324
Views
11
CrossRef citations to date
0
Altmetric
Articles

Detection-Identification of multiple unknown time-dependent point sources in a 2D transport equation: application to accidental pollution

Pages 1423-1447 | Received 13 May 2016, Accepted 20 Nov 2016, Published online: 20 Dec 2016

Abstract

We address the nonlinear inverse source problem of identifying multiple unknown time-dependent point sources occurring in a two-dimensional evolution advection–dispersion–reaction equation. Provided to be available within the monitored domain interfaces for recording the generated state and its flux crossing each suspected zone where a source could occur, we establish a constructive identifiability theorem based on an introduced dispersion-current function that yields uniqueness of the unknown elements defining all occurring sources. Then, the established theorem leads to develop a detection-identification method that goes throughout the monitored domain to detect in each suspected zone whether there exists or not an occurring source. Once a source is detected, the developed method determines lower and upper bounds of the mean value discharged by its unknown time-dependent intensity function. Thereafter, the method localizes the sought position of the detected source as the unique solution of an equation satisfied by the introduced dispersion-current function and identifies its unknown intensity function from solving an associated deconvolution problem. Ultimately, the unknown number of occurring sources is deduced as the sum of all detected-identified active sources. Some numerical experiments on a variant of the surface water BOD pollution model are presented.

1. Introduction

Non-linear inverse source problems are usually known to be somewhat delicate since in general there is no identifiability (uniqueness) of unknown sources under an abstract form using only local boundary or/and internal measurements related to the associated state, see [Citation1] for a counterexample. In the literature, to overcome this difficulty authors generally assume to be available some a priori information on the form of the sought source: for example, time-independent sources are treated by Cannon [Citation2] using spectral theory, then by Engl et al. [Citation3] using the approximated controllability of the heat equation. The results of this last paper are generalized by Yamamoto in [Citation4] to sources of separated time and space variables where the time-dependent part is assumed to be known and null at the initial time, then recently in [Citation5] to sources where the known time-dependent part of the sought source could also depend on the space variables and the involved differential operator is a time fractional parabolic equation. Hettlich and Rundell addressed in [Citation6] the two-dimensional (2D) inverse source problem for the heat equation with as a source the characteristic function associated to a subset of a disk. El Badia and Hamdi treated in [Citation7,Citation8] the case of a time-dependent point source occurring in a one-dimensional (1D) evolution transport equation with constant coefficients. Then, Hamdi and Mahfoudhi extended this study in [Citation9] to the case of equations with spatially varying coefficients and Ben Belgacem et al. [Citation10,Citation11] to the case of a moving time-dependent point source. A detection-identification method for multiple unknown time-dependent point sources in a one-dimensional evolution transport equation was recently developed by Hamdi in [Citation12]. Hamdi and Mahfoudhi treated in [Citation13] the identification of a single time-dependent point source in a two-dimensional evolution advection–dispersion–reaction equation. El Badia et al. studied in [Citation14] the case of multiple time-dependent moving point sources in a two-dimensional evolution advection–diffusion–reaction equation.

The present study is motivated by the results established in the two recent papers.[Citation12,Citation14] Firstly, in [Citation14] authors reported some difficulties encountered by the proposed identification procedure that consists of minimizing the two objective functions least squares and Kohn–Vogelius. In fact, the identifiability result in [Citation14] is based on the unique continuation Theorem that suggests to identify the unknown elements defining whatever finite number of time-dependent point sources using only local boundary measurements taken at any non-empty part of the boundary. In our view, that is an ideal theoretical framework which doesn’t give any visibility on how to proceed in terms of identification and in practice doesn’t take into account the flow properties: for example, the case of an advection dominant flow would imply that using only data reaching back the inflow boundary doesn’t enable to cover the entire activity of downstream occurring sources as well as the case of a flow defined by a velocity field fulfilling the no mass penetrate a solid interface and no-slip conditions along the lateral boundaries which leads to a similar conclusion for using only data reaching these boundaries. Moreover, in terms of identification although in [Citation14] the authors employed measurements taken on the whole boundary of the considered domain, the least squares approach doesn’t enable to identify more than one active source whereas the Kohn–Vogelius approach identifies two sources only under some particular conditions: sources should be well separated and diffusion coefficient should be very small to avoid rapid intermixing. Secondly, the recent results established in [Citation12] prove that in the case of a one-dimensional advection–diffusion–reaction equation, local measurements taken only upstream and/or downstream all occurring sources don’t yield uniqueness of the unknown elements defining more than one time-dependent point source.

The originality of the present study consists in establishing a constructive identifiability theorem that leads to develop a detection-identification method able to identify accurately multiple unknown time-dependent point sources occurring in a two-dimensional evolution advection–dispersion–reaction equation from some local measurements taking into account the flow properties. A motivation for our study is a typical problem associated with environmental monitoring that aims to identify unknown pollution sources occurring in surface water. For example, in a river the involved unknown pollution sources can be identified from measuring the BOD (Biochemical Oxygen Demand) concentration which represents the amount of dissolved oxygen consumed by the micro-organisms during the oxidation process.[Citation15,Citation16] The paper is organized as follows: Section 2 is devoted to stating the problem, assumptions and proving some technical results for later use. In Section 3, based on an introduced dispersion-current function, we establish a constructive identifiability theorem that yields uniqueness of the unknown elements defining the occurring sources. Section 4 is reserved to develop a detection-identification method that goes throughout the monitored domain to detect the presence of all occurring active sources. Once a source is detected, the developed method determines lower and upper bounds of the mean value discharged by its unknown time-dependent intensity function, localizes its sought position and identifies its intensity function. Some numerical experiments on a variant of the surface water BOD pollution model are presented in Section5.

2. Problem statement

Let T>0 be a final monitoring time and Ω be a bounded and connected open subset of IR2 with a sufficiently smooth boundary Ω:=ΓinΓLΓout where Γin denotes the inflow boundary, Γout designates the outflow boundary whereas ΓL regroups the two lateral lower and upper boundaries ΓL=ΓLlΓLu. The evolution of the BOD concentration, denoted here by u, is governed by the following equation, see [Citation15,Citation17,Citation18]:(1) L[u](x,t)=F(x,t)inΩ×(0,T)(1)

where F represents the set of all occurring pollution sources and L is the second-order linear parabolic partial differential operator defined as follows:(2) L[u](x,t):=tu(x,t)-div(D(x)u(x,t))+V(x)·u(x,t)+Ru(x,t)(2)

where D represents the hydrodynamic dispersion tensor, V is the flow velocity field and R is a real number that designates the first-order decay reaction coefficient. The velocity field V is defined by V=(V1,V2) where V1 and V2 are two Lipschitz functions in an open subset O of IR2 containing Ω¯ i.e. Ω¯O that satisfy:(3) div(V)=curl(V)=0inOandV20,V1>0a.e. inΩ(3)

The two first conditions in (Equation3) imply that the flow is incompressible and irrotational. Hydrodynamic dispersion occurs as a consequence of two processes: molecular diffusion resulting from the random molecular motion and mechanical dispersion which is caused by non-uniform velocities. The summation of these two processes defines as follows the hydrodynamic dispersion tensor [Citation19]:(4) D:=DMI+D1D0D0D2(4)

where DM is a positive real number that represents the molecular diffusion coefficient, I is the 2×2 identity matrix and the coefficients Di=0,1,2 are such that, see [Citation17,Citation20]:(5) D0=V1V2(DL-DT)V22,D1=DLV12+DTV22V22andD2=DLV22+DTV12V22(5)

where DTDL are two positive real numbers that represent the mean transverse and longitudinal dispersion coefficients. Moreover, using (Equation4)-(Equation5) the hydrodynamic dispersion tensor D can be rewritten as follows:(6) D=(DM+DT)I+DL-DTV22V·VDM+DTDX·XX22DM+DL(6)

for all XIR2\{(0,0)}. Hence, (Equation6) implies that D is uniformly elliptic and bounded.

For later use and as in view of (Equation6) the matrix D is invertible then, there exists a unique vector field X solution of the linear system DX+V=0 in O. Moreover, we have(7) D-1=1det(D)DM+D2-D0-D0DM+D1X=-1det(D)DMV1+D2V1-D0V2DMV2+D1V2-D0V1(7)

Furthermore, according to (Equation5) we find(8) det(D)=(DM+DL)(DM+DT)D2V1-D0V2=V1DTD1V2-D0V1=V2DT(8)

Thus, using the results (Equation8) to substitute the associated terms in the right-hand side of the second equation in (Equation7) gives X=-1DM+DLV. Since from (Equation3) it holds curl(V)=0 in O then, X is a gradient field derived from a scalar potential ψ that solves(9) Dψ+V=0ψ=-1DM+DLVinO(9)

Hence, the potential ψ that yields ψ(a,b)=0 where (a,b)Ω is defined by(10) ψ(x1,x2)=-1DM+DLax1V1(η,x2)dη+bx2V2(a,ζ)dζ(10)

In addition, using the vector V=(-V2,V1) we consider the following equation:(11) Dψ+V=0ψ=-1DM+DTVinO(11)

The second equation in (Equation11) is obtained using similar techniques as in (Equation7) and (Equation8). Moreover, as from (Equation3) we have div(V)=curl(V)=0 in O then, the scalar potential ψ that solves (Equation11) with ψ(a,b)=0 is defined by(12) ψ(x1,x2)=1DM+DTax1V2(η,x2)dη-bx2V1(a,ζ)dζ(12)

Therefore, from (Equation9)–(Equation11) and according to (Equation3) it follows that the three following functions:(13) z(x,t)=eRtz(x)forz=z0,z=eψandz=ψ(13)

where z0 is a real number, solve the adjoint equation(14) -tz-div(Dz)-V·z+Rz=0inΩ×(0,T)(14)

Lemma 2.1:

Provided (Equation3) holds true, the dispersion-current function Ψ defined by(15) Ψ:ΩIR2x=(x1,x2)Ψ(x)=ψ(x)ψ(x)(15)

is injective.

Let x^=(x^1,x^2) and x~=(x~1,x~2) be two elements of Ω. For the clearness of our proof, we proceed in two steps. Step 1: We prove that we have(16) Ψ(x^)=Ψ(x~)x^1x~1V1(η,x~2)dη=-x^2x~2V2(x^1,ζ)dζx^1x~1V2(η,x~2)dη=x^2x~2V1(x^1,ζ)dζ(16)

To this end, using the definition of the function ψ introduced in (Equation10), we obtain(17) ψ(x^)=ψ(x~)x^2x~2V2(a,ζ)dζ=ax^1V1(η,x^2)dη-ax~1V1(η,x~2)dη(17)

Therefore, using Chasles’s relation by introducing the variable x^1 on the integral over (a,x~1) in (Equation17), we obtain(18) x^2x~2V2(a,ζ)dζ=-x^1x~1V1(η,x~2)dη+ax^1[V1(η,ζ)]ζ=x~2ζ=x^2dη(18)

Moreover, since V1, V2 are two Lipschitz functions in Ω and curl(V)=0 in Ω which implies ζV1(η,ζ)=ηV2(η,ζ) for all (η,ζ)Ω it follows using Fubini’s rule that(19) ax^1[V1(η,ζ)]ζ=x~2ζ=x^2dη=ax^1x~2x^2ζV1(η,ζ)dζdη=x~2x^2ax^1ηV2(η,ζ)dηdζ=-x^2x~2[V2(η,ζ)]η=aη=x^1dζ(19)

Thus, from substituting the last term in (Equation18) by its value obtained in (Equation19) we find the first equation announced in the system (Equation16). Besides, from (Equation12) we get(20) ψ(x^)=ψ(x~)x^2x~2V1(a,ζ)dζ=ax~1V2(η,x~2)dη-ax^1V2(η,x^2)dη(20)

Then, using Chasles’s relation by introducing x^1 on the integral over (a,x~1) in (Equation20) gives(21) x^2x~2V1(a,ζ)dζ=x^1x~1V2(η,x~2)dη+ax^1[V2(η,ζ)]ζ=x^2ζ=x~2dη(21)

In addition, as V1, V2 are two Lipschitz functions in Ω and div(V)=0 in Ω which implies ηV1(η,ζ)=-ζV2(η,ζ) for all (η,ζ)Ω it follows using Fubini’s rule that(22) ax^1[V2(η,ζ)]ζ=x^2ζ=x~2dη=ax^1x^2x~2ζV2(η,ζ)dζdη=-x^2x~2ax^1ηV1(η,ζ)dηdζ=-x^2x~2[V1(η,ζ)]η=aη=x^1dζ(22)

Hence, from substituting the last term in the right hand-side of (Equation21) by its value found in (Equation22), we obtain the second equation in the system (Equation16).

Step 2: Since from (Equation3) we have V1>0 and V20 a.e. in Ω, it follows that the four integrals defining the system (Equation16) are all null either because V2=0 or because the two left-hand side terms have the same sign whereas the two right-hand side terms have two opposite signs. That implies x^1=x~1 and x^2=x~2 which means x^=x~.

Besides, to (Equation1) and (Equation2) one has to append initial and boundary conditions. We could consider without loss of generality that there is no pollution occurring at the initial monitoring time and thus, use a null initial BOD concentration. As far as the boundary conditions are concerned, an homogeneous Dirichlet condition on the inflow boundary Γin seems to be reasonable since the convective transport generally dominates the diffusion process. However, other physical considerations suggest using rather a Neumann homogeneous condition on the two remaining parts ΓL and Γout of the boundary Ω. Therefore, we employ the following initial and boundary conditions:(23) u(·,0)=0inΩu=0onΓin×(0,T)Du·ν=0on(ΓoutΓL)×(0,T)(23)

where ν is the unit outward vector normal to Ω. Notice that due to the linearity of the operator L introduced in (Equation2) and according to the superposition principle, the use of a non-zero initial condition and/or inhomogeneous boundary conditions do not affect the results established in this paper.

As far as F is concerned, we are interested in the case of multiple time-dependent point sources occurring in the main problem (Equation1), (Equation2) and (Equation23) i.e. F is defined as follows:(24) F(x,t)=n=1Nλn(t)δ(x-Sn)inΩ×(0,T)(24)

where NN, δ denotes the Dirac mass, Sn=1,,N are N distinct interior locations in Ω that represent the positions of the occurring sources and λn=1,,NL2(0,T) designate their associated time-dependent source intensity functions that satisfy:(25) 0Tλn(t)eRtdt0andT0(0,T)/λn=0in(T0,T),n=1,,N(25)

Then, using the transposition method introduced by Lions [Citation21] it follows that the problem (Equation1), (Equation2) and (Equation23), (Equation24) admits a unique solution u that belongs to(26) L2(0,T;L2(Ω))C0(0,T;H-1(Ω))(26)

For later use, given NmaxN, let Γi=1,,Nmax be sufficiently smooth interfaces in Ω such that by setting Γ0=Γin, ΓNmax+1=Γout and ωi=1,,Nmax+1 the open subdomains of Ω defined by ωi=Γi-1ΓLiΓi, the three following criteria hold true:(27) (C1)ΓiΓj=,ij{0,,Nmax+1}.(C2)|ΓiΓLl|=|ΓiΓLu|=1,i=1,,Nmax.(C3)Ω¯=i=1Nmax+1ω¯iwhereωNmax+1Ω\{Sn,n=1,,N}.(27) ΓLi is the lateral boundary ΓL situated between Γi-1 and Γi whereas || denotes the cardinality of a set. The criteria (C1,2,3) in (Equation27) are well satisfied if the Nmax interfaces Γi=1,,Nmax are taken distinct and, for example, "parallel” to the inflow and outflow boundaries Γin and Γout such that the involved sources Sn=1,,N occur upstream ΓNmax.

The present study is motivated by the inverse source problem of identifying accidental pollution sources in surface water. That is the case where the unknown pollution sources occur in some given suspected zones within the monitored domain and thus, interfaces for measuring data generated by the occurring unknown sources coming in and out of each suspected pollution zone can be available. For example, in a monitored portion of a river, pollution is usually the consequence of discharging wastewater by one or several sources among industrials, farmers and city sewers occurring in some suspected zones that can be separated by interfaces for measuring the generated data crossing each interzone. In the remainder, we denote ωi=1,,Nmax the suspected pollution zones in the monitored domain Ω and assume the interfaces Γi=ωiωi+1 to be available for measuring the state and its flux crossing each interzone. In addition, we consider that at most only one pollution source could occur in each ωiΩ (we rediscuss this assumption later in the outlook). Therefore, the source positions Sn=1,,N defining F in (Equation24) are assumed to be such that:(28) i=1,,Nmax,there existsatmost̲one elementn{1,,N}/Snωi(28)

Developing an accurate identification method that leads to localize the responsable(s) and to determine the amount of each discharged accidental pollution would be of great use namely ecologically: by treating wastes to preserve the aquatic biodiversity, sanitary: by alerting downstream drinking water stations, legislatively: by holding accountable the responsable(s) and imposing penalties to prevent all not allowed future discharges, etc.

Besides, since in view of (Equation28) the source positions Sn=1,,N are interior locations in ωi=1,,NmaxΩ where ωi=Γi-1ΓLiΓi it follows that the state u solution of the problem (Equation1), (Equation2) and (Equation23), (Equation24) is smooth enough on Ω as well as on the Nmax interfaces Γi=1,,Nmax. That allows to define the following observation operator:(29) M[F]:={Du·νonin;V2uonL;(u,Du·ν)oni=1,,Nmax}(29)

where in=Γin×(0,T), i=Γi×(0,T) for i=1,,Nmax and L=ΓL×(0,T). This is the so-called forward problem.

The inverse problem with which we are concerned here is: assuming (Equation25) and (Equation28) hold true and given the measurements ddin of Du·ν on in, dL of u on L and (di,ddi) of (u,Du·ν) on i=1,,Nmax, determine the unknown elements N, Sn=1,,N and λn=1,,N defining F in (Equation24) that yield(30) M[F]={ddinonin;V2dLonL;(di,ddi)oni=1,,Nmax}(30)

Remark 2.2:

If along the lateral boundary ΓL the flow velocity field V fulfils: V·ν=0 as mass cann’t penetrate a solid interface and V·τ=0 which is the no-slip condition [Citation22], where τ is a unit vector tangent to ΓL then, V2=0 on ΓL and thus, the observation operator M[F] in (Equation29) does not require any measurement related to the state u along ΓL.

Besides, for later use we recall the following technical lemma established in [Citation13]:

Lemma 2.3:

For all xΩ, there exists two boundary controls γinL2(Γin×(0,T)) and γoutL2(Γout×(0,T)) that maintain the solution v of the system:(31) tv-div(Dv)-V·v+Rv=0inΩ×(0,T)v(·,0)=0inΩv=γinonΓin×(0,T)(Dv+vV)·ν=0onΓL×(0,T)(Dv+vV)·ν=γoutonΓout×(0,T)(31) (32) such that:v(x,·)0a.e. in(0,T)(32)

3. Identifiability

In this section, we prove that under some reasonable assumptions the observation operator M[F] introduced in (Equation29) determines in a unique manner the unknown elements N, Sn=1,,N and λn=1,,N defining in (Equation24) the sought source F occurring in the problem (Equation1), (Equation2) and (Equation23). This result is given by the following identifiability theorem:

Theorem 3.1:

Let F(x,t)=n=1Nλn(t)δ(x-Sn) be NN unknown time-dependent point sources occurring in the problem (Equation1), (Equation2) and (Equation23). Given NmaxN, let Γi=1,,Nmax and ωi=1,,Nmax+1 be as in (Equation27). Provided Lemmas 2.1 and 2.3 apply, λn=1,,NL2(0,T) satisfy (Equation25) and Sn=1,,N fulfill (Equation28), the observation operator M[F] in (Equation29) determines uniquely the unknown elements N, Sn=1,,N and λn=1,,N defining F.

For k=1,2, let u(k) be the state solution of the problem (Equation1), (Equation2) and (Equation23) involving the source F(k)(x,t)=n=1N(k)λn(k)(t)δ(x-Sn(k)) where 1N(k)Nmax, λn=1,,N(k)(k) and Sn=1,,N(k)(k) satisfy (Equation25) and (Equation28). We denote M[F(k)] the associated observation operator as in (Equation29). Then, from setting M[F(2)]=M[F(1)], the variable w=u(2)-u(1) satisfies(33) Dw·ν=0onin,V2w=0onL,w=Dw·ν=0oni=1,,Nmax(33)

The aim is to prove that implies F(2)=F(1). Since for k=1,2 the source positions Sn=1,,N(k)(k) satisfy (Equation28), it follows that for all i=1,,Nmax we have in ωi one of the three following cases: Case 1. There is no source occurring in ωi and thus, F(2)=F(1) in ωi×(0,T). Case 2. There is a source λn(2)δ(x-Sn(2)) from F(2) and a source λn(1)δ(x-Sn(1)) from F(1) occurring in ωi. Case 3. There is an only one source λn(k)δ(x-Sn(k)) where k{1,2} occurring in ωi. We start by treating Case 2. which leads to deduce that Case 3. is absurd in view of (Equation25). In Case 2. and according to (Equation33), the variable w solves(34) L[w](x,t)=λn(2)(t)δ(x-Sn(2))-λn(1)(t)δ(x-Sn(1))inωi×(0,T)w(·,0)=0inωiw=0onΓi-1×(0,T)Dw·ν=0on(ΓLiΓi)×(0,T)(34)

Then, multiplying the first equation in (Equation34) by z introduced in (Equation13) that solve the adjoint equation (Equation14) and integrating by parts over ωi using Green’s formula and (Equation33) gives(35) λn(2)(t)z(Sn(2),t)-λn(1)(t)z(Sn(1),t)=ddtw(·,t),z(·,t)+ΓLiw(Dz+zV)·νdx(35)

for all t(0,T). Besides, according to (C3) of (Equation27), w solves in ωNmax+1×(0,T) a system similar to (Equation34) where the first equation becomes homogeneous. That implies w=0 in ωNmax+1×(0,T) which, in view of (Equation25) and by considering the problem solved by w in Ω×(T0,T) gives from applying the unique continuation Theorem in [Citation23] that w(·,T)=0 in Ω. Therefore, by integrating (Equation35) over (0, T) we obtain(36) 0T(λn(2)(t)z(Sn(2),t)-λn(1)(t)z(Sn(1),t))dt=ΓLi×(0,T)w(Dz+zV)·νdxdt(36)

Hence, employing in (Equation36) the functions z=eRtz for z=z0, z=eψ and z=ψ leads to(37) λ¯n(2)z(Sn(2))-λ¯n(1)z(Sn(1))=ΓLi×(0,T)eRtw(Dz+zV)·νdxdt(37)

where λ¯n(k)=0Tλn(k)(t)eRtdt. Furthermore, the second equation in (Equation33) implies that either V=0 or w=0 on ΓL×(0,T). Consequently, from (Equation9)–(Equation11) the term w(Dz+zV)·ν=0 on ΓLi×(0,T). Thus, it follows from (Equation37) that Sn(2) and Sn(1) are subject to:(38) ψ(Sn(2))=ψ(Sn(1))ψ(Sn(2))=ψ(Sn(1))Ψ(Sn(2))=Ψ(Sn(1))Sn(2)=Sn(1)(38)

The function Ψ and the implication in (Equation38) are given by Lemma 2.1. We set Sn(2)=Sn(1)=Sn in (Equation34). Let γin and γout be such that the solution v of (Equation31) yields (Equation32) for x=Sn i.e. v(Sn,·)0 a.e. in (0, T). Then, given θ(0,T) the variable vθ(·,t)=v(·,θ-t) solves(39) -tvθ-div(Dvθ)-V·vθ+Rvθ=0inΩ×(0,θ)vθ(·,θ)=0inΩvθ(·,t)=γin(·,θ-t)onΓin×(0,θ)(Dvθ+vθV)·ν=0onΓL×(0,θ)(Dvθ(·,t)+vθ(·,t)V)·ν=γout(·,θ-t)onΓout×(0,θ)(39)

Furthermore, from multiplying the first equation in (Equation34) by vθ and integrating by parts over ωi×(0,θ) using Green’s formula and (Equation33), we find(40) 0θ(λn(2)(t)-λn(1)(t))v(Sn,θ-t)dt=0,for allθ(0,T)(40)

According to Titchmarsh’s Theorem on convolution of L1 functions [Citation24] and since it holds v(Sn,·)0 a.e. in (0, T), the equation (Equation40) implies that λn(2)=λn(1) a.e. in (0, T). Hence, in Case 2. we get also F(2)=F(1) in ωi×(0,T). Besides, in Case 3. w solves a system similar to (Equation34) where in the right-hand side of the first equation occurs only the source λn(k)δ(x-Sn(k)) where k{1,2}. Then, using the same techniques employed in Case 2., we obtain from (Equation37) for z=z0 that λ¯n(k)=0 which is absurd since λn=1,,N(k)(k) satisfy (Equation25). Therefore, it follows that F(2)=F(1) in ωi×(0,T) for all i=1,,Nmax.

4. Detection-identification method

Given NmaxN, let Γi=1,,Nmax be the interfaces introduced in (Equation27) and ωi=1,,NmaxΩ be the source suspected zones defined by ωi=Γi-1ΓLiΓi where Γ0=Γin. Then, assuming the unknown elements λn=1,,N and Sn=1,,N defining in (Equation24) F involved in the problem (Equation1)-(Equation2) and (Equation23) to be such that (Equation25) and (Equation28) hold true, we focus in this section on developing from the observation operator M[F] introduced in (Equation29) a detection-identification method that goes throughout the monitored domain Ω to detect for all i=1,,Nmax whether there exists or not an active source occurring in the suspected zone ωiΩ. Once a source is detected, the developed method determines lower and upper bounds of the mean value discharged by its unknown time-dependent intensity function λn. Thereafter, it localizes the position Sn of the detected source as the unique solution of an equation satisfied by the dispersion-current function introduced in (Equation15) and identifies the historic of λn. Ultimately, the unknown number N of occurring sources is deduced as the sum of all detected-identified active sources. We proceed as follows:

4.1. Detection of active sources

For i=1,,Nmax and in view of (Equation28), multiplying the main equation (Equation1)–(Equation2) in ωi×(0,T) by the adjoint function z=eRtz0 for z0=1/T that solves the adjoint equation introduced in (Equation14), and integrating by parts using Green’s formula gives(41) SDi:=1TP1i=1Tλ¯nifSnωi0otherwise(41)

where λ¯n=0Tλn(t)eRtdt and the coefficient P1i is defined by(42) P1i=eRTωiu(x,T)dx+(Γi-1Γi)×(0,T)eRt(uV-Du)·νdxdt+ΓLi×(0,T)eRtuV·νdxdt(42)

The notation SDi introduced in (Equation41) is what we refer to in the remainder as the Source Detection index. Notice that for i=1,,Nmax, the coefficient P1i obtained in (Equation42) can be completely determined from the measurements M[F]. Then, according to (Equation41) and since λn satisfies (Equation25) which implies that λ¯n0 it follows that whether SDi=P1i/T is null or not corresponds to whether there is no or there is an active source occurring in ωiΩ. Moreover, if the source intensity function λn keeps a constant sign, for example, λn0 a.e. in (0, T) and the reaction coefficient R0 then, we have λn(t)λn(t)eRtλn(t)eRT in (0, T). Therefore, from (Equation41) we deduce the following lower and upper bounds of the mean value discharged by λn:(43) e-RTSDie-RT0SDi1T0Tλn(t)dtSDi(43)

In practice, usually the reaction coefficient R is small enough and thus, (Equation43) provides an approximation of the mean value discharged by the involved unknown time-dependent source intensity function λn without having to identify its historic. The active source detection procedure yields a useful tool that in practice leads to decide whether to go further with the identification of a detected source or to consider that it is not a significant source since the mean value discharged by its unknown intensity function remains below a certain tolerance/criterion. Besides, the knowledge of the time active limit T0 in (Equation25), using for example the results established in [Citation25], improves the lower bound in (Equation43).

4.2. Identification of the detected source occurring in ωi

Once an active time-dependent point source is detected to be occurring in ωiΩ i.e. in (Equation42) the coefficient P1i0, we apply the following identification procedure to determine the two unknown elements Sn and λn defining the involved source: From multiplying Equations (Equation1)–(Equation2) in ωi×(0,T) by the adjoint functions z=eRtz introduced in (Equation13) that solve the adjoint Equation (Equation14) and integrating by parts using Green’s formula, we obtain(44) λ¯nz(Sn)=eRTωiu(x,T)zdx+ωi×(0,T)eRt(u[zV+Dz]-zDu)·νdxdt(44)

Since on ΓLi×(0,T) we have Du·ν=0 then, from substituting in (Equation44) z by the adjoint functions z=z00, z=eψ and z=ψ introduced in (Equation13) it follows that the unknown elements Sn and λn defining the detected source occurring in ωi are subject to:(45) λ¯n=P1i,λ¯neψ(Sn)=Peψiandλ¯nψ(Sn)=Pψi(45)

where P1i is as defined in (Equation42) and the two coefficients Peψi, Pψi are such that(46) Peψi=eRTωieψu(x,T)dx-(Γi-1Γi)×(0,T)eψ+RtDu·νdxdtPψi=eRTωiψu(x,T)dx+(Γi-1Γi)×(0,T)eRt(u[ψV-V]-ψDu)·νdxdt+ΓLi×(0,T)eRtu(ψV-V)·νdxdt(46)

Hence, in view of (Equation45) and (Equation46) it follows that the sought source position Sn solves(47) ψ(Sn)=lnPeψiP1iandψ(Sn)=PψiP1iΨ(Sn)=ln(PeψiP1i)PψiP1i(47)

where Ψ is the injective dispersion-current function introduced in (Equation15). Then, to determine the two unknown elements Sn and λn defining the time-dependent point source detected to be occurring in ωiΩ, we proceed in the two following steps:

4.2.1. Step 1: Localization of the sought source position Sn

Since the coefficients P1i,Peψi,Pψi defined in (Equation42)–(Equation46) and involved in (Equation47) can be completely determined from the measurements M[F] then, due to the injectivity of the dispersion-current function Ψ introduced in (Equation15) one could perform Newton’s iterations to determine the unique solution Sn of the last equation in (Equation47) as the zero of the associated vector function. To this end, we compute the determinant det(JΨ) of the 2×2 Jacobian matrix JΨ associated to the function Ψ. Then, from (Equation15) and using the expressions of ψ and ψ given in (Equation9) and (Equation11) we obtain in Ω(48) det(JΨ)=x1ψx2ψ-x2ψx1ψ=V22det(D)>0(48)

whereas established in (Equation8) the determinant of the uniformly elliptic matrix D introduced in (Equation4)-(Equation6) is det(D)=(DM+DL)(DM+DT)>0 and V22=V12+V22>0 from (Equation3). Thus, the 2×2 Jacobian matrix JΨ is invertible in Ω and we have(49) JΨ-1=-det(D)V22V1DM+DT-V2DM+DLV2DM+DTV1DM+DL(49)

That leads to perform the following Newton’s iterations to determine the unique solution of the last equation in (Equation47): given an initial guess x0 in ωi, we iterate(50) xk+1=xk-JΨ-1(xk)ψ(xk)-ln(PeψiP1i)ψ(xk)-PψiP1i,k0(50)

Remark 4.1:

Since in Ω the state u is subject to only knowledge of its value and the value of its flux on the interfaces Γi=1,,Nmax it follows that the developed active source detection procedure as well as the localization of the unknown detected source are not yet available due to the not given term u(·,T) involved in the coefficients P1i, Peψi and Pψi.

To determine the final state u(·,T) in Ω of the main problem (Equation1), (Equation2) and (Equation23), (Equation24), we employ the change of variables u(x,t)=e12ψ(x)u(x,t) in Ω×(0,T). Then, since λn=1,,N satisfy (Equation25) it follows that the variable u solves for all T(T0,T) the system:(51) tu-div(Du)+ρu=0inΩ×(T,T)u(·,T)L2(Ω)inΩu=0onΓin×(T,T)(Du+12uV)·ν=0on(ΓLΓout)×(T,T)(51)

where ρ=R+14VD-1V=R+V22/(4(DM+DL)). Besides, in view of the condition (C3) in (Equation27), the variable u solves in ωNmax+1×(0,T) a system similar to (Equation51) where the initial condition becomes u(·,0)=0 in ωNmax+1 and the first boundary condition is replaced by u=e12ψu on ΓNmax×(0,T), whereas the last boundary condition is same and holds on (ΓLiΓout)×(0,T). Therefore, given measurements of the state u on ΓNmax×(0,T) we employ a numerical method to compute u in ωNmax+1×(0,T). Then, we consider the data assimilation problem [Citation26] of determining the final state u(·,T) in Ω of the problem (Equation1)-(Equation2) and (Equation23)-(Equation24) from the computed data u in ωNmax+1×(T,T). To this end, we establish the following result:

Proposition 4.2:

Let u be the solution of the problem (Equation1), (Equation2) and (Equation23), (Equation24), u=e12ψu in Ω×(0,T) and ωNmax+1 be the open subdomain of Ω introduced in (Equation27). Provided the assertion (Equation25) holds true, the data u in ωNmax+1×(T,T) where T(T0,T) determines in a unique manner the final state u(·,T) in Ω and we have(52) u(·,T)L2(Ω)CuL2(ωNmax+1×(T,T))(52)

where C is a Carleman constant [Citation27] that does not depend on u(·,T).

See the Appendix 1.

To determine the final state u(·,T) in Ω using the data u in ωNmax+1×(T,T), we denote (ek)k0 the normalized eigenfunctions of the following problem:(53) -div(De)+ρe=μeinΩe=0onΓin(De+12eV)·ν=0onΓLΓout(53)

that form a complete orthonormal family of L2(Ω), see [Citation13]. Then, the solution of the problem (Equation51) is expressed as follows:(54) u(x,t)=k0u(·,T),ekL2(Ω)eμk(T-t)ek(x)inΩ×(T,T)(54)

where μk is the eigenvalue associated to ek and ,L2(Ω) denotes the inner product in L2(Ω). Therefore, from trancating the series in (Equation54) using a sufficiently large number K of initial terms, we determine the coefficients yk:=u(·,T),ekL2(Ω) for k=0,,K from solving the following least squares problem:(55) miny0,,yK12k=0Kykeμk(T-t)ek-uL2(ωNmax+1×(T,T))2(55)

That leads using the change of variables u(x,T)=e-12ψ(x)u(x,T) in Ω and in view of (Equation54) to deduce the following approximation of the final state u(·,T) in Ω associated to the main problem (Equation1), (Equation2) and (Equation23), (Equation24):(56) u(x,T)e-12ψ(x)k=0Kykek(x),xΩ(56)

4.2.2. Step 2: Identification of the unknown time-dependent intensity λn

Assuming now the position Sn of the detected active source occurring in ωiΩ to be already localized, we focus here on identifying the historic of its unknown time-dependent intensity function λn. To this end, let γin and γout be two boundary controls that lead the solution v of the problem (Equation31) to satisfy the assertion (Equation32) for x=Sn. Then, from multiplying the main Equations (Equation1)–(Equation2) in ωi×(0,θ) by the solution vθ of the system (Equation39) and integrating by parts using Green’s formula, we obtain(57) 0θλn(t)v(Sn,θ-t)dt=Qθi,θ(0,T)(57)

where the coefficient Qθi is defined as follows:(58) Qθi=(Γi-1Γi)×(0,θ)(u(x,t)[Dv(x,θ-t)+v(x,θ-t)V]-v(x,θ-t)Du(x,t))·νdxdt(58)

Therefore, the identification of the unknown time-dependent intensity function λn can be transformed into solving the deconvolution problem in (Equation57)–(Equation58). To this end, given a desired number of time steps M, we employ the regularly distributed discrete times tm=mΔt for m=0,,M where Δt=T/M. Then, using the trapezoidal rule we get(59) 0tm+1λn(t)v(Sn,tm+1-t)dt=k=0mtktk+1λn(t)v(Sn,tm+1-t)dtΔt2k=0m(λn(tk)v(Sn,tm+1-k)+λn(tk+1)v(Sn,tm-k))=Δtk=1mλn(tk)v(Sn,tm+1-k)(59)

where according to (Equation31), we used v(Sn,0)=0 and λn(0)=0. Thus, using the notation λnmλn(tm) and in view of (Equation59) we obtain a discrete version of the deconvolution problem defined from (Equation57)–(Equation58) that leads to the following recursive formula:(60) λnm=1v(Sn,t1)Qtm+1iΔt-k=1m-1λnkv(Sn,tm+1-k),for allm1(60)

For the clearness of our presentation, we summarize the different steps of the developed detection-identification method in the following algorithm:

5. Numerical experiments

We carry out numerical experiments in the case of a rectangular domain Ω defined byΩ:={x=(x1,x2)such that0<x1<Land0<x2<}

with the associated boundaries:(61) Γin:={x=(x1,x2)such thatx1=0and0x2}Γout:={x=(x1,x2)such thatx1=Land0x2}ΓL:={x=(x1,x2)such thatx2=0and0x1L}{x=(x1,x2)such thatx2=and0x1L}(61)

where L=1000m and =100m. Then, we consider a flow defined by a unidirectional mean velocity vector V=(V1,0) where V1>0. Therefore, from setting a=b=0 in (Equation10)–(Equation12) it follows that the dispersion-current function Ψ introduced in (Equation15) is defined for all x=(x1,x2)Ω as follows:(62) Ψ(x)=-V1DM+DLx1-V1DM+DTx2(62)

Furthermore, in view of (Equation4)-(Equation5) the hydrodynamic dispersion tensor D is reduced to the 2×2 diagonal matrix defined by D=diag(DL+DM,DT+DM). To generate the boundary and internal measurements defining in (Equation29) the observation operator M[F], we solve the problem (Equation1)-(Equation2) and (Equation23), where F is as introduced in (Equation24)-(Equation25) for N=3 time-dependent point sources located at Sn=1,2,3 and loading the intensity functions defined in (0,T0) by(63) λ1(t)=χ(T05,T0)(t),λ2(t)=4T02t(T0-t)andλ3(t)=k=13cke-vk(t-τk)2(63)

whereas λn=1,2,3(t)=0 for all tT0. In (Equation63), χ is the characteristic function and the coefficients are such that: c1=1.2, c2=0.4, c3=0.6, v1=10-6, v2=5×10-5, v3=10-6 and τ1=4500s, τ2=6500s, τ3=9000s. Moreover, the mean values of the source intensity functions introduced in (Equation63) are given as follows:(64) 1T0Tλ1(t)dt=4T05T,1T0Tλ2(t)dt=2T03Tand1T0Tλ3(t)dt=π2Tk=13ckvk(erf(vk(T0-τk))-erf(-vkτk))(64)

where erf(θ)=(2/π)0θe-s2ds is the Gauss error function. Besides, we employ the following approximation of the Dirac mass δ(x-Sn) where Sn=(Snx1,Snx2):(65) δ(x-Sn)1πε2e-(x1-Snx1)2ε2-(x2-Snx2)2ε2(65)

with ε=10-5 in (Equation65). As far as the interfaces Γi=1,,Nmax introduced in (Equation27) are concerned, we set Nmax=4 and consider(66) Γi:={x=(x1,x2)such thatx1=i×200and0x2},fori=1,,Nmax(66)

That implies ωi=((i-1)×200,i×200)×(0,) for i=1,,Nmax+1. Notice that provided the involved source positions Sn=1,2,3 are occurring upstream Γ4 and do not lie on Γi=1,,4 then, the three criteria (C1,2,3) introduced in (Equation27) are well satisfied.

To identify the unknown elements Sn=1,2,3 and λn=1,2,3 defining in (Equation24) F that generated the observations M[F], we apply the detection-identification method developed in the previous section by following the steps described in Algorithm. Therefore, we start in view of the criterion (C3) by solving numerically the forward problem satisfied by the state u in ωNmax+1×(0,T). Then, using the computed data u in ωNmax+1×(0,T), we solve the least squares problem introduced in (Equation55) to determine the final state u(·,T) in Ω. To this end, we compute the normalized eigenfunctions (enm)n,m0 of the problem (Equation53) for a mean velocity vector V=(V1,0) and thus, a reduced dispersion tensor D=diag(DL+DM,DT+DM) with ρ=R+V12/(4(DL+DM)). That gives(67) enm(x1,x2)=cnmsin(αnx1)cos(mπx2)for alln,m0(67)

where for all n0, the coefficient αn is determined from solving the equation(68) tan(αnL)=-2(DL+DM)V1LαnL,in((2n+1)π/2,(2n+3)π/2)(68)

and the coefficient cnm is defined by(69) cnm=2Lβnifm=02Lβnifm0whereβn=1-sin(2αnL)2αnL(69) μnm=ρ+(DL+DM)αn2+(DT+DM)(mπ/)2 is the eigenvalue associated to enm.

To carry out numerical experiments and as far as the discretization of Ω×(0,T) is concerned, we employ uniform mesh sizes: Δx1=L/N1, Δx2=/N2 with a time-step Δt=1/Nt and use a 5-points finite differences Crank–Nicholson scheme for N1=10, N2=10 and Nt=240. Then, we set T=14400s (4 hr ), T0=10800s (3 hr ) and use mean longitudinal and transverse dispersion coefficients DL=29m2s-1, DT=1m2s-1 with a molecular diffusion DM=10-5m2s-1, V1=0.01ms-1 and R=10-5s-1. To generate the measurements defining in (Equation29) the observation operator M[F], we solve the problem (Equation1)-(Equation2) and (Equation23) using F as defined in (Equation24) for N=3 time-dependent point sources located at S1=(100,50), S2=(300,70), S3=(700,30) and loading the intensity functions introduced in (Equation63). To identify the final state u(·,T) in Ω, we use the records of the state u on ΓNmax×(0,T) and solve by employing a 5-points finite differences Crank–Nicholson scheme the forward problem satisfied by u in ωNmax+1×(0,T). That leads to generate the data u in ωNmax+1×(T,T). Then, using the Conjugate Gradient Method, we solve the minimization problem introduced in (Equation55) and thus, we define the identified final state as given in (Equation56). The obtained numerical results are presented by drawing in a same graphic both identified and simulated final states. We indicate for each experiment the L2 relative error computed from Error=usimulated(·,T)-uidentified(·,T)L2(Ω)/usimulated(·,T)L2(Ω). We carried out numerical experiments for T=T0 and different values of DL and V1.

Figure 1. (a) V1=0.01, DL=29: Error 0.17 (b) V1=0.001, DL=29: Error 0.08.

Figure 1. (a) V1=0.01, DL=29: Error 0.17 (b) V1=0.001, DL=29: Error 0.08.

Figure 2. (a) V1=0.01, DL=60: Error 0.10 (b) V1=0.01, DL=20: Error 0.22.

Figure 2. (a) V1=0.01, DL=60: Error 0.10 (b) V1=0.01, DL=20: Error 0.22.

The numerical results presented in Figures and show that the procedure described in (Equation51)–(Equation56) identifies accurately the final state u(·,T) in Ω associated to the main problem (Equation1)–(Equation2) and (Equation23)–(Equation24). The analysis of those results indicates that the L2 relative error on the identified final state depends on the dimensionless Peclet number Pe=V1L/(DL+DM). In fact, the reconstruction of u(·,T) in Ω is more accurate for smaller Peclet number.

Then, we apply the developed detection-identification method to identify the unknown elements Sn=1,,N and λn=1,,N defining in (Equation24) the sought source F occurring in Ω=(0,L)×(0,) that generated the measurements M[F] given in (Equation29). To this end, we employ the developed active source detection procedure to go throughout Ω by detecting for all i=1,,Nmax whether there is or not an active source occurring in the subdomain ωi=((i-1)×200,i×200)×(0,). Therefore, we compute for i=1,,Nmax the Source Detection index SDi from (Equation41)-(Equation42) and determine as in (Equation43) the estimated lower and upper bounds of the mean value loaded by the unknwon intensity function λn occurring in ωi. To carry out numerical experiments, we use DL=60m2s-1, V1=0.01ms-1 and generate the simulated measurements M[F] defined in (Equation29) using N=3 point sources loading the time-dependent intensity functions λn=1,,3 introduced in (Equation63) for different source locations Sn=1,2,3. As far as the obtained numerical results are concerned, we present in a same table for each simulated measurements the computed Source Detection index SDi in ωi for i=1,,Nmax=4, the interval [SDie-RT;SDi] obtained in (Equation43) and the corresponding mean value of the intensity function λn, used for simulation, computed from (Equation64) for T0=(3/4)T. We give in the caption of each table the three source positions Sn=1,2,3 used to generate the simulations.

Table 1. Source Detection: simulation with S1=(300,50);S2=(500,70);S3=(700,30).

Table 2. Source Detection: simulation with S1=(100,50);S2=(500,70);S3=(700,30).

Table 3. Source Detection: simulation with S1=(100,50);S2=(300,70);S3=(700,30).

Table 4. Source Detection: simulation with S1=(100,50);S2=(300,70);S3=(500,30).

The analysis of the numerical experiments presented in Tables shows that the developed active source detection procedure determines accurately throughout the monitored domain Ω whether there is or not an active source occurring in each ωi=1,,NmaxΩ. Furthermore, in view of (Equation43) this procedure provides estimated lower and upper bounds SDie-RT and SDi of the mean value loaded by the unknwon time-dependent intensity function λn of the deceted source occurring in ωi. In Tables , the two estimated frame bounds appear to be accurate and yield an approximation of the mean value loaded by the unknown time-dependent intensity function λn without having to identify its historic.

Besides, once an active time-dependent point source is detected in ωiΩ where i{1,,Nmax}, we employ the developed identification procedure to localize its position Sn from (Equation47)-(Equation50) and to identify its time-dependent intensity function λn using the recursive formula (Equation60). To carry out numerical experiments, we use for example the simulations M[F] generated for the experiments in Table 4. Since the results of the detection procedure presented in Table 4 indicate that the involved sources are occurring in ωi for i=1,,3 then, we apply the identification procedure to determine for each i=1,,3 the two unknown elements λn and Sn defining the detected source occurring in ωi. To this end, we compute for i=1,,3 the coefficients P1i, Peψi, Pψi from (Equation42) and (Equation46). Then, according to (Equation47) and in view of (Equation62), the two unknown coordinates Snx1 and Snx2 defining the position Sn of the detected source occurring in ωi are localized as follows:(70) Snx1=-DL+DMV1lnPeψiP1iSnx2=-DT+DMV1PψiP1i(70)

Furthermore, using the localized source position Sn we identify its time-dependent intensity function λn from (Equation60). To this end, we need to solve the boundary control problem introduced in (Equation31)–(Equation32) for x=Sn. By considering the following adjoint problem:(71) -tξ-div(Dξ)+V·ξ+Rξ=δ(x-Sn)inΩ×(0,T)ξ(·,T)=0inΩξ=0onΓin×(0,T)Dξ·ν=0on(ΓLΓout)×(0,T)(71)

and using similar techniques as in [Citation13] it follows that setting γin=-Dξ·ν on Γin×(0,T) and γout=ξ on Γout×(0,T) leads the solution of the problem (Equation31) to yield the assertion (Equation32) for x=Sn. We present the obtained identification results for the three detected sources in Table 4 while introducing a Gaussian noise on the used simulations. We give for each introduced intensity noise the three identified source positions and present on a same graphic for each involved source, the intensity function used for simulation given in (Equation63) and the identified intensity computed from (Equation60). We indicate also for each drawing the L2-relative error computed from λsimulation-λidentifiedL2(0,T)/λsimulationL2(0,T).

Figure 3. (a) Gaussian noise 5%: Error 0.253 (b) Gaussian noise 10%: Error 0.301.

Figure 3. (a) Gaussian noise 5%: Error 0.253 (b) Gaussian noise 10%: Error 0.301.

Figure 4. (a) Gaussian noise 5%: Error 0.141 (b) Gaussian noise 10%: Error 0.221.

Figure 4. (a) Gaussian noise 5%: Error 0.141 (b) Gaussian noise 10%: Error 0.221.

Figure 5. (a) Gaussian noise 5%: Error 0.191 (b) Gaussian noise 10%: Error 0.321.

Figure 5. (a) Gaussian noise 5%: Error 0.191 (b) Gaussian noise 10%: Error 0.321.

The identified source positions are as follows: For a Gaussian noise 5%, we obtained S1=(97.31,48.3), S2=(303.4,65.7) and S3=(498.2,27.4). For a Gaussian noise 10%, we found S1=(112.61,39.8), S2=(309.2,78.4) and S3=(506.2,37.5).

The numerical results obtained for the localization of the sought source positions as well as for the identification of their unknown time-dependent intensity functions presented in Figures show that the developed identification procedure determines accurately the unknown elements defining the occurring sources and is relatively stable with respect to the introduction of a Gaussian noise on the used measurements.

6. Conclusion and outlook

We addressed the non-linear inverse source problem of identifying multiple unknown time-dependent point sources occurring in a two-dimensional evolution advection–dispersion–reaction equation. We developed a source detection procedure that goes throughout the monitored domain to detect the presence of all occurring active sources. Once a source is detected, this procedure determines lower and upper bounds of the mean value discharged by its unknown time-dependent intensity function without having to identify the historic of this latest. In practice, this procedure leads to decide whether to go further with the identification of a detected source or to consider that it is not a significant source since the mean value discharged by its unknown intensity function remains below a certain tolerance/criterion. Then, we developed an identification method that localizes the sought position of a detected source as the unique solution of an equation satisfied by the introduced dispersion-current function and identifies its intensity function from solving an associated deconvolution problem. Some numerical experiments on a variant of the surface water BOD pollution model were carried out. The obtained numerical results show that the developed detection–identification method is accurate and relatively stable with respect to the introduction of a Gaussian noise on the used measurements.

6.1. Outlook

We carried out some numerical experiments where the assumption (Equation28) is not respected. That is we considered a domain Ω containing an only one suspected zone ω1 i.e. Nmax=1 where multiple point sources could occur. Then, we applied the detection–identification method using measurements generated firstly by one, secondly by two and thirdly by three point sources occurring in ω1 by discharging three different positive constants. The obtained numerical results are as follows: in the case of one point source: the localized position Sn and λ¯n are accurate. In the case of two point sources: the localized position Sn is between the two involved source positions. In the case of three point sources: the localized position Sn is in the triangle of summits the three involved source positions. Moreover, in these two last cases, the localized position Sn seems to be nearer to the source position of biggest discharge constant. However, in all cases the determined λ¯n corresponds to the sum of all discharged amounts.

Although it is expected, as in (Equation41)–(Equation42), that the computed λ¯n corresponds to the sum of all amounts discharged by the different involved point sources, it is not clear whether the localized source position is always in the polygon of summits the occurring source positions. An outlook of the present study could be to investigate this claim which would give another application of the developed detection–identification method: to be applied on a monitored domain Ω containing suspected zones ωi where multiple time-dependent point sources of known positions could occur. Then, the method determines the total amount discharged by all sources occurring in ωi. The localized position would be interpreted as follows: (1) if it coincides with one among the already known positions then, it is the unique source responsible of the detected pollution; (2) If it is in a polygon of summits some of the known source positions then, there were more than one active source among the already known positions (the number of active sources is it the number of summits?); (3) Otherwise, the responsable of the detected pollution is rather a new unknown source(s).

Notes

No potential conflict of interest was reported by the authors.

References

  • Hamdi A. Inverse source problem in a 2D linear evolution transport equation. Inverse Prob Sci Eng. 2012;20:401–421.
  • Cannon JR. Determination of an unknown heat source from overspecified boundary data. SIAM J Numer Anal. 1968;5:275–286.
  • Engl HW, Scherzer O, Yamamoto M. Uniqueness of forcing terms in linear partial differential equations with overspecified boundary data. Inverse Prob. 1994;10:1253–1276.
  • Yamamoto M. Conditional stability in determination of densities of heat sources in a bounded domain. Int Ser Numer Math. 1994;18:359–370. Birkhuser, Verlag Basel.
  • Sakamoto K, Yamamoto M. Inverse source problem with a final overdetermination for a fractional diffusion equation. Math Control Related Fields. 2012;1:509–518.
  • Hettlich F, Rundell W. Identification of a discontinuous source in the heat equation. Inverse Prob. 2001;17:1465–82.
  • El Badia A, Ha-Duong T, Hamdi A. Identification of a point source in a linear advection-dispersion-reaction equation: application to a pollution source problem. Inverse Prob. 2005;21:1121–39.
  • Hamdi A. The recovery of a time-dependent point source in a linear transport equation: application to surface water pollution. Inverse Prob. 2009;25:75006–75023.
  • Hamdi A, Mahfoudhi I. Inverse source problem in a one-dimensional evolution linear transport equation with spatially varying coefficients: application to surface water pollution. Inverse Prob Sci Eng. 2013;21:1007–1031.
  • Andrle M, Ben Belgacem F, El Badia A. Identification of moving pointwise sources in an advection–dispersion–reaction equation. Inverse Prob. 2011;27:025007.
  • Ben Belgacem F. Uniqueness for an ill-posed reaction-dispersion model. Application to organic pollution in stream-waters (One-dimensional model). Inverse Prob Imaging. 2012;6–2:163–181.
  • Hamdi A. Detection and identification of multiple unknown time-dependent point sources occurring in 1D evolution transport equations. Inverse Prob Sci Eng. 2016;1–23. doi:10.1080/17415977.2016.1172224.
  • Hamdi A, Mahfoudhi I. Inverse source problem based on two dimensionless dispersion-current functions in 2D evolution transport equations. J Inverse Ill Prob. 2016;24:663–685. doi:10.1515/JIIP-2014-0051.
  • Andrle M, El Badia A. Identification of multiple moving pollution sources in surface waters or atmospheric media with boundary observations. Inverse Prob. 2012;28:075009.
  • Cox BA. A review of currently available in-stream water-quality models and their applicability for simulating dissolved oxygen in lowland rivers. Sci Total Environ. 2003;314–316:335–77.
  • Brown LC, Branwell TO. The enhanced stream water quality models QUAL2E and QUAL2E-UNCAS: Documentation and user manual. EPA: 600/3-87/007. 1987.
  • Myung L, Won S. 2D Finite element pollutant transport model for accidental mass release in rivers. KSCE J Civil Eng. 2010;14:77–86.
  • Akira Okubo. Diffusion and ecological problems: mathematical models. New York (NY): Springer-Verlag; 1980.
  • Bear J, Bachmat Y. Introduction to modeling of transport phenomena in porous media. Dordrecht: Kluwer Academic Publishers; 1991.
  • Oelkers EH. Physical and chemical properties of rocks and fluids for chemical mass transport calculations. In: Lichtner PC, Steefel CC, Oelkers EH, Editors. Reactive transport in porous media. Washington (DC): The Mineralogical Society of America. 1996. p. 131–191.
  • Lions JL. Pointwise control for distributed systems in control and estimation in distributed parameters systems. In: Banks HT, Editor. Frontiers in Applied Mathematics. Philadelphia: Edited by Banks H. T. SIAM; 1992. p. 1–39.
  • Lauga E, Brenner PM, Stone AH. Microfluidics: the no-slip boundary condition, Chapter 15 in Handbook of Experimental Fluid Dynamics. New-York (NY): Springer; 2005.
  • Lin FH. A uniqueness theorem for parabolic equations. Commun Pure Appl Math. 1990;43:125–136. MR 90j:35106.
  • Titchmarsh EC. Introducton to the theory of Fourier Integrals. London: Oxford University Press; 1939.
  • Hamdi A, Mahfoudhi I, Rejaiba A. Identification of the time active limit with lower and upper bounds of the total amount loaded by unknown sources in 2D transport equations. J Eng Math. 2015;97:101–17.
  • Garca G, Osses A, Puel J-P. A null controllability data assimilation methodology applied to a large scale ocean circulation model. M2AN Math Model Numer Anal. 2011;45:361–386.
  • Fernandez-Cara E, Guerrero S. Global Carleman inequalities for parabolic systems and applications to controllability. SIAM J Control Optim. 2006;45:1395–1446.
  • Fursikov A, Imanuvilov OY. Controllability of evolution equations. Lecture notes: Research Institute of Mathematics, Seoul National University, Korea; 1996.
  • Rasmussen JM. Boundary control of linear evolution PDEs-continuous and discrete [Ph.D. thesis]. Technical University of Denmark, 2004.

Appendix 1

Here, we establish the proof of Proposition 4.2:

In view of [Citation28], we consider the null controllability problem of determining a distributed control γL2(ωNmax+1×(T,T)) that to a given initial state φ0L2(Ω) drives the solution φ of the following system:(A1) tφ-div(Dφ)+ρφ=γ(x,t)χωNmax+1(x)inΩ×(T,T)φ(·,T)=φ0inΩφ=0onΓin×(T,T)(Dφ+12φV)·ν=0on(ΓLΓout)×(T,T)(A1) (A2) to the final state:φ(·,T)=0inΩ(A2)

where χωNmax+1 is the characteristic function of ωNmax+1Ω. Multiplying the first equation in (EquationA1) by ξ the solution of the adjoint problem that to a given ξ0L2(Ω) associates(A3) -tξ-div(Dξ)+ρξ=0inΩ×(T,T)ξ(·,T)=ξ0inΩξ=0onΓin×(T,T)(Dξ+12ξV)·ν=0on(ΓLΓout)×(T,T)(A3)

and integrating by parts over Ω×(T,T) using Green’s formula gives(A4) ωNmax+1×(T,T)ξγdxdt=Ω[φ(x,t)ξ(x,t)]TTdx,ξ0L2(Ω)(A4)

Therefore, a distributed control γL2(ωNmax+1×(T,T)) solves the null controllability problem introduced in (EquationA1)–(EquationA2) if and only if it yields(A5) ωNmax+1×(T,T)ξγdxdt=-Ωφ(x,T)ξ(x,T)dx,ξ0L2(Ω)(A5)

The necessary condition in (EquationA5) is implied by (EquationA4) whereas the sufficient condition is obtained from observing that if the solution φ of (EquationA1) with a given γL2(ωNmax+1×(T,T)) satisfies (EquationA5) then, from (EquationA4) we get Ωφ(·,T)ξ0dx=0 for all ξ0L2(Ω). That leads to φ(·,T)=0 in Ω and thus, γ solves the null controllability problem (EquationA1)–(EquationA2).

Besides, let ξ^ be the solution of the adjoint problem introduced in (Equation71) using the initial data ξ^(·,T)=ξ^0 that yields the minimizer of the functional J:L2(Ω)IR which to a given ξ0 associates, for the coercivity and the strict convexity of J see [Citation13,Citation29]:(A6) J(ξ0)=12ξL2(ωNmax+1×(T,T))2+Ωφ(x,T)ξ(x,T)dx(A6)

Then, J(ξ^0)ξ0=ωNmax+1×(T,T)ξ^ξdxdt+Ωφ(x,T)ξ(x,T)dx=0 for all ξ0L2(Ω) and thus, in view of (EquationA5) the so-called Hilbert Uniqueness method (HUM) distributed control defined by: γ=ξ^ in ωNmax+1×(T,T) solves the null controllability problem (EquationA1)–(EquationA2). Furthermore, from setting ξ0=ξ^0 in (EquationA5) the HUM distributed control fulfills(A7) γL2(ωNmax+1×(T,T))Cφ(·,T)L2(Ω)(A7)

where C is a Carleman constant [Citation27]. Let φ^(x,t)=φ(x,T+T-t) in Ω×(T,T) where φ is the solution of the null controllability problem (EquationA1)–(EquationA2) involving the HUM distributed control γ=ξ^ in ωNmax+1×(T,T). Then, from multiplying the first equation in (Equation51) by φ^ and integrating by parts over Ω×(T,T) using Green’s formula we find(A8) Ωu(x,T)φ(x,T)dx=-ωNmax+1×(T,T)u(x,t)γ(x,T+T-t)dxdt(A8)

Moreover, using φ(·,T)=u(·,T) in Ω and according to (EquationA7)–(EquationA8) we getu(·,T)L2(Ω)CuL2(ωNmax+1×(T,T))

That is the result announced in Proposition 4.2.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.