survival function of the PDMP and the relaxed mild formulation of the solution of (15)
{ddtχμt(z)=−χμt(z)λd(ϕμt(z),μzt),χμ0(z)=1,ϕμt(z)=S(t)v+∫t0∫ZS(t−s)fd(ϕμs(z),u)μzs(du)ds, | (17) |
for
{χγt(z)=exp(−∫t0λd(ϕγs(z),γ(s))ds),ϕγt(z)=S(t)v+∫t0∫ZS(t−s)fd(ϕγs(z),u)γ(s)(du)ds, |
The following proposition is a direct consequence of Theorem 2.5b).
Proposition 2. For every compact set
supt∈[0,T]||vμt||H≤cK. |
The relaxed transition measure is given in the next section through the relaxed stochastic kernel of the MDP associated to our relaxed PDMP.
Let
Q′(B×C×E|z,γ)=∫T−h0˜ρtdt, | (18) |
for Borel sets
˜ρt:=χγt(z)1E(h+t)1B(ϕγt(z))∫Zλd(ϕγt(z),u)Q(C|ϕγt(z),d,u)γ(t)(du),=χγt(z)1E(h+t)1B(ϕγt(z))λd(ϕγt(z),γ(t))Q(C|ϕγt(z),d,γ(t)) |
and
Here, we are interested in finding optimal controls for optimization problems involving infinite-dimensional PDMPs. For instance, we may want to track a targeted "signal" (as a solution of a given PDE, see Section 5). To do so, we are going to study the optimal control problem of the imbedded MDP defined in Section 2.3. This strategy has been for example used in [12] in the particular setting of a decoupled finite-dimensional PDMP, the rate function being constant.
Thanks to the preceding sections we can consider ordinary or relaxed costs for the PDMP
(H(
c(v,u)=a||v||2H+bˉd(0,u)2+c||v||Hˉd(0,u)+d||v||H+eˉd(0,u)+f,g(v)=h||v||2H+i||v||H+j, |
with
Remark 3. This assumption might seem a bit restrictive, but it falls within the framework of all the applications we have in mind. More importantly, it can be widely loosened if we slightly change the assumptions of Theorem 4.3. In particular, all the following results, up to Lemma 4.14, are true and proved for continuous functions
Definition 4.1. Ordinary value function for the PDMP
For
Vα(z):=Eαz[∫Thc(Xαs(ϕ),α(Xαs))ds+g(XαT(ϕ))],z:=(v,d,h)∈Υ, | (19) |
V(z)=infα∈AVα(z),z∈Υ. | (20) |
Assumption (H(c)) ensures that
Definition 4.2. Relaxed value function for the PDMP
For
Vμ(z):=Eμz[∫Th∫Zc(Xμs(ϕ),u)μ(Xμs)(du)ds+g(XμT(ϕ))],z:=(v,d,h)∈Υ, | (21) |
˜V(z)=infμ∈ARVμ(z),z∈Υ. | (22) |
We can now state the main result of this section.
Theorem 4.3. Under assumptions (H(
˜V(z)=Vμ∗(z), ∀z∈Υ. |
Remark 4. All the subsequent results that lead to Theorem 4.3 would be easily transposable to the case of a lower semicontinuous cost function. We would then obtain a lower semicontinuous value function.
The next section is dedicated to proving Theorem 4.3 via the optimal control of the MDP introduced before. Let us briefly sum up what we are going to do. We first show that the optimal control problem of the PDMP is equivalent to the optimal control problem of the MDP and that an optimal control for the latter gives an optimal control strategy for the original PDMP. We will then build up a framework, based on so called bounding functions (see [12]), in which the value function of the MDP is the fixed point of a contracting operator. Finally, we show that under the assumptions of Theorem 4.3, the relaxed PDMP
Let us define the ordinary cost
c′(z,a):=∫T−h0χas(z)c(ϕas(z),a(s))ds+χaT−h(z)g(ϕaT−h(z)), | (23) |
and
Assumption (H(c)) allows
c′(z,γ)=∫T−h0χγs(z)∫Zc(ϕγs(z),u)γ(s)(du)ds+χγT−h(z)g(ϕγT−h(z)), | (24) |
and
Definition 4.4. Cost and value functions for the MDP
For
Jα(z)=Eαz[∞∑n=0c′(Z′n,α(Z′n))],Jμ(z)=Eμz[∞∑n=0c′(Z′n,μ(Z′n))],J(z)=infα∈AJα(z),J′(z)=infμ∈ARJμ(z), |
for
The finiteness of these sums will by justified later by Lemma 4.9.
In the following theorem we prove that the relaxed expected cost function of the PDMP equals the one of the associated MDP. Thus, the value functions also coincide. For the finite-dimensional case we refer the reader to [19] or [12] where the discrete component of the PDMP is a Poisson process and therefore the PDMP is entirely decoupled. The PDMPs that we consider are fully coupled.
Theorem 4.5. The relaxed expected costs for the PDMP and the MDP coincide:
Remark 5. Since we have
Proof. Let
Vμ(z)=Eμz[∞∑n=0∫T∧Tn+1T∧Tn∫Zc(Xμs(ϕ),u)μns−Tn(du)ds+1{Tn≤T<Tn+1}g(XμT(ϕ))]=∞∑n=0Eμz[Eμz[∫T∧Tn+1T∧Tn∫Zc(Xμs(ϕ),u)μns−Tn(du)ds+1{Tn≤T<Tn+1}g(XμT(ϕ))|Hn]], |
all quantities being non-negative. We want now to examine the two terms that we call
I1:=Eμz[∫T∧Tn+1T∧Tn∫Zc(Xμs(ϕ),u)μns−Tn(du)ds|Hn] |
that we split according to
I1=1{Tn≤T}Eμz[∫TTn∫Zc(Xμs(ϕ),u)μns−Tn(du)1{Tn+1>T}ds|Hn]+Eμz[1{Tn+1≤T}∫Tn+1Tn∫Zc(Xμs(ϕ),u)μns−Tn(du)ds|Hn]. |
By the strong Markov property and the flow property, the first term on the RHS is equal to
1{Tn≤T}Eμz[∫T−Tn0∫Zc(XμTn+s(ϕ),u)μns(du)1{Tn+1−Tn>T−Tn}ds|Hn]=1{Tn≤T}χμT−Tn(Zn)∫T−Tn0∫Zc(ϕμs(Zn),u)μns(du)ds. |
Using the same arguments, the second term on the RHS of
1{Tn≤T}∫T−Tn0∫Zλdn(ϕμt(Zn),u)μnt(du)χμt(Zn)∫t0∫Zc(ϕμs(Zn),u)μnt(du)dsdt, |
An integration by parts yields
I1=1{Tn≤T}∫T−Tn0χμt(Zn)∫Zc(ϕαt(Zn),u)μnt(du)dt. |
Moreover
I2:=Eμz[1{Tn≤T<Tn+1}g(XμT)|Hn]=1{Tn≤T}χμT−Tn(Zn)g(ϕμT−Tn(Zn)) |
By definition of the Markov chain
Vμ(z)=∞∑n=0Eμz[1{Tn≤T}∫T−Tn0χμt(Zn)∫Zc(ϕαt(Zn),u)μnt(du)dt+1{Tn≤T}χμT−Tn(Zn)g(ϕμT−Tn(Zn))]=Eμz[∞∑n=0c′(Z′n,μ(Z′n))]=Jμ(z). |
We now show existence of optimal relaxed controls under a contraction assumption. We use the notation
•
•
•
•
•
The classical way to address the discrete-time stochastic control problem that we introduced in Definition 4.4 is to consider an additional control space that we will call the space of Markovian policies and denote by
Jπ(z):=Eπz[∞∑n=0c′(Z′n,μn(Z′n))]. |
Now denote by
J∗(z)=J′(z). |
Let us now define some operators that will be useful for our study and state the first theorem of this section. Let
Rw(z,γ):=c′(z,γ)+(Q′w)(z,γ),Tμw(z):=c′(z,μ(z))+(Q′w)(z,μ(z))=Rw(z,μ(z)),(Tw)(z):=infγ∈R{c′(z,γ)+(Q′w)(z,γ)}=infγ∈RRw(z,γ), |
where
∫T−h0χγt(z)∫Zλd(ϕγt(z),u)∫Dw(ϕγt(z),r,h+t)Q(dr|ϕγt(z),d,u)γ(t)(du)dt. |
Theorem 4.6. Assume that there exists a subspace
J′(z)=Jμ∗(z), ∀z∈Υ. |
All the results needed to prove this Theorem can be found in [6]. We break down the proof into the two following elementary propositions, suited to our specific problem. Before that, recall that from [6,Proposition 9.1 p.216],
Let us now consider the
Jnπ(z):=Eπz[n−1∑i=0c′(Z′i,μi(Z′i))] Jn(z):=infπ∈ΠEπz[n−1∑i=0c′(Z′i,μi(Z′i))] |
for
Proposition 3. Let assumptions of Theorem 4.3 hold. Let
Jn(z)=infπ∈Π(Tμ0Tμ1…Tμn−10)(z)=(Tn0)(z), |
with
Proof. The first relation is straightforward since all quantities defining
Let
Tμn−1I≤TI+ε, Tμn−2TI≤TTI+ε,…, Tμ0Tn−1I≤TTn−1I+ε. |
We then get
TnI≥Tμ0Tn−1I−ε≥Tμ0Tμ1Tn−2I−2ε≥⋯≥Tμ0Tμ1…Tμn−1I−nε≥infπ∈ΠTμ0Tμ1…Tμn−1I−nε. |
Since this last inequality is true for any
TnI≥infπ∈ΠTμ0Tμ1…Tμn−1I, |
and by definition of
TnI≤Tμ0Tμ1…Tμn−1I. |
Finally,
Proposition 4. There exists
Proof. By definition, for every
The next section is devoted to proving that the assumptions of Theorem 4.6 are satisfied for the MDP.
The concept of bounding function that we define below will ensure that the operator
Definition 4.7. Bounding functions for a PDMP.
Let
(ⅰ)
(ⅱ)
(ⅲ)
Given a bounding function for the PDMP we can construct one for the MDP with or without relaxed controls, as shown in the next lemma (cf. [13,Definition 7.1.2 p.195]).
Lemma 4.8. Let
c′(z,γ)≤Bζ(z)cϕ(ccδ+cg), | (25) |
∫ΥBζ(y)Q′(dy|z,γ)≤Bζ(z)cϕMλ(ζ+δ). | (26) |
Proof. Take
c′(z,γ)≤∫T−h0e−δscccϕb(v)ds+e−δ(T−h)cgcϕb(v) ≤Bζ(z)e−ζ(T−h)cϕ(cc1−e−δ(T−h)δ+e−δ(T−h)cg), |
which immediately implies (25). On the other hand
∫ΥBζ(y)Q′(dy|z,γ)=∫T−h0χγs(z)b(ϕγs(z))eζ(T−h−s)∫Zλd(ϕγs(z),u)Q(D|ϕγs(z),u)γs(du)ds≤eζ(T−h)b(v)cϕMλe−ζτ∫T−h0e−δse−ζsds=Bζ(z)cϕMλζ+δ(1−e−(ζ+δ)(T−h)) |
which implies (26).
Let
L∗:={v:Υ→R continuous ;||v||∗:=supz∈Υ|v(z)||B∗(z)|<∞}. | (27) |
The following two lemmas give an estimate on the expected cost of the MDP that justifies manipulations of infinite sums.
Lemma 4.9. The inequality
Proof. We proceed by induction on
Eγz[B∗(Z′k+1)]=Eγz[Eγz[B∗(Z′k+1)|Z′k]] =Eγz[∫ΥB∗(y)Q′(dy|Z′k,γ)] =Eγz[B∗(Z′k)∫ΥB∗(y)Q′(dy|Z′k,γ)B∗(Z′k)]. |
Using (26) and the definition of
Lemma 4.10. There exists
Eμz[∞∑k=nc′(Z′k,μ(Z′k))]≤κCn1−CB∗(z). |
Proof. The results follows from Lemma 4.9 and from the fact that
c′(Z′k,μ(Z′k))≤B∗(Zk)cϕ(ccδ+cg) |
for any
We now state the result on the operator
Lemma 4.11.
||Tv−Tw||B∗≤C||v−w||B∗, |
where
Proof. We prove here the contraction property. The fact
supγ∈Rf(γ)−supγ∈Rg(γ)≤supγ∈R(f(γ)−g(γ)). |
Moreover since
Tv(z)−Tw(z)≤supγ∈R∫T−h0χγs(z)∫Zλd(ϕγs(z),u)I(u,s)γ(s)(du)ds, |
where
I(u,s):=∫D(v(ϕγs(z),r,h+s)−w(ϕγs(z),r,h+s))Q(dr|ϕγs(z),d,u), |
so that
||Tv−Tw||B∗≤sup(z,γ)∈Υ×R∫T−h0χγs(z)∫Zλd(ϕγs(z),u)J(s,u)γ(s)(du)ds |
where
J(s,u):=∫DB∗(ϕγs(z),r,h+s)B∗(z)||v−w||B∗Q(dr|ϕγs(z),d,u) |
We then conclude that
||Tv−Tw||B∗≤sup(z,γ)∈Υ×R∫T−h0e−δsMλcϕe−ζ∗sds||v−w||B∗≤Mλcϕ||v−w||B∗∫T−h0e−(δ+ζ∗)sds≤C||v−w||B∗. |
Here we prove that the trajectories of the relaxed PDMP are continuous w.r.t. the control and that the operator
Lemma 4.12. Assume that (H(
ϕ:(z,γ)∈Υ×R→ϕγ⋅(z)=S(0)v+∫⋅0∫ZS(⋅−s)fd(ϕγs(z),u)γ(s)(du)ds |
is continuous from
Proof. This proof is based on the result of Theorem 3.1. Here we add the joint continuity on
ϕγnt(zn)−ϕγt(z)=S(t)vn−S(t)v+∫t0∫ZS(t−s)fd(ϕγnt(zn),u)γn(s)(du)ds−∫t0∫ZS(t−s)fd(ϕγt(z),u)γ(s)(du)ds=S(t)vn−S(t)v+∫t0∫ZS(t−s)[fd(ϕγnt(zn),u)γn(s)(du)−fd(ϕγt(z),u)γn(s)(du)]ds+∫t0∫ZS(t−s)[fd(ϕγt(z),u)γn(s)(du)−fd(ϕγt(z),u)γ(s)(du)]ds. |
From(H(
||ϕγnt(zn)−ϕγt(z)||H≤MS||vn−v||H+MSlf∫t0||ϕγns(zn)−ϕγs(z)||Hds+||ℓn(t)||H |
where
||ϕγnt(zn)−ϕγt(z)||H≤C(||vn−v||H+sups∈[0,T]||ℓn(s)||H). |
Since
Let us denote by
(h,ℓn(t))H=∫t0∫Z(h,S(t−s)fd(ϕγt(z),u)))Hγn(s)(du)ds−∫t0∫Z(h,S(t−s)fd(ϕγt(z),u)))Hγ(s)(du)ds→n→∞0, |
since
The next lemma establishes the continuity property of the operator
Lemma 4.13. Suppose that assumptions (H(
(z,γ)→∫T−h0χγs(z)(∫Zw(ϕγs(z),d,h+s,u)γ(s)(du))ds |
is continuous on
(z,γ)→Rw(z,γ)=c′(z,γ)+Q′w(z,γ) |
is continuous on
Proof. See Appendix C.
It now remains to show that there exists a bounding function for the PDMP. This is the result of the next lemma.
Lemma 4.14. Suppose assumptions (H(
b(v):={max||x||H≤M3maxu∈U˜c(x,u)+max||x||H≤M3˜g(x), if ||v||H≤M3,maxu∈U˜c(v,u)+˜g(v), if ||v||H>M3, | (28) |
is a continuous bounding function for the PDMP.
Proof. For all
• If
• If
b(ϕγt(z)))=maxu∈U˜c(ϕγt(z),u)+˜g(ϕγt(z))≤b(M3M2v)≤M23M22b(v), |
since
Remark 6. Lemma 4.14 ensures the existence of a bounding function for the PDMP. To broaden the class of cost functions considered, we could just assume the existence of a bounding for the PDMP in Theorem 4.3 and then, the assumption on
Ordinary strategies are of crucial importance because they are the ones that the controller can implement in practice. Here we give convexity assumptions that ensure the existence of an ordinary optimal control strategy for the PDMP.
(A) (a) For all
(b) For all
(c) The cost function
Theorem 4.15. Suppose that assumptions (H(
Proof. This result is based on the fact that for all
∫Zλd(ϕγs(z),u)γ(s)(du)≤λd(ϕˉγs(z),ˉγ(s)),∫Zλd(ϕγs(z),u)Q(E|ϕγs(z),d,u)γ(s)(du)≥λd(ϕˉγs(z),ˉγ(s))Q(E|ϕˉγs(z),d,ˉγ(s)),∫Zc(ϕγs(z),u)γs(du)≥c(ϕˉγs(z),¯γs), |
for all
(Lw)(z,γ)=∫T−h0χγs(z)∫Zc(ϕγs(z),u)γ(s)(du)ds+χγT−h(z)g(ϕγT−h(z))+∫T−h0χγs(z)∫Zλd(ϕγs(z),u)∫Dw(ϕγs(z),r,h+s)Q(dr|ϕγs(z),d,u)γ(s)(du)ds≥∫T−h0χˉγs(z)c(ϕˉγs(z),ˉγ(s))ds+χˉγT−h(z)g(ϕˉγT−h(z))+∫T−h0χˉγs(z)∫Zλd(ϕˉγs(z),u)∫Dw(ϕˉγs(z),r,h+s)Q(dr|ϕˉγs(z),d,u)γ(s)(du)ds. |
Furthermore,
∫Zλd(ϕˉγs(z),u)∫Dw(ϕˉγs(z),r,h+s)Q(dr|ϕˉγs(z),d,u)γ(s)(du)≥λd(ϕˉγs(z),ˉγ(s))∫Dw(ϕˉγs(z),r,h+s)Q(dr|ϕˉγs(z),d,ˉγ(s)), |
so that
(Lw)(z,γ)≥∫T−h0χˉγs(z)c(ϕˉγs(z),ˉγ(s))ds+χˉγT−h(z)g(ϕˉγT−h(z))+∫T−h0χˉγs(z)λd(ϕˉγs(z),ˉγ(s))∫Dw(ϕˉγs(z),r,h+s)Q(dr|ϕˉγs(z),d,ˉγ(s))=(Lw)(z,ˉγ). |
Here we treat an elementary example that satisfies the assumptions made in the previous two sections.
Let
(v,w)V:=∫10v(x)w(x)+v′(x)w′(x)dx. |
We consider the following PDE for the deterministic evolution between jumps
∂∂tv(t,x)=Δv(t,x)+(d+u)v(t,x), |
with Dirichlet boundary conditions. We define the jump rate function for
λ1(v,u)=1e−||v||2+1+u2, λ−1(v,u)=e−1||v||2+1+u2, |
and the transition measure by
Finally, we consider a quadratic cost function
Lemma 4.16. The PDMP defined above admits the continuous bounding function
b(v):=||Vref||2H+||v||2H+1. | (29) |
Furthermore, the value function of the optimal control problem is continuous and there exists an optimal ordinary control strategy.
Proof. The proof consists in verifying that all assumptions of Theorem 4.15 are satisfied. Assumptions (H(
⟨−Δ(v−w),v−w⟩=∫10((v−w)′(x))2dx≥0. |
|⟨−Δv,w⟩|2=|∫10v′(x)w′(x)dx|2≤∫10(v′(x))2dx∫10(w′(x))2dx≤||v||2V||w||2V, |
and so
Now, define for
On
S(t)v=∑k≥1e−(kπ)2t(v,fk)Hfk. |
For
For
||fd(v,u)−fd(w,u)||H≤2||v−w||H, ||fd(v,u)||H≤2||v||H. |
This means that for every
We begin this section by making some comments on Definition 1.1.
In (1),
φ(x):={Ce1x2−1, if |x|<1,0, if |x|≥1, | (30) |
with
C:=(∫1−1exp(1x2−1)dx)−1 |
such that
Now, let
ξNz(x):={φN(x−z), if x∈UN0, if x∈[0,1]∖UN. | (31) |
For all
The following lemma is a direct consequence of [10,Proposition 7] and will be very important for the model to fall within the theoretical framework of the previous sections.
Lemma 5.1. For every
V−≤y(t,x)≤V+, ∀x∈I. |
Physiologically speaking, we are only interested in the domain
The next lemma shows that the stochastic controlled infinite-dimensional Hodgkin-Huxley-ChR2 model defines a controlled infinite-dimensional PDMP as defined in Definition 2.3 and that Theorem 2.5 applies.
Lemma 5.2. For
V−≤vαt(x)≤V+, ∀(t,x)∈[0,T]×I. |
Proof. The local Lipschitz continuity of
fd(y1)−fd(y2)=1N∑i∈IN(gK1{di=n4}+gNa1{di=m3h1}+gChR2(1{di=O1}+ρ1{di=O2})+gL)(ξNiN,y2−y1)HξNiN. |
We then get
||fd(y1)−fd(y2)||H≤4N2(gK+gNa+gChR2(1+ρ)+gL)||(y2−y1)||H. |
Finally, since the continuous component
Proof of Theorem 1.2. In Lemma 5.2 we already showed that assumptions (H(
We end this section with an important remark that significantly extends the scope of this example. Up to now, we only considered stationary reference signals but nonautonomous ones can be studied as well, as long as they feature some properties. Indeed, it is only a matter of incorporating the signal reference
This way, the part on the control problem is not impacted at all and we consider the continuous cost function
˜c(v,ˉv,u)=κ||v−ˉv||2H+u+cmin, | (32) |
the result and proof of lemma 1.2 remaining unchanged with the continuous bounding function defined for
b(v):={κM23+κsupt∈[0,T]||Vref(t)||2H+umax,if ||v||H≤M3,κ||v||2H+κsupt∈[0,T]||Vref(t)||2H+umax,if ||v||H>M3. |
In the next section, we present some variants of the model and the corresponding results in terms of optimal control.
We begin this section by giving arguments showing that the results of Theorem 4.3 remain valid for the model of Definition 1.1, which does not exactly fits into our theoretical framework. Then, the variations we present concern the model of ChR2, the addition of other light-sensitive ionic channels, the way the control acts on the three local characteristics and the control space. The optimal control problem itself will remain unchanged. First of all, let us mention that since the model of Definition 1.1 satisfies the convexity conditions (A), the theoretical part on relaxed controls is not necessary for this model. Nevertheless, the model of ChR2 presented on Figure 1 is only one among several others, some of which do not enjoy a linear, or even concave, rate function
We will not present them here, but the previous results for the Hodgkin-Huxley model remain straightforwardly unchanged for other neuron models such as the FitzHugh-Nagumo model or the Morris-Lecar model.
Optimal control for the original model.
In the original model, the function
|v(iN)−w(iN)|≤supx∈I|v(x)−w(x)|≤C||v−w||V. |
Finally, [10,Proposition 5] states that the bounds of Lemma 5.2 remain valid with Dirac masses.
Modifications of the ChR2 model.
We already mentioned the paper of Nikolic and al. [33] in which a three states model is presented. It is a somehow simpler model that the four states model of Figure 1 but it gives good qualitative results on the photocurrents produced by the ChR2. In first approximation the model can be considered to depend linearly in the control as seen on Figure 2.
This model features one open state
Some mathematical comments are needed here. On Figure 3, the control
limu→0Kr+c2log(u)=−∞, limu→01τd−log(u)=0. |
The first limit is not physical since rate jumps between states are positive numbers. The second limit is not physical either because it would mean that, in the dark, the proteins are trapped in the open state
The four states model of Figure 1 is also an approximation of a more accurate model that we represent on Figure 4 below. The transition rates can depend on either the membrane potential
Kd1(v)=K(1)d1−K(2)d1tanh((v+20)/20), |
e12(u)=e12d+c1ln(1+u/c),e21(u)=e21d+c2ln(1+u/c),Kr(v)=K(1)rexp(−K(2)rv), |
with
Addition of other light-sensitive ion channels.
Channelrhodopsin-2 has a promoting role in eliciting action potentials. There also exists a chlorine pump, called Halorhodopsin (NpHR), that has an inhibitory action. NpHR can be used along with ChR2 to obtain a control in both directions. Its modelisation as a multistate model was considered in [34]. The transition rates between the different states have the same shape that the ones of the ChR2 and the same simplifications are possible. This new light-sensitive channel can be easily incorporated in our stochastic model and we can state existence of optimal relaxed and/or ordinary control strategies depending on the level of complexity of the NpHR model we consider. It is here important to remark that since the two ionic channels do not react to the same wavelength of the light, the resulting control variable would be two-dimensional with values in
Modification of the way the control acts on the local characteristics.
Up to now, the control acts only on the rate function, and also on the measure transition via its special definition from the rate function. Nevertheless, we can present here a modification of the model where the control acts linearly on the PDE. This modification amounts to considering that the control variable is directly the gating variable of the ChR2. Indeed, we show in [38] that the optimal control of the deterministic counterpart of the stochastic Hodgkin-Huxley-ChR2 model, in finite dimension and with the three states ChR2 model of Figure 2, is closely linked to the optimal control of
{dVdt=gKn4(t)(VK−V(t))+gNam3(t)h(t)(VNa−V(t))+gChR2u(t)(VChR2−V(t))+gL(VL−V(t)),dndt=αn(V(t))(1−n(t))−βn(V(t))n(t),dmdt=αm(V(t))(1−m(t))−βm(V(t))m(t),dhdt=αh(V(t))(1−h(t))−βh(V(t))h(t), |
where the control variable is the former gating variable
Modification of the control space.
In all models discussed previously, the control has no spatial dependence. Any light-stimulation device, such as a laser, has a spatial resolution and it is possible that we do not want or cannot stimulate the entire axon. For this reason, spatial dependence of the control should be considered. Now, as long as the control space remains a compact Polish space, spatial dependence of the control could be considered. We propose here a control space defined as a subspace of the Skorohod space
U:={u:[0,1]→[0,umax]∣u is constant on [i/p,(i+1)/p),i=0,..,p−1,u(1)=u((p−1)/p)}. |
Lemma 5.3.
Proof. We tackle this proof by remarking that
In this case, the introduction of the space
˜U:={u:[0,1]→[0,umax]∣∃{xi}0≤i≤p subdivision of [0,1],u is constant on [xi,xi+1),i=0,..,p−1,u(1)=u(xp−1)}. |
Now
Lemma 5.4.
Proof. For this proof, we need to introduce some notation and a critera of compactness in
Let
wu([xi,xi+1)):=supx,y∈[xi,xi+1)|u(x)−u(y)|, |
and for
w′u(δ):=inf{xi}max0≤i<nwu([xi,xi+1)), |
the infimum being taken on all the subdivisions
limδ→0supu∈˜Uw′u(δ)=0. |
Let
With either
Perspectives.
Theorem 4.3 proves the existence of optimal controls for a wide class of infinite-dimensional PDMPs and Theorem 4.15 gives sufficient conditions to retrieve ordinary optimal controls. Nevertheless, these theorems do not indicate how to compute optimal controls or approximations of optimal controls. Thus, it would be very interesting, in a further work, to implement numerical methods in order to compute at least approximation of optimal controls for the control problems defined in this paper. One efficient way to address numerical optimal control problems for PDMPs is to use quantization methods that consist in replacing the state and control spaces by discrete spaces and work with approximations of the processes on these discrete spaces ([29], [35], [20]).
Let
• Let
Pαx(T1>t)=exp(−∫t0λd(ϕαs(x),α(νs,ds,hs)(τs))ds). |
• For
•When a jump occurs at time
ˆQ({ϕαT1(x)}×B×{0}×{h+τT−1}×{ϕαT1(x)}|ϕαT1(x),dT−1,τT−1,hT−1,νT−1,α(T−1))=Q(B|ϕαT1s(s),d,α(ν,d,h)(τ+T1)), | (33) |
where we use the notation
• The construction iterates after time
Formally the expressions of the jump rate and the transition measures on
λ(x,u):=λd(v,u),ˆQ(F×B×E×G×J|x,u):=1F×E×G×J(v,0,h+τ,ν)Q(B|v,d,u), |
with
There are two filtered spaces on which we can define the enlarged process
Given the sample path
{Tk,k=1,…,n}={s∈(0,T]|hs≠hs−}. |
Moreover we can associate to
Zαt:=(νTk,dTk,Tk), Tk≤t<Tk+1. | (34) |
Conversely, given the sample path of
{Xαt=(ϕαt(Zα0),dZ0,t,TZ0,νZ0),t<T1,Xαt=(ϕαt−Tk(ZαTk),dTk,t−Tk,TZTk,νZTk),Tk≤t<Tk+1. | (35) |
Let us note that
Part 1. The canonical space of jump processes with values in
ω=(γ0,s1,γ1,s2,γ2,…), | (36) |
where
Accordingly we introduce
{Si:ΩΥ⟶R+∪{∞},ω⟼Si(ω)=si,for i∈N∗,Γi:ΩΥ⟶Υ∪{Δ},ω⟼Γi(ω)=γi,for i∈N. | (37) |
We also introduce
ωi(ω):=(Γ0(ω),S1(ω),Γ1(ω),…,Si(ω),Γi(ω)) |
for
T0(ω):=0,Ti(ω):={i∑k=1Sk(ω),if Sk(ω)≠∞ and Γk(ω)≠Δ,k=1,…,i,∞if Sk(ω)=∞ or Γk(ω)=Δ for some k=1,…,i,T∞(ω):=limi→∞Ti(ω). |
and the sample path
xt(ω):={Γi(ω) Ti(ω)≤t<Ti+1(ω),Δ t≥T∞(ω). | (38) |
A relevant filtration for our problem is the natural filtration of the coordinate process
FΥt:=σ{xs|s≤t}, |
for all
μ1(α(γ0);({0}×Υ)∪(R+×{γ0}))=0. |
For
1.
2.
3.
4.
We need to extend the definition of
Now for a given control strategy
∫ΩΥf(ωi(ω))Pαγ0(dω)=∫YΥ1…∫YΥif(y1,…,yi)μi(y1,…,yi−1,α(yi−1);dyi)×μi−1(y1,…,yi−2,α(yi−2);dyi−1)…μ1(α(γ0);dy1), |
with
FΥ:=⋂γ∈Υα∈AFΥγ,α, |
FΥt:=⋂γ∈Υα∈AFΥ,γ,αt for all t≥0. |
Then (
Part 2. The canonical space of cdlg functions with values in
We start with the definition of
χαt(x):=exp(−∫t0λd(ϕαs(x),ατ+s(ν,d,h))ds), |
such that for
This procedure thus provides for each control
FΞt:=⋂α∈A,x∈ΞαFx,αt. | (39) |
The right-continuity of
Now that we have a filtered probability space that satisfies the usual conditions, let us show that the simple Markov property holds for
Pα[Tk+1−Tk>s|FTk]=exp(−∫s0λdTk(ϕαt(XαTk),αu(νTk,dTk,hTk))du)=χαs(XαTk). |
Now for
Pαx[Tk+1>t+s|Ft]1{Tk≤t<Tk+1}=Pαx[Tk+1−Tk>t+s−Tk|Ft]1{0≤t−Tk<Tk+1−Tk}=exp(−∫t+s−Tkt−TkλdTk(ϕαu(XαTk),αu(νTk,dTk,hTk))du)1{0≤t−Tk<Tk+1−Tk}(∗)=exp(−∫s0λdTk(ϕαu+t−Tk(XαTk),αu+t−Tk(νTkdTk,hTk))du)1{0≤t−Tk<Tk+1−Tk}. |
The equality (*) is the classical formula for jump processes (see Jacod [32]). On the other hand,
χαs(Xαt)1{Tk≤t<Tk+1}=exp(−∫s0λdt(ϕαu(Xαt),αu+τt(νt,dt,ht))du)1{Tk≤t<Tk+1}=exp(−∫s0λdTk(ϕαu(Xαt),αu+t−Tk(νTk,dTk,hTk))du)1{Tk≤t<Tk+1}=exp(−∫s0λdTk(ϕαu+t−Tk(XαTk),αu+t−Tk(νTkdTk,hTk))du)1{Tk≤t<Tk+1}, |
because
Thus we showed that for all
Pαx[Tk+1>t+s|Ft]1{Tk≤t<Tk+1}=χαs(Xαt)1{Tk≤t<Tk+1}. |
Now if we write
Pαx[Tαt>t+s|Ft]=χαs(Xαt), | (40) |
which means that, conditionally to
To extend the proof to the strong Markov property, the application of Theorem (25.5) (Davis [19]) on the characterization of jump process stopping times on Borel spaces is straightforward.
From the results of [10], there is no difficulty in finding the expression of the extended generator
• Let
• Let
dfdv[v,d,τ,h,ν](y)=⟨hv(v,d,τ,h,ν),y⟩V∗,V ∀y∈V, |
where
Gαf(v,d,τ,h,ν)=∂∂τf(v,d,τ,h,ν)+⟨hv(v,d,τ,h,ν),Lv+fd(v,ατ(ν,d,h))⟩V∗,V+λd(v,ατ(ν,d,h))∫D[f(v,p,0,h+τ,v)−f(v,d,τ,h,ν)]Qα(dp|v,d). | (41) |
The bound on the continuous component of the PDMP comes from the following estimation. Let
||vαt||H≤||S(t)v||H+∫t0||S(t−s)fds(vαs,ατs(νs,ds,hs))||Hds≤MS||v||H+∫t0MS(b1+b2||vαs||H)ds≤MS(||v||H+b1T)eMSb2T, | (42) |
by Gronwall's inequality.
Part 1. Let's first look at the case when
W(z,γ)=∫T−h0χγs(z)(∫Zw(ϕγs(z),d,h+s,u)γ(s)(du))ds |
Now take
|W(zn,γn)−W(z,μ)|≤|∫bnanχγns(zn)∫Zwn(s,u)γn(s)(du)ds|+∫T−h0χγns(zn)∫Z|wn(s,u)−w(s,u)|γn(s)(du)ds+|∫T−h0χγns(zn)∫Zw(s,u)γn(s)(du)ds−∫T−h0χγs(z)∫Zw(s,u)γn(s)(du)ds|+|∫T−h0χγs(z)∫Zw(s,u)γn(s)(du)ds−∫T−h0χγs(z)∫Zw(s,u)γ(s)(du)ds| |
The first term on the right-hand side converges to zero for
by dominated convergence and the continuity of
again by dominated convergence, provided that for
It is enough to take
By the local Lipschitz property of
and
Part 2. In the general case where
is bounded, continuous, decreasing and converges to
which is thus upper semicontinuous. Since
is continuous so that in fact
Now the continuity of the applications
[1] |
A. S. Ackleh and L. Ke, Existence-uniqueness and long time behavior for a class on nonlocal nonlinear parabolic evolution equations, Proc. Amer. Math. Soc., 128 (2000), 3483-3492. doi: 10.1090/S0002-9939-00-05912-8
![]() |
[2] | V. Anaya, M. Bendahmane and M. Sepúlveda, Mathematical and numerical analysis for reaction-diffusion systems modeling the spread of early tumors, Bol. Soc. Esp. Mat. Apl., (2009), 55-62. |
[3] |
V. Anaya, M. Bendahmane and M. Sepúlveda, A numerical analysis of a reaction-diffusion system modelling the dynamics of growth tumors, Math. Models Methods Appl. Sci., 20 (2010), 731-756. doi: 10.1142/S0218202510004428
![]() |
[4] |
B. Ainseba, M. Bendahmane and A. Noussair, A reaction-diffusion system modeling predator-prey with prey-taxis, Nonlinear Anal. Real World Appl., 128 (2008), 2086-2105. doi: 10.1016/j.nonrwa.2007.06.017
![]() |
[5] |
L. Bai and K. Wang, A diffusive stage-structured model in a polluted environment, Nonlinear Anal. Real World Appl., 7 (2006), 96-108. doi: 10.1016/j.nonrwa.2004.11.010
![]() |
[6] |
M. Bendahmane, K. H. Karlsen and J. M. Urbano, On a two-sidedly degenerate chemotaxis model with volume-filling effect, Math. Models Methods Appl. Sci., 17 (2007), 783-804. doi: 10.1142/S0218202507002108
![]() |
[7] |
M. Bendahmane and M. Sepúlveda, Convergence of a finite volume scheme for nonlocal reaction-diffusion systems modelling an epidemic disease, Discrete Contin. Dyn. Syst. Ser. B, 11 (2009), 823-853. doi: 10.3934/dcdsb.2009.11.823
![]() |
[8] |
M. Chipot and B. Lovat, Some remarks on nonlocal elliptic and parabolic problem, Nonlinear Anal., 30 (1997), 4619-4627. doi: 10.1016/S0362-546X(97)00169-7
![]() |
[9] |
B. Dubey and J. Hussain, Modelling the interaction of two biological species in a polluted environment, J. Math. Anal. Appl., 246 (2000), 58-79. doi: 10.1006/jmaa.2000.6741
![]() |
[10] |
B. Dubey and J. Hussain, Models for the effect of environmental pollution on forestry resources with time delay, Nonlinear Anal. Real World Appl., 5 (2004), 549-570. doi: 10.1016/j.nonrwa.2004.01.001
![]() |
[11] | R. Eymard, Th. Gallouët and R. Herbin, "Finite Volume Methods. Handbook of Numerical Analysis," vol. VII, North-Holland, Amsterdam, 2000. |
[12] |
R. Eymard, D. Hilhorst and M. Vohralík, A combined finite volume-nonconforming/mixed-hybrid finite element scheme for degenerate parabolic problems, Numer. Math., 105 (2006), 73-131. doi: 10.1007/s00211-006-0036-z
![]() |
[13] |
H. I. Freedman and J. B. Shukla, Models for the effects of toxicant in single-species and predator-prey systems, J. Math. Biol., 30 (1991), 15-30. doi: 10.1007/BF00168004
![]() |
[14] |
T. G. Hallam, C. E. Clark and R. R. Lassider, Effects of toxicants on populations: A qualitative approach I. Equilibrium environment exposured, Ecol. Model, 18 (1983), 291-304. doi: 10.1016/0304-3800(83)90019-4
![]() |
[15] | T. G. Hallam, C. E. Clark and G. S Jordan, Effects of toxicants on populations: A qualitative approach II. First order kinetics, J. Math. Biol., 18 (1983), 25-37. |
[16] |
T. G. Hallam and J. T. De Luna, Effects of toxicants on populations: A qualitative approach III. Environment and food chains pathways, J. Theor. Biol., 109 (1984), 11-29. doi: 10.1016/S0022-5193(84)80090-9
![]() |
[17] | J.-L. Lions, "Quelques Méthodes de Résolution des Problèmes aux Limites Non Linéaires," Dunod, 1969. |
[18] |
C. A. Raposo, M. Sepúlveda, O. Vera, D. Carvalho Pereira and M. Lima Santos, Solution and asymptotic behavior for a nonlocal coupled system of reaction-diffusion, Acta Appl. Math. 102 (2008), 37-56. doi: 10.1007/s10440-008-9207-5
![]() |
[19] |
J. Simon, Compact sets in the space , Ann. Mat. Pura Appl., 146 (1987), 65-96. doi: 10.1007/BF01762360
![]() |
[20] |
J. B. Shukla and B. Dubey, Simultaneous effect of two toxicants on biological species: A mathematical model, J. Biol. Syst., 4 (1996), 109-130. doi: 10.1142/S0218339096000090
![]() |
[21] | R. Temam, "Navier-Stokes Equations, Theory and Numerical Analysis," 3rd revised edition, North-Holland, Amsterdam, reprinted in the AMS Chelsea series, AMS, Providence, 2001. |
[22] | M. Vohralik, "Numerical Methods for Nonlinear Elliptic and Parabolic Equations. Application to Flow Problems in Porous and Fractured Media," Ph.D. dissertation, Université de Paris-Sud Czech Technical University, Prague, 2004. |
[23] |
X. Yang, Z. Jin and Y. Xue, Weak average persistence and extinction of a predator-prey system in a polluted environment with impulsive toxicant input, Chaos Solitons Fractals, 31 (2007), 726-735. doi: 10.1016/j.chaos.2005.10.042
![]() |
[24] | K. Yosida, "Functional Analysis and its Applications," New York, Springer-Verlag, 1971. |
1. | Nathalie Krell, Émeline Schmisser, Nonparametric estimation of jump rates for a specific class of piecewise deterministic Markov processes, 2021, 27, 1350-7265, 10.3150/20-BEJ1312 | |
2. | Alessandro Calvia, Stochastic filtering and optimal control of pure jump Markov processes with noise-free partial observation, 2020, 26, 1292-8119, 25, 10.1051/cocv/2019020 | |
3. | Elena Bandini, Michèle Thieullen, Optimal Control of Infinite-Dimensional Piecewise Deterministic Markov Processes: A BSDE Approach. Application to the Control of an Excitable Cell Membrane, 2021, 84, 0095-4616, 1549, 10.1007/s00245-020-09687-y | |
4. | Benoîte de Saporta, Aymar Thierry d’Argenlieu, Régis Sabbadin, Alice Cleynen, Inés P. Mariño, A Monte-Carlo planning strategy for medical follow-up optimization: Illustration on multiple myeloma data, 2024, 19, 1932-6203, e0315661, 10.1371/journal.pone.0315661 |
| |
| |
| |