This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Existence and uniqueness of solution to a hyperbolic-parabolic free boundary problem for biofilm growth

Dieudonné Zirhumanana Balike Luigi Frunzo Maria Rosaria Mattei Fabiana Russo
Abstract

This work presents the existence and uniqueness of solution to a free boundary value problem related to biofilm growth. The problem consists of a system of nonlinear hyperbolic partial differential equations governing the microbial species growth, and a system of parabolic partial differential equations describing the substrate dynamics. The free boundary evolution is governed by an ordinary differential equation that accounts for the thickness of the biofilm. We use the method of characteristics and fixed point strategies to prove the existence and uniqueness theorem in small and all times. All the equations are converted into integral equations, in particular this transformation is made for the parabolic equations by using the Green’s functions. We consider Dirichlet-Neumann and Neumann-Robin boundary conditions for the substrates equations and their extension to the case with variable diffusivity.

keywords:
Biofilms , free boundary problem , Existence and uniqueness , Integral equations, Modeling
MSC:
35A20 , 35A08 , 35A07 , 35A22, 35C15 , 35R35
\affiliation

[inst1]organization=Dipartimento di Matematica e Applicazioni ”Renato Caccioppoli”,Università degli di Studi di Napoli ”Federico II”,addressline=Via Cintia 1, Complesso di Monte Sant’Angelo, city=Naples, postcode=80126, state=Naples, country=Italy \affiliation[inst2]organization=Faculté des Sciences et Technologies, Université de Limoges,addressline=123 Avenue Albert Thomas, city=Limoges, postcode=87100, state=Haute-Vienne, country=France

1 Introduction

Biofilms are defined as a community of microorganisms embedded in a matrix of extracellular polymeric substances [1]. The experimental study of biofilms is relatively complex and expensive. Yet, a good understanding of the metabolic and chemical mechanisms taking place in and around the biofilm is important in several fields. In the medical field, for example, this understanding can facilitate the development of high-powered antibiotics since biofilms are known to be resistant to antibiotics. In the environmental field, biofilms are very important in wastewater treatment as they are responsible for various metabolic processes participating in waste degradation.
For several decades mathematical models have been proposed to contribute to the understanding of biofilms, evolving from 1D to 2D to 3D. Early models treated the biofilm as a continuum in one dimension and are exemplified by the founding stone biofilm model introduced by Wanner and Gujer [2]. Such model is based on a free boundary problem comprising a system of nonlinear hyperbolic systems which accounts for the biofilm growth, a system of diffusion-reaction equations for the substrates diffusion and conversion within the biofilm, and an ordinary differential equation that accounts for the biofilm thickness evolution. Although more sophisticated 2D and 3D models have been designed starting from the late 1990s with the aim to catch the highly complex heterogeneous 3D biofilm structure, these models have not found applications in engineering practice due to their complexity, and instead, there has been a pull-back towards 1D biofilm models for engineering applications [3].
In line with this trend, 1D Wanner-Gujer based models have been formulated and advocated for modelling biofilms in engineering systems. The phenomena modelled include, but are not limited to, biosorption [4], precipitation [5], invasion and attachment of new species [6], the role and interactions between the suspended microbial populations and the biofilm [7, 8]. In most of the models in literature, the forcing terms for the hyperbolic equations are modeled according to Monod kinetics which account for the microbial growth as a nonlinear function of the available nutrients. In the diffusion equations, these terms express the uptake or production of substrates by microbial activity. Due to the complexity of biofilm models, both with respect to mathematical structure and number of parameters, many works are limited to numerical simulations without ever considering the mathematical justification of the models or a rigorous analysis of its dynamics [9, 10, 7]. To cover this lack, recent works have been dedicated to the qualitative analysis of biofilm models of this type. Some of them focus on existence results for simplified models [11, 12]; others consider the pseudo stationarity of the diffusion equations due to time-scale arguments [13, 14].
In this work we study the existence and uniqueness of the complete 1D biofilm model as originally formulated in [2]. This study can be adapted to any kind of 1D biofilm from sorption to precipitation models. To achieve this, we use the method of characteristics and transpose the partial differential equations into integral systems. Then we state and prove an existence and uniqueness theorem of the solution to the problem in small and large time. Next, we briefly consider the case of the parabolic equation with a variable diffusion coefficient using the single layer potential.
The rest of this manuscript is organized as follows: in section 2, we present the model subject to this study, section 3 is devoted to the reduction of the PDEs to integral equations and to the existence theorem in small times. The extension of this result in large time is presented in section 4 while the case of parabolic equations with variable diffusivity case is discussed in section 6. We consider the Neumann-Robin boundary conditions in section 5 and conclude the paper and propose future recommendations in section 7.

2 The mathematical model

The biofilm model considered in this work is formulated as a free boundary problem made of a system of nonlinear hyperbolic PDEs, a system of quasi-linear parabolic PDEs, a PDE for the velocity growth of the microbial mass, and the free boundary is given by an ODE modelling the evolution of the biofilm thickness over time.
In the sequel, we will denote by

  • 1.

    Xi(z,t)=ρifi(z,t)X_{i}(z,t)=\rho_{i}f_{i}(z,t) the concentration of biomass i,i{1,..,n},𝑿=(X1,,Xn),i,\ i\in\{1,..,n\},\ \boldsymbol{X}=(X_{1},...,X_{n}), where fif_{i} are the volume fractions occupied by the biomasses in the biofilm and ρi\rho_{i} the density of the biomass, assumed constant for simplicity. In addition, the volume fractions fif_{i} are supposed to satisfy the incompressibility constraint

    i=1nfi=1;\sum_{i=1}^{n}f_{i}=1; (1)
  • 2.

    Cj(z,t),j{1,..,m},𝑪=(C1,,Cm),C_{j}(z,t),\ j\in\{1,..,m\},\ \boldsymbol{C}=(C_{1},...,C_{m}), the concentration of nutrients (dissolved components) in the biofilm;

  • 3.

    u(z,t)u(z,t) the convective growth velocity of the biofilm;

  • 4.

    rM,i(𝑿,𝑪)r_{M,i}(\boldsymbol{X},\boldsymbol{C}) the reaction term in the hyperbolic equations which accounts for the biomass growth and decay rate;

  • 5.

    rC,j(𝑿,𝑪)r_{C,j}(\boldsymbol{X},\boldsymbol{C}) the reaction term of the parabolic equations which accounts for the impact of the consumption/production of the substrate (nutrient) jj by the biomass;

  • 6.

    L(t)L(t) the thickness of the biofilm, which is the unknown boundary (free boundary) of the mathematical problem, and L0L_{0} the initial biofilm thickness.

The complete model considered can be stated as follows (details on the derivation and some applications can be found in [2]):

Xit+z(uXi)=H^i(𝑿,𝑪),i{1,..,n},0zL(t),t>0;\frac{\partial X_{i}}{\partial t}+\frac{\partial}{\partial z}(uX_{i})=\hat{H}_{i}(\boldsymbol{X},\boldsymbol{C}),~{}~{}~{}i\in\{1,..,n\},~{}~{}~{}0\leq z\leq L(t),~{}~{}t>0; (2)

with the following initial condition:

Xi(z,0)=ϕi(z),0zL0.X_{i}(z,0)=\phi_{i}(z),~{}~{}~{}0\leq z\leq L_{0}. (3)

By writing equations (2) in terms of volume fractions fi(z,t)f_{i}(z,t), summing over ii and using the constraint (1), we get the velocity equation below:

uz=i=1nH^i(𝑿,𝑪)=R^i(𝑿,𝑪),0<zL(t),t0,\frac{\partial u}{\partial z}=\sum_{i=1}^{n}\hat{H}_{i}(\boldsymbol{X},\boldsymbol{C})=\hat{R}_{i}(\boldsymbol{X},\boldsymbol{C}),~{}~{}~{}0<z\leq L(t),~{}~{}~{}t\geq 0,\\ (4)

with boundary condition reflecting the zero biomass flux at the substratum (z=0z=0)

u(0,t)=0,t0.u(0,t)=0,~{}~{}t\geq 0. (5)

The equation for the thickness is given by:

dLdt=u(L(t),t)+σ(t),L(0)=L0,t>0.\frac{dL}{dt}=u(L(t),t)+\sigma(t),~{}~{}~{}L(0)=L_{0},~{}~{}t>0. (6)

where σ(t)\sigma(t) is the attachment and detachment function which can assume positive or negative values. In last position we have the system of diffusion equations for the substrates, with initial and boundary conditions:

CjtDj2Cjz2=F^j(𝑿,𝑪)j{1,..,m},0<z<L(t),t>0Cj(z,0)=φj(z),j{1,..,m},0zL0,Cjz(0,t)=0,j{1,..,m},t>0,Cj(L(t),t)=ψj(t),j{1,..,m},t>0.\begin{array}[]{l}\frac{\partial C_{j}}{\partial t}-D_{j}\frac{\partial^{2}C_{j}}{\partial z^{2}}=\hat{F}_{j}(\boldsymbol{X},\boldsymbol{C})~{}~{}~{}j\in\{1,..,m\},~{}~{}~{}0<z<L(t),~{}~{}t>0\\ \\ C_{j}(z,0)=\varphi_{j}^{*}(z),~{}~{}~{}j\in\{1,..,m\},~{}~{}~{}0\leq z\leq L_{0},\\ \\ \frac{\partial C_{j}}{\partial z}(0,t)=0,~{}~{}~{}j\in\{1,..,m\},~{}~{}t>0,\\ \\ C_{j}(L(t),t)=\psi^{*}_{j}(t),~{}~{}~{}j\in\{1,..,m\},~{}~{}~{}t>0.\end{array} (7)

3 Reduction to integral equations and local existence and uniqueness

The diffusion system in a planar biofilm is now given by 7. We make a change of variable by setting Sj=DjCjS_{j}=D_{j}C_{j}, and after substituting in (7) we get the following equation:

Sjt2Sjz2=Fj(𝑿,𝑺),0<z<L(t),\displaystyle\frac{\partial S_{j}}{\partial t}-\frac{\partial^{2}S_{j}}{\partial z^{2}}=F_{j}(\boldsymbol{X},\boldsymbol{S}),~{}~{}~{}0<z<L(t), (8a)
Sj(z,0)=φj(z),0zL0,\displaystyle S_{j}(z,0)=\varphi_{j}(z),~{}~{}~{}0\leq z\leq L_{0}, (8b)
Sjz(0,t)=0,t>0\displaystyle\frac{\partial S_{j}}{\partial z}(0,t)=0,~{}~{}t>0 (8c)
Sj(L(t),t)=ψj(t),t>0;\displaystyle S_{j}(L(t),t)=\psi_{j}(t),~{}~{}~{}t>0; (8d)

where Fj(𝑿,𝑺)=F^j(𝑿,𝑪,D),φj(z)=Djφj(z)F_{j}(\boldsymbol{X},\boldsymbol{S})=\hat{F}_{j}(\boldsymbol{X},\boldsymbol{C},D),~{}~{}\varphi_{j}(z)=D_{j}\varphi_{j}^{*}(z) and ψj(t)=Djψj(t).\psi_{j}(t)=D_{j}\psi_{j}^{*}(t). In the sequel, we will also use H~i(𝑿,𝑺)=H^i(𝑿,𝑪,D),\tilde{H}_{i}(\boldsymbol{X},\boldsymbol{S})=\hat{H}_{i}(\boldsymbol{X},\boldsymbol{C},D), and R(𝑿,𝑺)=R^(𝑿,𝑪,D)R(\boldsymbol{X},\boldsymbol{S})=\hat{R}(\boldsymbol{X},\boldsymbol{C},D) and the notation SjxS_{jx} will stand for Sjx,\frac{\partial S_{j}}{\partial x}, with xx any variable when no confusion is possible.
Consider the following kernel of the heat equation [15, 16, 17] :

K(z,t;ξ,τ)=14π(tτ)exp((zξ)24(tτ)).K(z,t;\xi,\tau)=\frac{1}{\sqrt{4\pi(t-\tau)}}\exp\left(-\frac{(z-\xi)^{2}}{4(t-\tau)}\right). (9)

Define the following Green and Neumann functions in the half plane z>0z>0 respectively given by:

G(z,t;ξ,τ)=K(z,t;ξ,τ)K(z,t;ξ,τ)G(z,t;\xi,\tau)=K(z,t;\xi,\tau)-K(-z,t;\xi,\tau) (10)
N(z,t;ξ,τ)=K(z,t;ξ,τ)+K(z,t;ξ,τ)N(z,t;\xi,\tau)=K(z,t;\xi,\tau)+K(-z,t;\xi,\tau) (11)

To derive the solution of (8), we integrate the following Green’s identity on 0<ξ<L(t)0<\xi<L(t), 0<τ<t0<\tau<t:

ξ(G(z,t;ξ,τ)Sjξ(ξ,τ)Gξ(z,t;ξ,τ)Sj(ξ,τ))τ(G(z,t;ξ,τ)Sj(ξ,τ))=\displaystyle\frac{\partial}{\partial\xi}\left(G(z,t;\xi,\tau)S_{j\xi}(\xi,\tau)-G_{\xi}(z,t;\xi,\tau)S_{j}(\xi,\tau)\right)-\frac{\partial}{\partial\tau}\left(G(z,t;\xi,\tau)S_{j}(\xi,\tau)\right)=
G(z,t;ξ,τ)Fj(𝑿,𝑺)\displaystyle-G(z,t;\xi,\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})

After introducing the boundary and initial conditions, we get:

Sj(z,t)=0L0φj(ξ)G(z,t;ξ,0)𝑑ξ0tψj(τ)Gξ(z,t;L(τ),τ)𝑑τ+0tθj(τ)G(z,t;L(τ),τ)𝑑τ+0tΦj(τ)Gξ(z,t;0,τ)𝑑τ+0t0L(t)G(z,t;ξ,τ)Fj(𝑿,𝑺)𝑑ξ𝑑τ,\begin{array}[]{l }S_{j}(z,t)=\int_{0}^{L_{0}}\varphi_{j}(\xi)G(z,t;\xi,0)d\xi-\int_{0}^{t}\psi_{j}(\tau)G_{\xi}(z,t;L(\tau),\tau)d\tau+\\ \\ \int_{0}^{t}\theta_{j}(\tau)G(z,t;L(\tau),\tau)d\tau+\int_{0}^{t}\Phi_{j}(\tau)G_{\xi}(z,t;0,\tau)d\tau\\ \\ +\int_{0}^{t}\int_{0}^{L(t)}G(z,t;\xi,\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\xi d\tau,\end{array} (12)

where

θj(t)=Sjz(L(t),t),\theta_{j}(t)=S_{jz}(L(t),t), (13)
Φj(t)=Sj(0,t),\Phi_{j}(t)=S_{j}(0,t), (14)

and in equation (12), we have to determine Φj\Phi_{j} and θj\theta_{j}. By letting z0z\rightarrow 0 in equation (12) we get:

Φj(t)=0tGξ(0,t,L(τ),τ)ψj(τ)𝑑τ.\Phi_{j}(t)=-\int_{0}^{t}G_{\xi}(0,t,L(\tau),\tau)\psi_{j}(\tau)d\tau. (15)

Next, we use the following property of the function K which is given in Theorem 1 of Chapter 5, page 137 of [15] and in Lemma 4.2.1, page 50 of [16] :

limzL(t)z0tK(z,t;L(τ),τ)g(τ)𝑑τ=12g(t)+0tg(τ)zK(z,t,L(τ),τ)|z=L(t)dτ,\left.\lim_{z\rightarrow L(t)}\frac{\partial}{\partial z}\int_{0}^{t}K(z,t;L(\tau),\tau)g(\tau)d\tau=\frac{1}{2}g(t)+\int_{0}^{t}g(\tau)\frac{\partial}{\partial z}K(z,t,L(\tau),\tau)\right|_{z=L(t)}d\tau, (16)

where gg is a continuous function. Taking the derivative of SjS_{j} over zz in the domain 0<ξ<L(τ)0<\xi<L(\tau), 0<τ<tϵ0<\tau<t-\epsilon and letting ϵ0\epsilon\rightarrow 0, and using relation 16 (or equivalently if we let zL(t)z\rightarrow L(t)), we get:

θj(t)=Uj(t)+20L0φj(ξ)N(L(t),t,ξ,0)𝑑ξ+20tψ˙j(τ)N(L(t),t,L(τ),τ)𝑑τ+20tθj(τ)Gz(L(t),t,L(τ),τ)𝑑τ20tΦ˙j(τ)N(L(t),t;0,τ)𝑑τ20tN(L(t),t,L(τ),τ)Fj(𝑿,𝑺)𝑑τ+20tN(L(t),t,0,τ)Fj(𝑿,𝑺)𝑑τ,\begin{array}[]{l l}\theta_{j}(t)=&U_{j}(t)+2\int_{0}^{L_{0}}\varphi_{j}^{\prime}(\xi)N(L(t),t,\xi,0)d\xi+2\int_{0}^{t}\dot{\psi}_{j}(\tau)N(L(t),t,L(\tau),\tau)d\tau\\ &\\ &+2\int_{0}^{t}\theta_{j}(\tau)G_{z}(L(t),t,L(\tau),\tau)d\tau-2\int_{0}^{t}\dot{\Phi}_{j}(\tau)N(L(t),t;0,\tau)d\tau\\ &\\ &-2\int_{0}^{t}N(L(t),t,L(\tau),\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\tau+2\int_{0}^{t}N(L(t),t,0,\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\tau,\end{array} (17)

where

Uj(t)=2N(L(t),t,0,0)[φj(0)Φj(0)]2ψj(0)N(L(t),t,L0,0).U_{j}(t)=2N(L(t),t,0,0)[\varphi_{j}(0)-\Phi_{j}(0)]-2\psi_{j}(0)N(L(t),t,L_{0},0). (18)

In this way we have just proved that for any solution SjS_{j} of (8) for all t<λt<\lambda, θj\theta_{j} (defined by (13)) has to satisfy the integral equation (17), where L(t)L(t) is given by (19) (for σ=0\sigma=0) below:

L(t)=L0+0tu(L(τ),τ)𝑑τ (for σ(t)=0,t).L(t)=L_{0}+\int_{0}^{t}u(L(\tau),\tau)d\tau~{}~{}~{}\text{ (for $\sigma(t)=0,~{}~{}\forall t$).} (19)

Conversely, we consider that for some λ>0,\lambda>0, θj(t)\theta_{j}(t) is a continuous solution of the integral equation (17) for 0t<λ0\leq t<\lambda and then we prove that Sj(z,t)S_{j}(z,t) form a solution of equations (8) for L(t)>0L(t)>0 for all t<λt<\lambda.
We can easily check that Sj(z,t)S_{j}(z,t) satisfies (8a), (8b), and (8d). To verify that it satisfies the boundary condition (8c), we integrate Green’s identity 12 as before and we get:

Sj(z,t)=0L0φj(ξ)G(z,t;ξ,0)𝑑ξ0tψj(τ)Gξ(z,t;L(τ),τ)𝑑τ0tρj(τ)G(z,t;0,τ)𝑑τ+0tΦj(τ)Gξ(z,t;0,τ)𝑑τ+0tθj(τ)G(z,t;L(τ),τ)𝑑τ+0t0L(t)G(z,t;ξ,τ)Fj(𝑿,𝑺)𝑑ξ𝑑τ\begin{array}[]{l }S_{j}(z,t)=\int_{0}^{L_{0}}\varphi_{j}(\xi)G(z,t;\xi,0)d\xi-\int_{0}^{t}\psi_{j}(\tau)G_{\xi}(z,t;L(\tau),\tau)d\tau\\ \\ -\int_{0}^{t}\rho_{j}(\tau)G(z,t;0,\tau)d\tau+\int_{0}^{t}\Phi_{j}(\tau)G_{\xi}(z,t;0,\tau)d\tau+\int_{0}^{t}\theta_{j}(\tau)G(z,t;L(\tau),\tau)d\tau\\ \\ +\int_{0}^{t}\int_{0}^{L(t)}G(z,t;\xi,\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\xi d\tau\end{array} (20)

where ρj(t)=Sjz(0,t).\rho_{j}(t)=S_{jz}(0,t). Equation (20) is the same as (12) if the third integral of the right hand side is null. Since G(z,t,0,τ)G(z,t,0,\tau) vanishes, the third integral is zero. So, the condition is fulfilled. However, we still need to prove that ρj(t)0\rho_{j}(t)\equiv 0 so that the boundary condition is satisfied. We thus have

0tρj(τ)G(z,t;0,τ)𝑑τ=0,\int_{0}^{t}\rho_{j}(\tau)G(z,t;0,\tau)d\tau=0, (21)

for 0<z<L(t)0<z<L(t), 0<t<λ.0<t<\lambda.
We recall that

|G(z,t;ξ,τ)|1|Gz(z,t;ξ,τ)||zξ|tτ.|G(z,t;\xi,\tau)|\leq 1\leq|G_{z}(z,t;\xi,\tau)|\leq\frac{|z-\xi|}{\sqrt{t-\tau}}.

Hence, for any continuous function hh

|0tG(z,t;ξ,τ)h(τ)𝑑τ|h.|\int_{0}^{t}G(z,t;\xi,\tau)h(\tau)d\tau|\leq\|h\|.

Since ρj\rho_{j} is continuous and bounded, we can write

limξ0|0tG(z,t;ξ,τ)ρj(τ)𝑑τ|ρj\lim_{\xi\rightarrow 0}|\int_{0}^{t}G(z,t;\xi,\tau)\rho_{j}(\tau)d\tau|\leq\|\rho_{j}\|

and in particular that

limξ0|0tG(z,t;ξ,τ)ρj(τ)𝑑τ|limξ0|0tzξtτρj(τ)𝑑τ|.\lim_{\xi\rightarrow 0}|\int_{0}^{t}G(z,t;\xi,\tau)\rho_{j}(\tau)d\tau|\leq\lim_{\xi\rightarrow 0}\big{|}\int_{0}^{t}\frac{z-\xi}{\sqrt{t-\tau}}\rho_{j}(\tau)d\tau\big{|}.

Using the continuity of ρj\rho_{j}, we get

|ρj(t)|C0t|ρj(τ)|tτ𝑑τC20tdτtτ0τ|ρj(ζ)|τζ𝑑ζ|\rho_{j}(t)|\leq C\int_{0}^{t}\frac{|\rho_{j}(\tau)|}{\sqrt{t-\tau}}d\tau\leq C^{2}\int_{0}^{t}\frac{d\tau}{\sqrt{t-\tau}}\int_{0}^{\tau}\frac{|\rho_{j}(\zeta)|}{\sqrt{\tau-\zeta}}d\zeta
=C20t|ρj(ζ)|𝑑ζζtdτ(tτ)(τζ)=πC20t|ρj(ζ)|𝑑ζ,=C^{2}\int_{0}^{t}|\rho_{j}(\zeta)|d\zeta\int_{\zeta}^{t}\frac{d\tau}{\sqrt{(t-\tau)(\tau-\zeta)}}=\pi C^{2}\int_{0}^{t}|\rho_{j}(\zeta)|d\zeta,

with C=C(t).C=C(t).
Using Grönwall’s lemma, we conclude that ρ(t)=0,t\rho(t)=0,~{}\forall t and thus, the boundary condition (8c) is satisfied.
In summary, we have proved the following theorem:

Theorem 1.

Let φC1[0,L0],\varphi\in C^{1}[0,L_{0}], ψC1[0,λ],\psi\in C^{1}[0,\lambda], and FF a Lipschitz function on [0,λ][0,\lambda] for some λ>0.\lambda>0. Let L(t)L(t) be given by equation (19) and assume L(t)>0,t[0,λ].L(t)>0,~{}~{}\forall t\in[0,\lambda]. The solution of the equations (8) is given by (12) where Φj\Phi_{j} and θj\theta_{j} must satisfy equations (15) and (17) respectively.

Next, we will derive integral equations for the hyperbolic equations rewritten here in compact form:

Xit+z(uXi)=H~i(𝑿,𝑺),0zL(t),t>0,\frac{\partial X_{i}}{\partial t}+\frac{\partial}{\partial z}(uX_{i})=\tilde{H}_{i}(\boldsymbol{X},\boldsymbol{S}),~{}~{}~{}0\leq z\leq L(t),~{}~{}t>0, (22)

with the following initial condition:

Xi(z,0)=ϕi(z),0zL0.X_{i}(z,0)=\phi_{i}(z),~{}~{}~{}0\leq z\leq L_{0}. (23)

Let us denote the characteristics of this system by η(z0,t)\eta(z_{0},t) so that they satisfy the following initial value problem:

ηt(z0,t)=u(η(z0,t),t),η(z0,0)=z0.\frac{\partial\eta}{\partial t}(z_{0},t)=u(\eta(z_{0},t),t),~{}~{}~{}\eta(z_{0},0)=z_{0}. (24)

Considering equation (6) for σ(t)=0\sigma(t)=0, it follows immediately that L(t)=η(L0,t).L(t)=\eta(L_{0},t). Thus, to get the integral equation for (22)–(23), we will use the characteristics coordinates z=η(z0,t)z=\eta(z_{0},t) that will be noted by η\eta for simplicity:

u(η,t)=0z0R(𝑿,𝑺)ηζ0𝑑ζ0,u(\eta,t)=\int_{0}^{z_{0}}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial\zeta_{0}}d\zeta_{0}, (25)
Xi(η,t)=ϕi(z0)+0tHi(𝑿,𝑺)𝑑τ,X_{i}(\eta,t)=\phi_{i}(z_{0})+\int_{0}^{t}H_{i}(\boldsymbol{X},\boldsymbol{S})d\tau, (26)
η(z0,t)=z0+0tu(η,τ)𝑑τ=z0+0t𝑑τ0z0R(𝑿,𝑺)ηζ0𝑑ζ0,\eta(z_{0},t)=z_{0}+\int_{0}^{t}u(\eta,\tau)d\tau=z_{0}+\int_{0}^{t}d\tau\int_{0}^{z_{0}}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial\zeta_{0}}d\zeta_{0}, (27)
ηz0=1+0tR(𝑿,𝑺)ηz0𝑑τ,\frac{\partial\eta}{\partial z_{0}}=1+\int_{0}^{t}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial z_{0}}d\tau, (28)

with Hi(𝑿,𝑺)=H~i(𝑿,𝑺)XiR(𝑿,𝑺).H_{i}(\boldsymbol{X},\boldsymbol{S})=\tilde{H}_{i}(\boldsymbol{X},\boldsymbol{S})-X_{i}R(\boldsymbol{X},\boldsymbol{S}). The integral equation for the free boundary is given by:

L(t)=L0+0t0𝑑τ0L0R(𝑿,𝑺)ηζ0𝑑ζ0.L(t)=L_{0}+\int_{0}^{t_{0}}d\tau\int_{0}^{L_{0}}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial\zeta_{0}}d\zeta_{0}. (29)

Finally, we convert equations (12) and (17) into characteristics coordinates. When z=η(z0,t),z=\eta(z_{0},t), L(t)=η(L0,t)L(t)=\eta(L_{0},t), L(τ)=η(L0,τ)L(\tau)=\eta(L_{0},\tau) and ξ=η(ξ0,t)\xi=\eta(\xi_{0},t) (to simplify the writing, we will drop the variable in brackets and write η\eta and η~\tilde{\eta} for η(z0,t)\eta(z_{0},t) and η(ξ0,t)\eta(\xi_{0},t) respectively when no confusion is possible), we have:

Sj(η,t)=0L0φj(η~)G(η,t;η~,0)η~ξ0𝑑ξ00tψj(τ)Gξ(η,t;η(L0,τ),τ)η~ξ0𝑑τ+0tθj(τ)G(η,t;η(L0,τ),τ)𝑑τ+0tΦj(τ)Gξ(η,t;0,τ)𝑑τ+0t0L0G(η,t;η~,τ)Fj(𝑿,𝑺)η~ξ0𝑑ξ0𝑑τ\begin{array}[]{l }S_{j}(\eta,t)=\int_{0}^{L_{0}}\varphi_{j}(\tilde{\eta})G(\eta,t;\tilde{\eta},0)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\xi_{0}-\int_{0}^{t}\psi_{j}(\tau)G_{\xi}(\eta,t;\eta(L_{0},\tau),\tau)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\tau\\ \\ +\int_{0}^{t}\theta_{j}(\tau)G(\eta,t;\eta(L_{0},\tau),\tau)d\tau+\int_{0}^{t}\Phi_{j}(\tau)G_{\xi}(\eta,t;0,\tau)d\tau\\ \\ +\int_{0}^{t}\int_{0}^{L_{0}}G(\eta,t;\tilde{\eta},\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\xi_{0}d\tau\end{array} (30)

and the equation for θj(t)\theta_{j}(t) is

θj(t)=Uj(t)+20L0φj(η~)N(η(L0,t),t,η~,0)η~ξ0𝑑ξ0+20tψ˙j(τ)N(η(L0,t),t,η(L0,τ),τ)𝑑τ+20tθj(τ)Gξ(η(L0,t),t,η(L0,τ),τ)ηz0𝑑τ20tΦ˙j(τ)N(η(L0,t),t,0,τ)𝑑τ20tN(η(L0,t),t,η(L0,τ),τ)Fj(𝑿,𝑺)𝑑τ+20tN(η(L0,t),t,0,τ)Fj(𝑿0,𝑺0)𝑑τ,\begin{array}[]{l}\theta_{j}(t)=U_{j}(t)+2\int_{0}^{L_{0}}\varphi_{j}^{\prime}(\tilde{\eta})N(\eta(L_{0},t),t,\tilde{\eta},0)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\xi_{0}\\ \\ +2\int_{0}^{t}\dot{\psi}_{j}(\tau)N(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)d\tau+2\int_{0}^{t}\theta_{j}(\tau)G_{\xi}(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)\frac{\partial\eta}{\partial z_{0}}d\tau\\ \\ -2\int_{0}^{t}\dot{\Phi}_{j}(\tau)N(\eta(L_{0},t),t,0,\tau)d\tau-2\int_{0}^{t}N(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\tau\\ \\ +2\int_{0}^{t}N(\eta(L_{0},t),t,0,\tau)F_{j}(\boldsymbol{X}^{0},\boldsymbol{S}^{0})d\tau,\end{array} (31)

where 𝑿0\boldsymbol{X}^{0} and 𝑺0\boldsymbol{S}^{0} denote respectively 𝑿(0,t)\boldsymbol{X}(0,t) and 𝑺0=(Φ1(t),,Φm(t))\boldsymbol{S}^{0}=(\Phi_{1}(t),...,\Phi_{m}(t)). Moreover, Uj(t)U_{j}(t) in characteristic coordinates assumes the following form:

Uj(t)=2N(η(L0,t),t,0,0)[φj(0)Φj(0)]2ψj(0)N(η(L0,t),t,L0,0).U_{j}(t)=2N(\eta(L_{0},t),t,0,0)[\varphi_{j}(0)-\Phi_{j}(0)]-2\psi_{j}(0)N(\eta(L_{0},t),t,L_{0},0).

Before we prove the existence and uniqueness theorem, we need the following lemmas to use the Banach-Caccioppoli fixed point theorem.
We consider the space of continuous functions

Cλ,M={𝒰=(u1(η,t)u2(η,t)u3(η,t)u4(η,t))ui:Ω=[0,L0]×[0,λ],ui continuous and 𝒰ΩM}C_{\lambda,M}=\left\{\mathcal{U}=\begin{pmatrix}u_{1}(\eta,t)\\ u_{2}(\eta,t)\\ u_{3}(\eta,t)\\ u_{4}(\eta,t)\end{pmatrix}\mid u_{i}:\Omega=[0,L_{0}]\times[0,\lambda]\rightarrow\mathbb{R},u_{i}\text{ continuous and }\|\mathcal{U}\|_{\Omega}\leq M\right\}

where .Ω\|.\|_{\Omega} is given by

𝒰Ω:=i=14max(η,t)Ω|ui|.\|\mathcal{U}\|_{\Omega}:=\sum_{i=1}^{4}\max_{(\eta,t)\in\Omega}|u_{i}|.

We choose the vector 𝒰\mathcal{U} such that 𝒰=(u1(η,t)u2(η,t)u3(η,t)u4(η,t))=(u(η,t)Xi(η,t)η(z0,t)θj(η,t))\mathcal{U}=\begin{pmatrix}u_{1}(\eta,t)\\ u_{2}(\eta,t)\\ u_{3}(\eta,t)\\ u_{4}(\eta,t)\end{pmatrix}=\begin{pmatrix}u(\eta,t)\\ X_{i}(\eta,t)\\ \eta(z_{0},t)\\ \theta_{j}(\eta,t)\end{pmatrix} and we define the map 𝒯:Cλ,MCλ,M\mathcal{T}:C_{\lambda,M}\rightarrow C_{\lambda,M} such that 𝒯(𝒰(η,t))=(𝒯1(𝒰(η,t))𝒯2(𝒰(η,t))𝒯3(𝒰(η,t))𝒯4(𝒰(η,t)))\mathcal{T}(\mathcal{U}(\eta,t))=\begin{pmatrix}\mathcal{T}_{1}\left(\mathcal{U}(\eta,t)\right)\\ \mathcal{T}_{2}\left(\mathcal{U}(\eta,t)\right)\\ \mathcal{T}_{3}\left(\mathcal{U}(\eta,t)\right)\\ \mathcal{T}_{4}\left(\mathcal{U}(\eta,t)\right)\\ \end{pmatrix}

where 𝒯i(𝒰(η,t))\mathcal{T}_{i}\left(\mathcal{U}(\eta,t)\right) (i=1,,4i=1,\cdots,4) are respectively given by

𝒯1(𝒰(η,t))=ϕi(z0)+0tHi(𝑿,𝑺)𝑑τ𝒯2(𝒰(η,t))=z0+0t0z0R(𝑿,𝑺)ηζ0𝑑ζ0𝑑τ𝒯3(𝒰(η,t))=1+0tR(𝑿,𝑺)ηz0𝑑τ𝒯4(𝒰(η,t))=Uj(t)+20L0φj(η~)N(η(L0,t),t,η~,0)η~ξ0𝑑ξ0+20tψ˙j(τ)N(η(L0,t),t,η(L0,τ),τ)𝑑τ+20tθj(τ)Gξ(η(L0,t),t,η(L0,τ),τ)ηz0𝑑τ20tΦ˙j(τ)N(η(L0,t),t,0,τ)𝑑τ20tN(η(L0,t),t,η(L0,τ),τ)Fj(𝑿,𝑺)𝑑τ+20tN(η(L0,t),t,0,τ)Fj(𝑿0,𝑺0)𝑑τ.content\begin{array}[]{l}\mathcal{T}_{1}\left(\mathcal{U}(\eta,t)\right)=\phi_{i}(z_{0})+\int_{0}^{t}H_{i}(\boldsymbol{X},\boldsymbol{S})d\tau\\ \\ \mathcal{T}_{2}\left(\mathcal{U}(\eta,t)\right)=z_{0}+\int_{0}^{t}\int_{0}^{z_{0}}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial\zeta_{0}}d\zeta_{0}d\tau\\ \\ \mathcal{T}_{3}\left(\mathcal{U}(\eta,t)\right)=1+\int_{0}^{t}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial z_{0}}d\tau\\ \\ \mathcal{T}_{4}\left(\mathcal{U}(\eta,t)\right)=U_{j}(t)+2\int_{0}^{L_{0}}\varphi_{j}^{\prime}(\tilde{\eta})N(\eta(L_{0},t),t,\tilde{\eta},0)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\xi_{0}\\ \\ +2\int_{0}^{t}\dot{\psi}_{j}(\tau)N(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)d\tau+2\int_{0}^{t}\theta_{j}(\tau)G_{\xi}(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)\frac{\partial\eta}{\partial z_{0}}d\tau\\ \\ -2\int_{0}^{t}\dot{\Phi}_{j}(\tau)N(\eta(L_{0},t),t,0,\tau)d\tau-2\int_{0}^{t}N(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\tau\\ \\ +2\int_{0}^{t}N(\eta(L_{0},t),t,0,\tau)F_{j}(\boldsymbol{X}^{0},\boldsymbol{S}^{0})d\tau.content...\end{array}

We shall prove the properties of 𝒯\mathcal{T}. We first need the following lemma whose proof is immediate.

Lemma 2.

Let uiC0([0,L0]×[0,λ]),maxΩ|ui(η,t)|Mu_{i}\in C^{0}([0,L_{0}]\times[0,\lambda]),\max_{\Omega}|u_{i}(\eta,t)|\leq M and 2MλL02M\lambda\leq L_{0} then L(t)L(t) defined by (29) satisfies

L(t)L(τ)M(tτ),τ,t[0,λ],\mid L(t)-L(\tau)\mid\leq M(t-\tau),~{}~{}~{}\forall\tau,t\in[0,\lambda], (32)
L(t)L0L02,t[0,λ].\mid L(t)-L_{0}\mid\leq\frac{L_{0}}{2},~{}~{}~{}\forall t\in[0,\lambda]. (33)
L02L(t)3L02\frac{L_{0}}{2}\leq L(t)\leq\frac{3L_{0}}{2} (34)

P1. 𝒯\mathcal{T} maps Cλ,MC_{\lambda,M} into itself. In what follows, we will mainly focus on the fourth component of 𝒯(𝒰(η,t)).\mathcal{T}(\mathcal{U}(\eta,t)). The proof for the other components is similar to the one proposed in [14], so we skip it.

Lemma 3.

Let λ1,\lambda\leq 1, M1M\geq 1, φjC1[0,L0],\varphi_{j}\in C^{1}[0,L_{0}], ψjC1[0,λ],\psi_{j}\in C^{1}[0,\lambda], and FjF_{j} a Lipschitz over C0(+)C^{0}(\mathbb{R}^{+}) with a Lipschitz constant K>0.K>0. Under the hypothesis of Lemma 34 we have the following properties

Uj(t)φj(0)\|U_{j}(t)\|\leq\|\varphi_{j}(0)\| (35)
0L0φj(η~)N(η(L0,t),t,η~,0)η~ξ0𝑑ξ0φj\int_{0}^{L_{0}}\mid\varphi_{j}^{\prime}(\tilde{\eta})N(\eta(L_{0},t),t,\tilde{\eta},0)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}\mid d\xi_{0}\leq\|\varphi_{j}^{\prime}\| (36)
0tψ˙j(τ)N(η(L0,t),t,η(L0,τ),τ)𝑑τψ˙jπλ.\int_{0}^{t}\mid\dot{\psi}_{j}(\tau)N(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)\mid d\tau\leq\frac{\|\dot{\psi}_{j}\|}{\sqrt{\pi}}\sqrt{\lambda}. (37)
0tθj(τ)Gz(η(L0,t),t,η(L0,τ),τ)ηz0𝑑τ14π(Mtτ+3L0(23eL02)3)\int_{0}^{t}\mid\theta_{j}(\tau)G_{z}(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)\frac{\partial\eta}{\partial z_{0}}\mid d\tau\leq\frac{1}{4\sqrt{\pi}}\left(\frac{M}{\sqrt{t-\tau}}+3L_{0}\sqrt{\left(\frac{2}{3eL_{0}^{2}}\right)^{3}}\right) (38)
0tΦ˙j(τ)N(η(L0,t),t,0,τ)𝑑τΦ˙j(τ)πλ\int_{0}^{t}\mid\dot{\Phi}_{j}(\tau)N(\eta(L_{0},t),t,0,\tau)\mid d\tau\leq\frac{\|\dot{\Phi}_{j}(\tau)\|}{\sqrt{\pi}}\sqrt{\lambda} (39)
0t|N(η(L0,t),t,η(L0,τ),τ)Fj(𝑿,𝑺)|𝑑τ2Kπλ.\int_{0}^{t}|N(\eta(L_{0},t),t,\eta(L_{0},\tau),\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})|d\tau\leq\frac{2K}{\sqrt{\pi}}\sqrt{\lambda}. (40)
0t|N(η(L0,t),t,0,τ)Fj(𝑿0,𝑺0)|𝑑τKλL02πe\int_{0}^{t}\big{|}N(\eta(L_{0},t),t,0,\tau)F_{j}(\boldsymbol{X}^{0},\boldsymbol{S}^{0})\big{|}d\tau\leq\frac{K\lambda}{L_{0}}\sqrt{\frac{2}{\pi e}} (41)

Proof: To prove (35), we choose aa such that 0<a<x<λ.0<a<x<\lambda.
We have

12πtexp(x24t)12πtexp(a24t)1.\frac{1}{2\sqrt{\pi t}}\exp\left(\frac{-x^{2}}{4t}\right)\leq\frac{1}{2\sqrt{\pi t}}\exp\left(\frac{-a^{2}}{4t}\right)\leq 1.

By using this last inequality and the definition of Uj,U_{j}, the inequality (35) follows easily. The inequality 36 is immediate and can be found in [18], same for 38.
To prove 37, we use the definition 11 and the inequality holds. The inequality 39 can also be proven in the same way. Finally, inequality 40 holds because FjF_{j} is Lipschitz and N(x,t,ξ,τ)1π(tτ).N(x,t,\xi,\tau)\leq\frac{1}{\sqrt{\pi(t-\tau)}}.
To prove inequality 41 we use the following exponential inequality [18, 19]:

exp(x2α(tτ))(tτ)n2(nα2ex2)n2 for α,x,t>τ and n\frac{\exp\bigg{(}\frac{-x^{2}}{\alpha(t-\tau)}\bigg{)}}{(t-\tau)^{\frac{n}{2}}}\leq\bigg{(}\frac{n\alpha}{2ex^{2}}\bigg{)}^{\frac{n}{2}}\text{ for }\alpha,~{}x,t>\tau\text{ and }n\in\mathbb{N} (42)

We have

0t|N(η(L0,t),t;0,τ)F(𝑿0,𝑺0)dτ||0t1π(tτ)exp(η(L0,t)24(tτ))F(𝑿0,𝑺0)𝑑τ||0t1π(tτ)2eη(L0,t)2F(𝑿0,𝑺0)𝑑τ||0tKL02πe𝑑τ|KλL02πe\begin{array}[]{ll}\int_{0}^{t}\big{|}N(\eta(L_{0},t),t;0,\tau)F(\boldsymbol{X}^{0},\boldsymbol{S}^{0})d\tau\big{|}\leq&\bigg{|}\int_{0}^{t}\frac{1}{\sqrt{\pi(t-\tau)}}\exp\bigg{(}\frac{-\eta(L_{0},t)^{2}}{4(t-\tau)}\bigg{)}F(\boldsymbol{X}^{0},\boldsymbol{S}^{0})d\tau\bigg{|}\\ &\\ &\leq\bigg{|}\int_{0}^{t}\frac{1}{\sqrt{\pi(t-\tau)}}\sqrt{\frac{2}{e~{}\eta(L_{0},t)^{2}}}F(\boldsymbol{X}^{0},\boldsymbol{S}^{0})d\tau\bigg{|}\\ &\\ &\leq\bigg{|}\int_{0}^{t}\frac{K}{L_{0}}\sqrt{\frac{2}{\pi e}}d\tau\bigg{|}\leq\frac{K\lambda}{L_{0}}\sqrt{\frac{2}{\pi e}}\\ \end{array}

P2. 𝒯\mathcal{T} is a contraction. To show that 𝒯\mathcal{T} is contraction in its fourth component, we consider that T1=𝒯θj1T_{1}=\mathcal{T}\theta_{j1} and T2=𝒯θj2T_{2}=\mathcal{T}\theta_{j2} and denote

θj1θj2=δ\mid\theta_{j1}-\theta_{j2}\mid=\delta (43)

If we consider two functions L1L_{1} and L2L_{2} corresponding respectively to θj1\theta_{j1} and θj2\theta_{j2} but by means of 29, then we have the following inequalities:

L1(t)L2(t)δt\mid L_{1}(t)-L_{2}(t)\mid\leq\delta t (44)
L˙1(t)L˙2(t)δ\mid\dot{L}_{1}(t)-\dot{L}_{2}(t)\mid\leq\delta (45)

Note that L1L_{1} and L2L_{2} also satisfy inequality 34.

Lemma 4.

Let φ[0,L0],ψC0[0,λ],\varphi\in[0,L_{0}],~{}\psi\in C^{0}[0,\lambda], and F a Lipschitz function over C0[0,L0]×[0,λ].C^{0}[0,L_{0}]\times[0,\lambda]. Then we have:

Uj1(t)Uj2(t)C1(φj(0)+ψj(0)).\mid\mid U_{j1}(t)-U_{j2}(t)\mid\mid\leq C_{1}\big{(}\mid\mid\varphi_{j}(0)\mid\mid+\mid\mid\psi_{j}(0)\mid\mid\big{)}. (46)
0L0φj(η~)[N(η1(L0,t),t,η~,0)N(η2(L0,t),t,η~,0)]𝑑ξ02φjδπ.\int_{0}^{L_{0}}\left|\left|\varphi_{j}^{\prime}(\tilde{\eta})[N(\eta_{1}(L_{0},t),t,\tilde{\eta},0)-N(\eta_{2}(L_{0},t),t,\tilde{\eta},0)]\right|\right|d\xi_{0}\leq\frac{2\mid\mid\varphi_{j}^{\prime}\mid\mid\delta}{\sqrt{\pi}}. (47)
0tψ˙(τ)[N(η1(L0,t),t,η1(L0,τ),τ)N(η2(L0,t),t,η2(L0,τ),τ)]𝑑τ4ψ˙δλπ.\int_{0}^{t}\left|\left|\dot{\psi}(\tau)[N(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)-N(\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)]\right|\right|d\tau\leq 4\|\dot{\psi}\|\frac{\delta\lambda}{\sqrt{\pi}}. (48)
0t|[θj1(τ)Gz(η1(L0,t),t,η1(L0,τ),τ)θj2(τ)Gz(η2(L0,t),t,η2(L0,τ),τ)]ηξ0|𝑑τ(Rλ+2)πρ1ρ2+δθj22λπ+λ\begin{array}[]{ll}\int_{0}^{t}\left|[\theta_{j1}(\tau)G_{z}(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)-\theta_{j2}(\tau)G_{z}(\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)]\frac{\partial\eta}{\partial\xi_{0}}\right|d\tau&\\ &\\ \leq\frac{(R\sqrt{\lambda}+2)}{\sqrt{\pi}}\|\rho_{1}-\rho_{2}\|+\delta\frac{\|\theta_{j2}\|}{2}\sqrt{\frac{\lambda}{\pi}}+\lambda\end{array} (49)
0tΦ˙j(τ)[N(η1(L0,t),t,0,τ)N(η2(L0,t),t,0,τ)]𝑑τΦ˙δ2λπ\int_{0}^{t}\mid\dot{\Phi}_{j}(\tau)[N(\eta_{1}(L_{0},t),t,0,\tau)-N(\eta_{2}(L_{0},t),t,0,\tau)]\mid d\tau\leq\|\dot{\Phi}\|\frac{\delta^{2}\lambda}{\sqrt{\pi}} (50)
0t[N(η1(L0,t),t,η1(L0,τ),τ)F(X1,S1)N(η2(L0,t),t,η2(L0,τ),τ)F(X2,S2)]𝑑τ2(A1+A2)\begin{array}[]{l}\int_{0}^{t}\mid[N(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)F(X_{1},S_{1})-N(\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)F(X_{2},S_{2})]\mid d\tau\\ \\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\leq 2(A_{1}+A_{2})\end{array} (51)

where

A1=(LS1S2+LX1X2)λπA_{1}=\left(L\|S_{1}-S_{2}\|+L\|X_{1}-X_{2}\|\right)\sqrt{\frac{\lambda}{\pi}}

and

A2=Mδ2t2π+3MδL02πtMδ2t+3MδL02πA_{2}=\frac{M\delta^{2}\sqrt{t}}{2\sqrt{\pi}}+\frac{3M\delta L_{0}}{2\sqrt{\pi t}}\leq\frac{M\delta^{2}\sqrt{t}+3M\delta L_{0}}{2\sqrt{\pi}}

(Here ρ1\rho_{1} and ρ2\rho_{2} correspond to η1\eta_{1} and η2\eta_{2}).

0t[N(η1(L0,t),t,0,τ)F(X10,S10)N(η2(L0,t),t,0,τ)F(X20,S20)]𝑑τ2KL0λe[X10X20+S10S20+1]\begin{array}[]{ll}\int_{0}^{t}\mid[N(\eta_{1}(L_{0},t),t,0,\tau)F(X_{1}^{0},S_{1}^{0})-N(\eta_{2}(L_{0},t),t,0,\tau)F(X_{2}^{0},S_{2}^{0})]\mid d\tau&\\ &\\ \leq\frac{2K}{L_{0}}\sqrt{\frac{\lambda}{e}}\Bigg{[}\mid X_{1}^{0}-X_{2}^{0}\mid+\mid S_{1}^{0}-S_{2}^{0}\mid+1\Bigg{]}\end{array} (52)

Proof: To prove the inequality 46, we can write

||Uj1(t)Uj2(t)||=2||φj(0)[N(L1(t),t,0,0)N(L2(t),t,0,0)]\left|\left|U_{j1}(t)-U_{j2}(t)\right|\right|=2\left|\left|\varphi_{j}(0)[N(L_{1}(t),t,0,0)-N(L_{2}(t),t,0,0)]\right.\right.
+ψj(0)[N(L1(t),t,L0,0)N(L2(t),t,L0,0)]||.\left.\left.+\psi_{j}(0)[N(L_{1}(t),t,L_{0},0)-N(L_{2}(t),t,L_{0},0)]\right|\right|.
=2||φj(0)[N(L1(t),t,0,0)N(L2(t),t,0,0)]+ψj(0)[N(L1(t),t,L0,0)=2\left|\left|\varphi_{j}(0)[N(L_{1}(t),t,0,0)-N(L_{2}(t),t,0,0)]+\psi_{j}(0)[N(L_{1}(t),t,L_{0},0)\right.\right.
N(L2(t),t,L0,0)]||=I1+I2\left.\left.-N(L_{2}(t),t,L_{0},0)]\right|\right|=I_{1}+I_{2}\\

By using the mean value theorem and inequalities 33, 34, and 44 we get inequality 46.
Let us denote by I the left hand side of the inequality 47. By taking η\eta* such that η2(L0,t)<η(L0,t)<η1(L0,t)\eta_{2}(L_{0},t)<\eta^{*}(L_{0},t)<\eta_{1}(L_{0},t) and by using the mean value theorem again, we have

I=20L0φj[η1(L0,t)η2(L0,t)]Nz(η(L0,t),t,η~,0)𝑑ξ02||φj||0L0||(η(L0,t)η2(L0,t))[(η(L0,t)η~)4tπtK(η(L0,t),t,η~,0)+(η(L0,t)+η~)4tπtK(η(L0,t),t,η~,0)]||dξ0\begin{array}[]{ll}I=2\int_{0}^{L_{0}}\left|\left|\varphi_{j}^{\prime}\left[\eta_{1}(L_{0},t)-\eta_{2}(L_{0},t)\right]N_{z}(\eta^{*}(L_{0},t),t,\tilde{\eta},0)\right|\right|d\xi_{0}\\ &\\ \leq 2\left|\left|\varphi_{j}^{\prime}\right|\right|\int_{0}^{L_{0}}\left|\left|\left(\eta(L_{0},t)-\eta_{2}(L_{0},t)\right)\left[\frac{\left(\eta^{*}(L_{0},t)-\tilde{\eta}\right)}{4t\sqrt{\pi t}}K(\eta^{*}(L_{0},t),t,\tilde{\eta},0)\right.\right.\right.\\ &\\ \left.\left.\left.+\frac{\left(\eta^{*}(L_{0},t)+\tilde{\eta}\right)}{4t\sqrt{\pi t}}K(-\eta^{*}(L_{0},t),t,\tilde{\eta},0)\right]\right|\right|d\xi_{0}\end{array}

Setting

I1=2φj0L0(η1(L0,t)η2(L0,t))[(η(L0,t)η~)4tπtK(η(L0,t),t,η~,0)]𝑑ξ0I_{1}=2\left|\left|\varphi_{j}^{\prime}\right|\right|\int_{0}^{L_{0}}\left|\left|\left(\eta_{1}(L_{0},t)-\eta_{2}(L_{0},t)\right)\left[\frac{\left(\eta^{*}(L_{0},t)-\tilde{\eta}\right)}{4t\sqrt{\pi t}}K(\eta^{*}(L_{0},t),t,\tilde{\eta},0)\right]\right|\right|d\xi_{0}

and

I2=2φj0L0(η1(L0,t)η2(L0,t))[(η(L0,t)+η~)4tπtK(η(L0,t),t,η~,0)]𝑑ξ0,I_{2}=2\left|\left|\varphi_{j}^{\prime}\right|\right|\int_{0}^{L_{0}}\left|\left|\left(\eta_{1}(L_{0},t)-\eta_{2}(L_{0},t)\right)\left[\frac{\left(\eta^{*}(L_{0},t)+\tilde{\eta}\right)}{4t\sqrt{\pi t}}K(-\eta^{*}(L_{0},t),t,\tilde{\eta},0)\right]\right|\right|d\xi_{0},

it can be checked that

I12φj0L0δt124π(η(L0,t)η~)K(η(L0,t),t,η~,0)𝑑ξ0I_{1}\leq 2\left|\left|\varphi_{j}^{\prime}\right|\right|\int_{0}^{L_{0}}\left|\left|\frac{\delta t^{-\frac{1}{2}}}{4\sqrt{\pi}}(\eta^{*}(L_{0},t)-\tilde{\eta})K(\eta^{*}(L_{0},t),t,\tilde{\eta},0)\right|\right|d\xi_{0}

Then, we make the following change of variable y=ηη~4ty=\frac{\eta^{*}-\tilde{\eta}}{4t} and the inequality becomes

I1φjδπt0L04t2yexp(y2)dyφj|δπt[exp(L024t)1]|φjδπtφjδπ.\begin{array}[]{l}I_{1}\leq\mid\mid\varphi_{j}^{\prime}\mid\mid\frac{\delta}{\sqrt{\pi t}}\int_{0}^{\frac{L_{0}}{\sqrt{4t}}}-2y\exp(-y^{2})dy\leq\mid\mid\varphi_{j}^{\prime}\mid\mid\left|\frac{\delta}{\sqrt{\pi t}}\left[\exp(-\frac{L_{0}^{2}}{4t})-1\right]\right|\\ \\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}\leq\frac{\mid\mid\varphi_{j}^{\prime}\mid\mid\delta}{\sqrt{\pi t}}\leq\frac{\mid\mid\varphi_{j}^{\prime}\mid\mid\delta}{\sqrt{\pi}}.\end{array}

The same reasoning can be applied to I2I_{2}. So, the original inequality holds .
Inequality 48 can be proved as follows again by using the MVT:

|0tψ˙(τ)[N(η1(L0,t),t,η1(L0,τ),τ)N(η2(L0,t),t,η2(L0,τ),τ)]𝑑τ|\left|\int_{0}^{t}\dot{\psi}(\tau)[N(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)-N(\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)]d\tau\right|\leq
ψ˙|0t[η1(L0,t)η2(L0,t)]Nz(η¯(L0,t),t,η¯(L0,τ),τ)𝑑τ|\mid\mid\dot{\psi}\mid\mid\left|\int_{0}^{t}[\eta_{1}(L_{0},t)-\eta_{2}(L_{0},t)]N_{z}(\bar{\eta}(L_{0},t),t,\bar{\eta}(L_{0},\tau),\tau)d\tau\right|
=ψ˙|0tδ(tτ)[η¯(L0,t)η¯(L0,τ)2(tτ)K(η¯(L0,t),t,η¯(L0,τ),τ)=\mid\mid\dot{\psi}\mid\mid\left|\int_{0}^{t}\delta(t-\tau)\left[\frac{\bar{\eta}(L_{0},t)-\bar{\eta}(L_{0},\tau)}{2(t-\tau)}K(\bar{\eta}(L_{0},t),t,\bar{\eta}(L_{0},\tau),\tau)\right.\right.
+η¯(L0,t)+η¯(L0,τ)2(tτ)K(η¯(L0,t),t,η¯(L0,τ),τ)]dτ|\left.\left.+\frac{\bar{\eta}(L_{0},t)+\bar{\eta}(L_{0},\tau)}{2(t-\tau)}K(-\bar{\eta}(L_{0},t),t,\bar{\eta}(L_{0},\tau),\tau)\right]d\tau\right|
ψ˙|0tδ2[(tτ)1/24π+3δL0(tτ)3/222π]𝑑τ|\leq\mid\mid\dot{\psi}\mid\mid\left|\int_{0}^{t}\delta^{2}\left[\frac{(t-\tau)^{1/2}}{4\sqrt{\pi}}+\frac{3\delta L_{0}(t-\tau)^{-3/2}}{2\sqrt{2\pi}}\right]d\tau\right|\\
δ2ψ˙λ3/26π+3δ2ψ˙L02π.\leq\frac{\delta^{2}\mid\mid\dot{\psi}\mid\mid\lambda^{3/2}}{6\sqrt{\pi}}+\frac{3\delta^{2}\mid\mid\dot{\psi}\mid\mid L_{0}}{\sqrt{2\pi}}.

Inequality 50 can also be proved in the same way. Denote the left hand side of 49, by J, we have:

J|0t[θj1(τ)θj2(τ))]Gz(η1(L0,t),t,η1(L0,τ),τ)ηξ0dτ|+|0tθj1(τ)[Gz(η1(L0,t),t,η1(L0,τ),τ)Gz(η2(L0,t),t,η2(L0,τ),τ)]𝑑τ|=J1+J2\displaystyle\begin{array}[]{l}J\leq\left|\int_{0}^{t}[\theta_{j1}(\tau)-\theta_{j2}(\tau))]G_{z}(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)\frac{\partial\eta}{\xi_{0}}d\tau\right|+\\ \\ \left|\int_{0}^{t}\theta_{j1}(\tau)[G_{z}(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)-G_{z}(\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)]d\tau\right|=J_{1}+J_{2}\\ \\ \end{array}

On one hand we have

J1Lθj1θj2|0tη1(L0,t)η2(L0,τ)2(tτ)K(η1(L0,t),t,η1(L0,τ),τ)+η1(L0,t)+η2(L0,τ)2(tτ)K(η1(L0,t),t,η1(L0,τ),τ)dτ|Lθj1θj2[|0t(tτ)124πdτ|+|3L04π0t(tτ)32dτ|]=Lθj1θj2(t4π+3L04πt),\begin{array}[]{l}J_{1}\leq L\mid\mid\theta_{j1}-\theta_{j2}\|\left|\int_{0}^{t}-\frac{\eta_{1}(L_{0},t)-\eta_{2}(L_{0},\tau)}{2(t-\tau)}K(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)\right.\\ \\ \left.+\frac{\eta_{1}(L_{0},t)+\eta_{2}(L_{0},\tau)}{2(t-\tau)}K(-\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)d\tau\right|\\ \leq L\mid\mid\theta_{j1}-\theta_{j2}\|\left[\left|\int_{0}^{t}\frac{(t-\tau)^{-\frac{1}{2}}}{4\sqrt{\pi}}d\tau\right|+\left|\frac{3L_{0}}{4\pi}\int_{0}^{t}(t-\tau)^{-\frac{3}{2}}d\tau\right|\right]=L\mid\mid\theta_{j1}-\theta_{j2}\|\left(\frac{\sqrt{t}}{4\sqrt{\pi}}+\frac{3L_{0}}{4\sqrt{\pi t}}\right),\end{array}

and on the other hand we have

J2|0tθj1(τ)[Gz(η1(L0,t),t,η1(L0,τ),τ)Gz(η2(L0,t),t,η2(L0,τ),τ)]ηξ0𝑑τ||0tθj1(τ))[η1(L0,t)η1(L0,τ)2(tτ)K(η1(L0,t),t,η1(L0,τ),τ)+η1(L0,t)+η1(L0,τ)2(tτ)K(η1(L0,t),t,η1(L0,τ),τ)]ηξ0dτ|+|0tθj1(τ))[η2(L0,t)η2(L0,τ)2(tτ)K(η2(L0,t),t,η2(L0,τ),τ)η2(L0,t)+η2(L0,τ)2(tτ)K(η2(L0,t),t,η2(L0,τ),τ)]ηξ0dτ|=J2+J2′′\displaystyle\begin{array}[]{l}J_{2}\leq\left|\int_{0}^{t}\theta_{j1}(\tau)[G_{z}(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)-G_{z}(\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)]\frac{\partial\eta}{\partial\xi_{0}}d\tau\right|\\ \\ \leq\left|\int_{0}^{t}\theta_{j1}(\tau))[-\frac{\eta_{1}(L_{0},t)-\eta_{1}(L_{0},\tau)}{2(t-\tau)}K(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)\right.\\ \\ \left.+\frac{\eta_{1}(L_{0},t)+\eta_{1}(L_{0},\tau)}{2(t-\tau)}K(-\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)]\frac{\partial\eta}{\partial\xi_{0}}d\tau\right|+\\ \\ \left|\int_{0}^{t}\theta_{j1}(\tau))[\frac{\eta_{2}(L_{0},t)-\eta_{2}(L_{0},\tau)}{2(t-\tau)}K(\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)-\right.\\ \\ \left.\frac{\eta_{2}(L_{0},t)+\eta_{2}(L_{0},\tau)}{2(t-\tau)}K(-\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)]\frac{\partial\eta}{\partial\xi_{0}}d\tau\right|=J_{2}^{\prime}+J_{2}^{\prime\prime}\\ \end{array}

By carrying out similar calculations as before and taking into account that J2′′J_{2}^{\prime\prime} is inferior to the product of its first term and θj1(τ),\theta_{j1}(\tau), we get:

J2M(3t2π+3L0π).\displaystyle J_{2}\leq M\left(\frac{3\sqrt{t}}{2\sqrt{\pi}}+\frac{3L_{0}}{\sqrt{\pi}}\right).

This completes the proof for the original inequality.
The left hand side of 51 is equal to:

|0t[F(X1,S2,η)F(X2,S2,η)]N(η1(L0,t),t,η1(L0,τ),τ)𝑑τ|+|0tF(X2,S2,η)[N(η1(L0,t),t,η1(L0,τ),τ)N(η2(L0,t),t,η2(L0,τ),τ)]|=B1+B2\begin{array}[]{l}\bigg{|}\int_{0}^{t}\left[F(X_{1},S_{2},\eta^{\prime})-F(X_{2},S_{2},\eta^{\prime})\right]N(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)d\tau\bigg{|}\\ \\ +\bigg{|}\int_{0}^{t}F(X_{2},S_{2},\eta^{\prime})\left[N(\eta_{1}(L_{0},t),t,\eta_{1}(L_{0},\tau),\tau)-N(\eta_{2}(L_{0},t),t,\eta_{2}(L_{0},\tau),\tau)\right]\bigg{|}\\ \\ ~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}=B_{1}+B_{2}\end{array}

It is easy to check that B1+B2A1+A2.B_{1}+B_{2}\leq A_{1}+A_{2}.
Inequality 52 can be also proved in the same way. This completes the proof.

Theorem 5.

The map 𝒯:CM,λCM,λ\mathcal{T}:~{}~{}C_{M,\lambda}\rightarrow C_{M,\lambda} is well defined and it is a contraction if it satisfies the following inequalities:

λ1;2Mλb;\begin{array}[]{c}\lambda\leq 1;\\ \\ 2M\lambda\leq b;\\ \end{array} (53)

K11;K_{1}\leq 1;\\ (54)

K2<1K_{2}<1 (55)

where

K1=M1+M2+M3+M4+M5+M6K_{1}=M_{1}+M_{2}+M_{3}+M_{4}+M_{5}+M_{6}

with

M1=φj(0)+ψj(0),M2=φj(η~),M3=ψ˙(t)πλ,M4=ψ˙(t)πλ,M_{1}=\|\varphi_{j}(0)\|+\|\psi_{j}(0)\|,M_{2}=\|\varphi_{j}^{\prime}(\tilde{\eta})\|,M_{3}=\frac{\|\dot{\psi}(t)\|}{\sqrt{\pi}}\sqrt{\lambda},M_{4}=\frac{\|\dot{\psi}(t)\|}{\sqrt{\pi}}\sqrt{\lambda},
M5=2Lπλ,M6=2KL0λeM_{5}=\frac{2L}{\sqrt{\pi}}\sqrt{\lambda},M_{6}=\frac{2K}{L_{0}}\sqrt{\frac{\lambda}{e}}

and K2K_{2} is the sum of right hand site of inequalities in lemma 4. Then, the original problem has a unique solution.

Proof. By lemma 3, we know that 𝒯uK1.\|\mathcal{T}u\|\leq K_{1}. Choosing M to satisfy M=1+2ϕ+ϕM=1+2\|\phi^{\prime}\|+\|\phi\| and if 55 holds, we have 𝒯u4M.\|\mathcal{T}u_{4}\|\leq M. To complete the proof, we define 𝒗:=(u1(η,t),u2(η,t),u3(η,t))T\boldsymbol{v}:=(u_{1}(\eta,t),u_{2}(\eta,t),u_{3}(\eta,t))^{T}, 𝒗0:=(ϕ(z0),z0,1)T,\boldsymbol{v}_{0}:=(\phi(z_{0}),z_{0},1)^{T}, and

(𝑿,𝑺)=(H(𝑿,𝑺),U¯(𝑿,𝑺),u(η,τ)) where U¯(𝑿,𝑺)=0z0R(𝑿,𝑺)ηζ𝑑ζ.\mathcal{B}(\boldsymbol{X},\boldsymbol{S})=\left(H(\boldsymbol{X},\boldsymbol{S}),\bar{U}(\boldsymbol{X},\boldsymbol{S}),u(\eta,\tau)\right)~{}~{}\text{ where }~{}~{}\bar{U}(\boldsymbol{X},\boldsymbol{S})=\int_{0}^{z_{0}}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial\zeta}d\zeta.

We can finally write:

𝒯v=u0+0t(𝑿,𝑺)𝑑τ.\mathcal{T}\textbf{v}=\textbf{u}_{0}+\int_{0}^{t}\mathcal{B}(\boldsymbol{X},\boldsymbol{S})d\tau.

Then by choosing two functions v1\textbf{v}_{1} and v2\textbf{v}_{2} corresponding to v we have the following

𝒯v1𝒯v20t(X1,S1)(X2,S2)𝑑τMλ(X1X2+S1S2)\displaystyle\|\mathcal{T}\textbf{v}_{1}-\mathcal{T}\textbf{v}_{2}\|\leq\int_{0}^{t}\|\mathcal{B}(X_{1},S_{1})-\mathcal{B}(X_{2},S_{2})\|d\tau\leq M\lambda(\|X_{1}-X_{2}\|+\|S_{1}-S_{2}\|)

By lemma 4, we have 𝒯u𝒯vMuv\|\mathcal{T}u-\mathcal{T}v\|\leq M\|u-v\| and by 55, and choosing M and λ\lambda such that Mλ<1,M\lambda<1, we have a contraction. Then the theorem is proved.

Remark 1.

When attachment or detachment is the most prevailing, then the function σ\sigma is different from zero. In either of these two cases the free boundary coincides with the characteristics-like lines. In case of detachment, the thickness equation is given by

L˙(t)=u(L(t),t)σd(L(t)),L(0)=L0.\dot{L}(t)=u(L(t),t)-\sigma_{d}(L(t)),~{}~{}L(0)=L_{0}. (56)

This case can be treated as in [14, 20]. When attachment is the most prevailing as at the earlier moments of the biofilm life, the free boundary is given by

L˙(t)=u(L(t),t)+σa(t),L(0)=L0.\dot{L}(t)=u(L(t),t)+\sigma_{a}(t),~{}~{}L(0)=L_{0}. (57)

This case has been widely studied in [21] especially for biomass equations and the steady-state substrates equations.
In both these cases, the free boundary and the characteristics-like lines do not coincide anymore.
All the equations presented here can be treated in the same way for each of these cases.

We will briefly prove the global existence.

4 Global existence

We want now extend our solution to t0.t\geq 0. We consider T1T_{1} such that T1>λ.T_{1}>\lambda.
We introduce the following notations L(tδ):=L1,L(t-\delta):=L_{1}, η(L(t1δ),t)=η(L1,t)\eta(L(t_{1}-\delta),t)=\eta(L_{1},t) to get

θ(t)=2S(0,t1δ)N(L(t),t,0,t1δ)+2S(L(t),t1δ)N(L(t),t,L(t1δ),t1δ)+H1+H2+H3+H4\begin{array}[]{l}\theta(t)=2S(0,t_{1}-\delta)N(L(t),t,0,t_{1}-\delta)+2S(L(t),t_{1}-\delta)N(L(t),t,L(t_{1}-\delta),t_{1}-\delta)\\ \\ +H_{1}+H_{2}+H_{3}+H_{4}\end{array}

where

H1=20L1Sjξ(ξ,t1δ)N(L(t),t,ξ,t1δ)ηξ0𝑑ξ0,H_{1}=2\int_{0}^{L_{1}}S_{j\xi}(\xi,t_{1}-\delta)N(L(t),t,\xi,t_{1}-\delta)\frac{\partial\eta}{\partial\xi_{0}}d\xi_{0},
H2=2t1δtψ˙(τ)N(η(L,t),t,η(L(τ),τ))𝑑τH_{2}=2\int_{t_{1}-\delta}^{t}\dot{\psi}(\tau)N(\eta(L,t),t,\eta(L(\tau),\tau))d\tau
H3=2t1δtθ(τ)Gz(η(L,t),t,η(τ,t),τ)𝑑τH_{3}=2\int_{t_{1}-\delta}^{t}\theta(\tau)G_{z}(\eta(L,t),t,\eta(\tau,t),\tau)d\tau~{}~{}~{}~{}
H4=2tτtN(η(L,t),t,η(L(τ),t),τ)F(𝑿,𝑺)ηξ0𝑑ξ0𝑑τ.H_{4}=2\int_{t-\tau}^{t}N(\eta(L,t),t,\eta(L(\tau),t),\tau)F(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial\xi_{0}}d\xi_{0}d\tau.

We also have

u(z,t)=0η(tδ)R(𝑿,𝑺)ηζ0𝑑ζ0,u(z,t)=\int_{0}^{\eta(t-\delta)}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial\zeta_{0}}d\zeta_{0}, (58)
X(z,t)=X(z,t1δ)+t1δtH(𝑿,𝑺)𝑑τ,X(z,t)=X(z,t_{1}-\delta)+\int_{t_{1}-\delta}^{t}H(\boldsymbol{X},\boldsymbol{S})d\tau, (59)
η(z0,t)=η(z0,t1δ)+t1δt0η(t1δ)R(𝑿,𝑺)ηζ0𝑑ζ0𝑑τ,\eta(z_{0},t)=\eta(z_{0},t_{1}-\delta)+\int_{t_{1}-\delta}^{t}\int_{0}^{\eta(t_{1}-\delta)}R(\boldsymbol{X},\boldsymbol{S})\frac{\partial\eta}{\partial\zeta_{0}}d\zeta_{0}d\tau, (60)

and

ηz0=1+t1δtu(η,τ)ηz0𝑑τ\frac{\partial\eta}{\partial z_{0}}=1+\int_{t_{1}-\delta}^{t}u(\eta,\tau)\frac{\partial\eta}{\partial z_{0}}d\tau (61)

Now, we check that these integrals are still bounded. First, we notice that V1V_{1} can be written as V1=V1(1)+V1(2)V_{1}=V_{1}^{(1)}+V_{1}^{(2)}

V1(1)=2S(0,t1δ)N(L(t),t,0,t1δ) and V_{1}^{(1)}=2S(0,t_{1}-\delta)N(L(t),t,0,t_{1}-\delta)\text{ and }
V1(2)=2S(L(t),t1δ)N(L(t),t,L(t1δ),t1δ).V_{1}^{(2)}=2S(L(t),t_{1}-\delta)N(L(t),t,L(t_{1}-\delta),t_{1}-\delta).

Then we notice that

|V1(1)|2S(0,t1δ)|V_{1}^{(1)}|\leq 2\left|\left|S(0,t_{1}-\delta)\right|\right|

and

|V1(2)|2S(L(t),t1δ)|K(η(L,t),t,η(L1,t),t1δ)+K(η(L,t),t,η(L1,t),t1δ)|2S(L(t),t1δ)1πδ;\begin{array}[]{l}\left|V_{1}^{(2)}\right|\leq 2\left|\left|S(L(t),t_{1}-\delta)\right|\right|\left|K(\eta(L,t),t,\eta(L_{1},t),t_{1}-\delta)+K(-\eta(L,t),t,\eta(L_{1},t),t_{1}-\delta)\right|\\ \leq 2\left|\left|S(L(t),t_{1}-\delta)\right|\right|\frac{1}{\sqrt{\pi\delta}};\end{array}

Thus,

|V(t)|2S(0,t1δ)+2S(L(t),t1δ)1πδ.|V(t)|\leq 2\left|\left|S(0,t_{1}-\delta)\right|\right|+2\left|\left|S(L(t),t_{1}-\delta)\right|\right|\frac{1}{\sqrt{\pi\delta}}. (62)

On the other hand we have:

|H1|2|0L1Sjξ(ξ,t1δ)N(L(t),t,ξ,t1δ)ηξ0𝑑ξ0|Sz(z,t1δ)\left|H_{1}\right|\leq 2\left|\int_{0}^{L_{1}}S_{j\xi}(\xi,t_{1}-\delta)N(L(t),t,\xi,t_{1}-\delta)\frac{\partial\eta}{\partial\xi_{0}}d\xi_{0}\right|\leq\left|\left|S_{z}(z,t_{1}-\delta)\right|\right|
|H2|ψ˙t1δtN(η(L,t),t,η(L(τ),τ))𝑑τψ˙πδ|H_{2}|\leq\|\dot{\psi}\|\int_{t_{1}-\delta}^{t}N(\eta(L,t),t,\eta(L(\tau),\tau))d\tau\leq\frac{\|\dot{\psi}\|}{\sqrt{\pi}}\sqrt{\delta}
|H3|2θ|t1δtGz(η(L,t),t,η(τ,t),τ)𝑑τ|=\displaystyle|H_{3}|\leq 2\|\theta\|\left|\int_{t_{1}-\delta}^{t}G_{z}(\eta(L,t),t,\eta(\tau,t),\tau)d\tau\right|=
2θ|t1δt((ηη~)2(tτ)K(η,t,η~,τ)+(η+η~)2(tτ)K(η,t,η~,τ))𝑑τ|\displaystyle 2\|\theta\|\left|\int_{t_{1}-\delta}^{t}\left(\frac{(\eta-\tilde{\eta})}{2(t-\tau)}K(\eta,t,\tilde{\eta},\tau)+\frac{(\eta+\tilde{\eta})}{2(t-\tau)}K(-\eta,t,\tilde{\eta},\tau)\right)d\tau\right|
θ|t1δtRdτ4π(tτ)|+θ|t1δt3L(t1δ)dτ4π(tτ)|Rθ2πδ+3θ2πδ\displaystyle\leq\|\theta\|\left|\int_{t_{1}-\delta}^{t}\frac{Rd\tau}{\sqrt{4\pi(t-\tau)}}\right|+\|\theta\|\left|\int_{t_{1}-\delta}^{t}\frac{3L(t_{1}-\delta)d\tau}{\sqrt{4\pi(t-\tau)}}\right|\leq\frac{R\|\theta\|}{2\sqrt{\pi}}\sqrt{\delta}+\frac{3\|\theta\|}{2\sqrt{\pi}}\sqrt{\delta}
|H4||2t1τtN(η(L,t),t,η(L(τ),t),τ)F(𝑿,𝑺)𝑑τ|2L|t1δtdτ4π(tτ))|+2L|t1δtdτ4π(tτ))|=2Lδπ.\begin{array}[]{l}|H_{4}|\leq\left|2\int_{t_{1}-\tau}^{t}N(\eta(L,t),t,\eta(L(\tau),t),\tau)F(\boldsymbol{X},\boldsymbol{S})d\tau\right|\leq 2L\left|\int_{t_{1}-\delta}^{t}\frac{d\tau}{\sqrt{4\pi(t-\tau))}}\right|\\ \\ +2L\left|\int_{t_{1}-\delta}^{t}\frac{d\tau}{\sqrt{4\pi(t-\tau))}}\right|=\frac{2L\sqrt{\delta}}{\sqrt{\pi}}.\end{array}

We conclude that

|θ|2S(0,t1δ)+2S(L(t),t1δ)1πδ+Sz(z,t1δ)+ψ˙πδ\displaystyle|\theta|\leq 2\left|\left|S(0,t_{1}-\delta)\right|\right|+2\left|\left|S(L(t),t_{1}-\delta)\right|\right|\frac{1}{\sqrt{\pi\delta}}+\left|\left|S_{z}(z,t_{1}-\delta)\right|\right|+\frac{\|\dot{\psi}\|}{\sqrt{\pi}}\sqrt{\delta}
+Rθ2πδ+3θ2πδ+2Lδπ.\displaystyle+\frac{R\|\theta\|}{2\sqrt{\pi}}\sqrt{\delta}+\frac{3\|\theta\|}{2\sqrt{\pi}}\sqrt{\delta}+\frac{2L\sqrt{\delta}}{\sqrt{\pi}}.

In the same way, we can show that each of uu, XX, η\eta, and ηz\frac{\partial\eta}{\partial z} in equations 5961 is bounded by a constant BiB_{i} (i=1,,4i=1,...,4) independent of tt. Thus, the solution exists up to T1T_{1} for T1TT_{1}-T small enough. We can iterate this procedure for a sequence of TiT_{i} (for any integer i>1i>1) and show that TnT_{n}\rightarrow\infty as nn\rightarrow\infty as it was done in [22].

5 Robin-Neumann boundaries

The formulation of the second principle is influenced by these analyses. In [23, 24], it is hypothesized that a specific measurement away from the substratum, defined as H(t)=L(t)+hH(t)=L(t)+h (where h stands as a specified positive constant), the concentration of the substrate Sj(H(t),t)S_{j}(H(t),t) to that within the main fluid, denoted by Γ(t).\Gamma(t). This particular substrate, once dissolved, moves from the primary liquid into the biofilm between 0zL(t)0\leq z\leq L(t) and is then utilized as per equation 7. It’s assumed that no biochemical processes take place on L(t)zH(t),L(t)\leq z\leq H(t), which leads to consider homogeneous parabolic equations for Sj(z,t)S_{j}(z,t). The resolution at steady-state results in

hDjDjSjz(L(t),t)+kSj(Lt),t)=ψj(t),h\frac{D_{j}}{D^{*}_{j}}\frac{\partial S_{j}}{\partial z}(L(t),t)+kS_{j}(Lt),t)=\psi_{j}(t), (63)

or equivalently

Sjz(L(t),t)=α1ψj(t)α2Sj(L(t),t)\frac{\partial S_{j}}{\partial z}(L(t),t)=\alpha_{1}\psi_{j}(t)-\alpha_{2}S_{j}(L(t),t) (64)

where DjD_{j}^{*} denotes the diffusion rate of substrate j in the bulk liquid. In equation 64 α1=DjhDj\alpha_{1}=\frac{D_{j}^{*}}{hD_{j}} and α2=kDjhDj.\alpha_{2}=\frac{kD^{*}_{j}}{hD_{j}}. For h=0h=0 we get the boundary conditions in the previous sections and for k=0k=0 we have a Neumann boundary condition which can be treated in the same way as in this section.
By integrating the Green’s identity as before we get

Sj(z,t)=0L0Sj(ξ,0)(ξ)G(z,t;ξ,0)𝑑ξ+0tSj(0,τ)G(z,t;0,τ)𝑑τ\displaystyle S_{j}(z,t)=\int_{0}^{L_{0}}S_{j}(\xi,0)(\xi)G(z,t;\xi,0)d\xi+\int_{0}^{t}S_{j}(0,\tau)G(z,t;0,\tau)d\tau
0tSj(L(τ),τ)Gξ(z,t;L(τ),τ)𝑑τ+0tSjz(L(τ),τ)G(z,t;L(τ),τ)𝑑τ\displaystyle-\int_{0}^{t}S_{j}(L(\tau),\tau)G_{\xi}(z,t;L(\tau),\tau)d\tau+\int_{0}^{t}\frac{\partial S_{j}}{\partial z}(L(\tau),\tau)G(z,t;L(\tau),\tau)d\tau
+0t0L(t)G(z,t;ξ,τ)Fj(𝑿,𝑺)𝑑ξ𝑑τ,\displaystyle+\int_{0}^{t}\int_{0}^{L(t)}G(z,t;\xi,\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\xi d\tau,

Introducing the initial and boundary conditions as before, we get

Sj(z,t)=0L0φj(ξ)G(z,t;ξ,0)𝑑ξ+0tΦj(τ)Gξ(z,t;0,τ)𝑑τ0tρj(τ)Gξ(z,t;L(τ),τ)𝑑τ+0t[α1ψj(τ)α2ρj(τ)]G(z,t;L(τ),τ)𝑑τ+0t0L(t)G(z,t;ξ,τ)Fj(𝑿,𝑺)𝑑ξ𝑑τ,\begin{array}[]{l}S_{j}(z,t)=\int_{0}^{L_{0}}\varphi_{j}(\xi)G(z,t;\xi,0)d\xi+\int_{0}^{t}\Phi_{j}(\tau)G_{\xi}(z,t;0,\tau)d\tau\\ \\ -\int_{0}^{t}\rho_{j}(\tau)G_{\xi}(z,t;L(\tau),\tau)d\tau+\int_{0}^{t}[\alpha_{1}\psi_{j}(\tau)-\alpha_{2}\rho_{j}(\tau)]G(z,t;L(\tau),\tau)d\tau\\ \\ +\int_{0}^{t}\int_{0}^{L(t)}G(z,t;\xi,\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\xi d\tau,\end{array} (65)

where

ρj(t)=Sj(L(t),t) and Φj(t)=Sj(0,t).\rho_{j}(t)=S_{j}(L(t),t)\text{ and }\Phi_{j}(t)=S_{j}(0,t). (66)

The function Φj\Phi_{j} is still given by equation 15 as follows

Φj(t)=0tρj(τ)Gξ(0,t;L(τ),τ)𝑑τ\Phi_{j}(t)=-\int_{0}^{t}\rho_{j}(\tau)G_{\xi}(0,t;L(\tau),\tau)d\tau (67)

and the function ρj\rho_{j} is determined by letting zL(t)z\rightarrow L(t) so that we can write

ρj(t)=0L0φj(ξ)G(L(t),t;ξ,0)𝑑ξ+0tΦj(τ)Gξ(L(t),t;0,τ)𝑑τ0tρj(τ)Gξ(L(t),t;L(τ),τ)𝑑τ+0t[α1ψj(τ)α2ρj(τ)]G(L(t),t;L(τ),τ)𝑑τ+0t0L(t)G(L(t),t;ξ,τ)Fj(𝑿,𝑺)𝑑ξ𝑑τ,\begin{array}[]{l }\rho_{j}(t)=\int_{0}^{L_{0}}\varphi_{j}(\xi)G(L(t),t;\xi,0)d\xi+\int_{0}^{t}\Phi_{j}(\tau)G_{\xi}(L(t),t;0,\tau)d\tau\\ \\ -\int_{0}^{t}\rho_{j}(\tau)G_{\xi}(L(t),t;L(\tau),\tau)d\tau+\int_{0}^{t}[\alpha_{1}\psi_{j}(\tau)-\alpha_{2}\rho_{j}(\tau)]G(L(t),t;L(\tau),\tau)d\tau\\ \\ +\int_{0}^{t}\int_{0}^{L(t)}G(L(t),t;\xi,\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})d\xi d\tau,\end{array} (68)

Hence, the substrates system 8a-8c together with Robin boundary condition 63 is equivalent to the simultaneous system of integral equations 65, 67 and 68 whenever LL satisfy the equation 19. Converting these integral equations into characteristics we get the following system

Sj(η,t)=0L0φj(η~)G(η,t;η~,0)η~ξ0𝑑ξ0+0tΦj(τ)Gξ(η,t;0,τ)𝑑τ0tρj(τ)Gξ(η,t;η(L0,τ),τ)η~ξ0𝑑τ+0t[α1ψj(τ)α2ρj(τ)]G(η,t;η(L0,τ),τ)𝑑τ+0t0L0G(η,t;η~,τ)Fj(𝑿,𝑺)η~ξ0𝑑ξ0𝑑τ\begin{array}[]{l}S_{j}(\eta,t)=\int_{0}^{L_{0}}\varphi_{j}(\tilde{\eta})G(\eta,t;\tilde{\eta},0)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\xi_{0}+\int_{0}^{t}\Phi_{j}(\tau)G_{\xi}(\eta,t;0,\tau)d\tau\\ \\ -\int_{0}^{t}\rho_{j}(\tau)G_{\xi}(\eta,t;\eta(L_{0},\tau),\tau)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\tau+\int_{0}^{t}[\alpha_{1}\psi_{j}(\tau)-\alpha_{2}\rho_{j}(\tau)]G(\eta,t;\eta(L_{0},\tau),\tau)d\tau\\ \\ +\int_{0}^{t}\int_{0}^{L_{0}}G(\eta,t;\tilde{\eta},\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\xi_{0}d\tau\end{array} (69)
Φj(t)=0tρj(τ)Gξ(η,t;η(L0,τ),τ)η~ξ0𝑑τ\Phi_{j}(t)=-\int_{0}^{t}\rho_{j}(\tau)G_{\xi}(\eta,t;\eta(L_{0},\tau),\tau)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\tau (70)
ρj(t)=0L0φj(η~)G(η(L0,t),t;η~,0)η~ξ0𝑑ξ0+0tΦj(τ)Gξ(η(L0,t),t;0,τ)𝑑τ+0t[α1ψj(τ)α2ρj(τ)]Gη(L0,t),t;η(L0,τ),τ)dτ0tρj(τ)Gξ(η(L0,t),t;η(L0,τ),τ)η~ξ0𝑑τ+0t0L0G(η(L0,t),t;η~,τ)Fj(𝑿,𝑺)η~ξ0𝑑ξ0𝑑τ\begin{array}[]{l }\rho_{j}(t)=\int_{0}^{L_{0}}\varphi_{j}(\tilde{\eta})G(\eta(L_{0},t),t;\tilde{\eta},0)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\xi_{0}+\int_{0}^{t}\Phi_{j}(\tau)G_{\xi}(\eta(L_{0},t),t;0,\tau)d\tau\\ \\ +\int_{0}^{t}[\alpha_{1}\psi_{j}(\tau)-\alpha_{2}\rho_{j}(\tau)]G\eta(L_{0},t),t;\eta(L_{0},\tau),\tau)d\tau\\ \\ -\int_{0}^{t}\rho_{j}(\tau)G_{\xi}(\eta(L_{0},t),t;\eta(L_{0},\tau),\tau)\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\tau\\ \\ +\int_{0}^{t}\int_{0}^{L_{0}}G(\eta(L_{0},t),t;\tilde{\eta},\tau)F_{j}(\boldsymbol{X},\boldsymbol{S})\frac{\partial\tilde{\eta}}{\partial\xi_{0}}d\xi_{0}d\tau\end{array} (71)

As for lemmas 3 and 4 we can establish estimates that guarantee the existence and uniqueness theorem 5.

6 Parabolic equation with variable diffusivity

We have considered so far equations with constant coefficients. However, experiments show that some phenomena happening inside a biofilm cannot be modeled by means of equations with constant coefficients. One of leading processes occurring in a biofilm is precipitation. We recently proposed a model on this topic (see [25]) where the following diffusion equation was introduced:

St=z(D(z,t)Sz)+F^(𝑿,𝑺)\frac{\partial S}{\partial t}=\frac{\partial}{\partial z}\left(D(z,t)\frac{\partial S}{\partial z}\right)+\hat{F}(\boldsymbol{X},\boldsymbol{S}) (72)

where

D(z,t)=D0exp((1p(x))12).D(z,t)=D_{0}\exp(-(1-p(x))^{\frac{1}{2}}). (73)

In the last equation, D0D_{0} is the diffusivity in water, and pp is the porosity of the biofilm. The formulation was made following the assumption that precipitates ”clog” the pores of the biofilm during their accumulation.
Since DD is a differentiable function, equation 72 can be rewritten as follows:

a(z,t)2Sz2+b(z,t)SzSt=F(𝑿,𝑺)a(z,t)\frac{\partial^{2}S}{\partial z^{2}}+b(z,t)\frac{\partial S}{\partial z}-\frac{\partial S}{\partial t}=F(\boldsymbol{X},\boldsymbol{S}) (74)

where a(z,t)=D(x,t)a(z,t)=D(x,t), b(x,t)=Dzb(x,t)=\frac{\partial D}{\partial z}, and F(𝑿,𝑺)=F^(𝑿,𝑺).F(\boldsymbol{X},\boldsymbol{S})=-\hat{F}(\boldsymbol{X},\boldsymbol{S}). The boundary and initial conditions considered here are the same as those considered in with constant coefficients. Note that this variable diffusivity does not affect the growth equation for biomass.
We are first interested in the homogeneous equation associated to 74 in the spirit of [26, 27, 28]:

Pa(z,t)2Sz2+b(z,t)SzSt=0.P\equiv a(z,t)\frac{\partial^{2}S}{\partial z^{2}}+b(z,t)\frac{\partial S}{\partial z}-\frac{\partial S}{\partial t}=0. (75)

From [15] this equation has the following fundamental solution :

Γ(z,t;ξ,τ)=Z(z,t;ξ,τ)+0t0LZ(z,t;σ,λ)Φ(σ,λ;ξ,τ)𝑑σ𝑑λ\Gamma(z,t;\xi,\tau)=Z(z,t;\xi,\tau)+\int_{0}^{t}\int_{0}^{L}Z(z,t;\sigma,\lambda)\Phi(\sigma,\lambda;\xi,\tau)d\sigma d\lambda (76)

where

Z(z,t;ξ,τ)=14πa(ξ,τ)(tτ)exp((xξ)24a(ξ,τ)(tτ))Z(z,t;\xi,\tau)=\frac{1}{\sqrt{4\pi a(\xi,\tau)(t-\tau)}}\exp\left(-\frac{(x-\xi)^{2}}{4a(\xi,\tau)(t-\tau)}\right) (77)

and

Φ(x,t;ξ,τ)=P(Z(x,t;ξ,τ))+0t0LP(Z(z,t;σ,λ))Φ(σ,λ,ξ,τ)𝑑σ𝑑λ\Phi(x,t;\xi,\tau)=P\left(Z(x,t;\xi,\tau)\right)+\int_{0}^{t}\int_{0}^{L}P\left(Z(z,t;\sigma,\lambda)\right)\Phi(\sigma,\lambda,\xi,\tau)d\sigma d\lambda (78)

We have the following properties of Z(z,t;ξ,τ)Z(z,t;\xi,\tau) and Γ(z,t;ξ,τ)\Gamma(z,t;\xi,\tau) adapted from ([28]):

|Γ(z,t;ξ,τ)|<M1(tτ)12exp(μ1(xξ)2tτ)\left|\Gamma(z,t;\xi,\tau)\right|<M_{1}\left(t-\tau\right)^{-\frac{1}{2}}\exp\left(-\frac{\mu_{1}(x-\xi)^{2}}{t-\tau}\right) (79)
|Γz(z,t;ξ,τ)|<M2(tτ)1exp(μ2(xξ)2tτ)\left|\Gamma_{z}(z,t;\xi,\tau)\right|<M_{2}\left(t-\tau\right)^{-1}\exp\left(-\frac{\mu_{2}(x-\xi)^{2}}{t-\tau}\right) (80)
|Γt(z,t;ξ,τ)|<M3(tτ)32exp(μ3(xξ)2tτ)\left|\Gamma_{t}(z,t;\xi,\tau)\right|<M_{3}\left(t-\tau\right)^{-\frac{3}{2}}\exp\left(-\frac{\mu_{3}(x-\xi)^{2}}{t-\tau}\right) (81)

Instead of following the method proposed in [29, 28]; we will construct a new Green’s function as we did for the constant coefficients equation. Define

H(z,t;ξ,τ)=Γ(z,t;ξ,τ)Γ(z,t;ξ,τ).H(z,t;\xi,\tau)=\Gamma(z,t;\xi,\tau)-\Gamma(-z,t;\xi,\tau). (82)

By using the Green’s relation as equality as before on Ω=[0,L]×[0,T],\Omega=[0,L]\times[0,T], and including the boundary and initial conditions and using characteristics coordinates, we end up with the following integral equation:

Sj(z,t)=0L0H(z,t;ξ,0)φj(ξ)𝑑ξ+0tψj(τ)Hξ(x,t;L(τ),τ)𝑑τ\displaystyle S_{j}(z,t)=-\int_{0}^{L_{0}}H(z,t;\xi,0)\varphi_{j}(\xi)d\xi+\int_{0}^{t}\psi_{j}(\tau)H_{\xi}(x,t;L(\tau),\tau)d\tau
0tH(z,t;L(τ),τ)Sjξ(L(τ),τ)𝑑τ+0t0LH(z,t,ξ,τ)F(𝑿,𝑺)𝑑ξ𝑑τ.\displaystyle-\int_{0}^{t}H(z,t;L(\tau),\tau)S_{j\xi}(L(\tau),\tau)d\tau+\int_{0}^{t}\int_{0}^{L}H(z,t,\xi,\tau)F(\boldsymbol{X},\boldsymbol{S})d\xi d\tau.

In addition, we can show that Γ(z,t;ξ,τ)\Gamma(z,t;\xi,\tau) satisfies 16. Then, we define V(t)=Sz(L(t),t)V(t)=S_{z}(L(t),t) as before.

V(t)=0L0Hz(L(t),t;ξ,0)φj(ξ)𝑑ξ+0tψj(τ)Hξz(L(t),t;L(τ),τ)𝑑τ0tHz(L(t),t;L(τ),τ)V(τ)𝑑τ+0t0LH(L(t),t,ξ,τ)F(𝑿,𝑺)𝑑ξ𝑑τ.\begin{array}[]{rl}V(t)=&-\int_{0}^{L_{0}}H_{z}(L(t),t;\xi,0)\varphi_{j}(\xi)d\xi+\int_{0}^{t}\psi_{j}(\tau)H_{\xi z}(L(t),t;L(\tau),\tau)d\tau\\ &\\ &-\int_{0}^{t}H_{z}(L(t),t;L(\tau),\tau)V(\tau)d\tau+\int_{0}^{t}\int_{0}^{L}H(L(t),t,\xi,\tau)F(\boldsymbol{X},\boldsymbol{S})d\xi d\tau.\end{array} (83)

By using inequalities 7981 we can prove propositions similar to lemmas 3 and 4 but we have to be careful because we don’t need a Neumann function such as the one defined in 11. In this way, we can prove the existence and uniqueness of the solution.

7 Conclusion and future works

We have presented in this paper a qualitative analysis of the biofilm model. The model takes into account the growth of the biofilm and the dynamics of the substrates in the biofilm. The work mainly showed the existence and uniqueness of solution to this free boundary problem. However, our system describing the growth of biofilm is not strictly hyperbolic due to a major simplification made on the model formulation. This simplification consists in assuming that the growth velocity is the same for all the biomass. But this is not the case in reality. Considering different growth velocities leads to a strictly hyperbolic system and this induces major changes in the study of the problem. Such considerations have never been taken into account and this comprises a long list of open problems.
In addition, we made strong regularity assumptions on the initial data and the boundary conditions. In some cases we can have functions that are not smooth enough and the existence and uniqueness criteria studied here are no longer valid.
In the case of variable diffusivity, we may have a coefficient which is not derivable, then the transition we made is not valid. Additionally, we considered the Dirichlet boundary conditions to be ”given” but in general, this not the case. The boundary conditions are derived following a mass balance low following the inlet and outlet from the bioreactor. For this reason, we need a supplementary system of ODE which has additional complications in the study of the system. Finally, in some works, the biofilms growth is modeled by a system of advection-diffusion-reaction equations. Combined with the diffusion equation, we have a completely new system whose solution may or not blow up.

Acknowledgements

This research has been supported by the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement N0N^{0} 861088.
Maria Rosaria Mattei acknowledges support from the project PRIN 2022 titled Spectral reectance signature of colored subaerial biolms as an indicator of stone heritage susceptibility to biodeterioration, project code: 2022KTBX3M, CUP: E53D23010850006.
Luigi Frunzo acknowledges support from the project PRIN 2022 titled
MOMENTA-Modelling complex biOlogical systeMs for biofuEl productioN and sTorAge: mathematics meets green industry, project code: 202248TY47, CUP: E53D23005430006.
This paper has been performed under the auspices of the G.N.F.M. of I.N.d.A.M.

References

  • [1] B. D’Acunto, L. Frunzo, V. Luongo, M. Mattei, Mathematical modeling of biofilms, in: Introduction to Biofilm Engineering, ACS Publications, 2019, pp. 245–273.
  • [2] O. Wanner, W. Gujer, A multispecies biofilm model, Biotechnology and bioengineering 28 (3) (1986) 314–328.
  • [3] M. Plattes, H. M. F. Lahore, Perspectives on the monod model in biological wastewater treatment, Journal of Chemical Technology & Biotechnology 98 (4) (2023) 833–837.
  • [4] B. D’Acunto, L. Frunzo, M. Mattei, On a free boundary problem for biosorption in biofilms, Nonlinear Analysis: Real World Applications 39 (2018) 120–141.
  • [5] H. Feldman, X. Flores-Alsina, P. Ramin, K. Kjellberg, U. Jeppsson, D. Batstone, K. Gernaey, Assessing the effects of intra-granule precipitation in a full-scale industrial anaerobic digester, Water Science and Technology 79 (7) (2019) 1327–1337.
  • [6] B. D’Acunto, L. Frunzo, I. Klapper, M. Mattei, Modeling multispecies biofilms including new bacterial species invasion, Mathematical biosciences 259 (2015) 20–26.
  • [7] A. Mašić, H. J. Eberl, A modeling and simulation study of the role of suspended microbial populations in nitrification in a biofilm reactor, Bulletin of mathematical biology 76 (2014) 27–58.
  • [8] A. Mašić, H. J. Eberl, A chemostat model with wall attachment: the effect of biofilm detachment rates on predicted reactor performance, in: Mathematical and Computational Approaches in Advancing Modern Science and Engineering, Springer, 2016, pp. 267–276.
  • [9] C. Picioreanu, J. Xavier, M. C. van Loosdrecht, Advances in mathematical modeling of biofilm structure, Biofilms 1 (4) (2004) 337–349.
  • [10] F. Clarelli, C. Di Russo, R. Natalini, M. Ribot, Mathematical models for biofilms on the surface of monuments, in: Applied and industrial mathematics in Italy III, World Scientific, 2010, pp. 220–231.
  • [11] B. Szomolay, Analysis of a moving boundary value problem arising in biofilm modelling, Mathematical methods in the applied sciences 31 (15) (2008) 1835–1859.
  • [12] A. Mašić, H. J. Eberl, Persistence in a single species cstr model with suspended flocs and wall attached biofilms, Bulletin of mathematical biology 74 (2012) 1001–1026.
  • [13] F. Russo, M. Mattei, A. Tenore, B. D’Acunto, V. Luongo, L. Frunzo, Analysis of a spherical free boundary problem modelling granular biofilms, arXiv preprint arXiv:2301.12263 (2023).
  • [14] B. D’Acunto, L. Frunzo, M. Mattei, Qualitative analysis of the moving boundary problem for a biofilm reactor model, Journal of Mathematical Analysis and Applications 438 (1) (2016) 474–491.
  • [15] A. Friedman, Partial differential equations of parabolic type, Courier Dover Publications, 2008.
  • [16] J. R. Cannon, The one-dimensional heat equation, no. 23, Cambridge University Press, 1984.
  • [17] D. V. Widder, The heat equation, Vol. 67, Academic Press, 1976.
  • [18] A. C. Briozzo, D. A. Tarzia, A one-phase stefan problem for a non-classical heat equation with a heat flux condition on the fixed face, Applied Mathematics and Computation 182 (1) (2006) 809–819.
  • [19] A. C. Briozzo, D. A. Tarzia, A stefan problem for a non-classical heat equation with a convective condition, Applied Mathematics and Computation 217 (8) (2010) 4051–4060.
  • [20] B. D’Acunto, L. Frunzo, Qualitative analysis and simulations of a free boundary problem for multispecies biofilm models, Mathematical and computer modelling 53 (9-10) (2011) 1596–1606.
  • [21] B. D’Acunto, G. Esposito, L. Frunzo, M. R. Mattei, F. Pirozzi, Analysis and simulations of the initial phase in multispecies biofilm formation, Communications in Applied and Industrial Mathematics 4 (2013).
  • [22] S. B. Cui, Analysis of a free boundary problem modeling tumor growth, Acta Mathematica Sinica 21 (5) (2005) 1071–1082.
  • [23] B. D’ACUNTO, L. Frunzo, V. Luongo, M. R. Mattei, Invasion moving boundary problem for a biofilm reactor model, European Journal of Applied Mathematics 29 (6) (2018) 1079–1109.
  • [24] I. Klapper, J. Dockery, Finger formation in biofilm layers, SIAM Journal on Applied Mathematics 62 (3) (2002) 853–869.
  • [25] Z. Balike, M. Mattei, V. Luongo, L. Frunzo, V. Deluchat, Mathematical modeling of trace-metals precipitation in biofilms, To appear.
  • [26] F. Dressel, The fundamental solution of the parabolic equation (1940).
  • [27] O. Ladyzhenskaya, V. Solonnikov, N. Ural’tseva, Linear and quasilinear equations of parabolic type, transl. math, Monographs, Amer. Math. Soc 23 (1968).
  • [28] A. Ilyin, A. Kalashnikov, O. Oleynik, Linear second-order partial differential equations of the parabolic type, Journal of Mathematical Sciences 108 (4) (2002) 435–542.
  • [29] W. Pogorzelski, Problèmes aux limites pour l’équation parabolique normale, in: Annales Polonici Mathematici, Vol. 4, Instytut Matematyczny Polskiej Akademii Nauk, 1957, pp. 110–126.