Plan
Comptes Rendus

Biological modelling / Biomodélisation
Phytoplankton dynamics
[La dynamique du phytoplancton.]
Comptes Rendus. Biologies, Volume 327 (2004) no. 11, pp. 961-969.

Résumés

We present a model of the phytoplankton dynamics. The distribution of the size of the phytoplankton aggregates is described by a non-linear transport equation that contains terms responsible for the growth of phytoplankton aggregates, their fragmentation and coagulation. We study asymptotic behaviour of moments of the solutions and we explain why phytoplankton tends to create large aggregates.

Nous présentons un modèle de dynamique du phytoplancton. La distribution de la taille des agrégats de phytoplancton est décrite par une équation de transport non linéaire, qui contient des termes responsables de la croissance des agrégats de phytoplancton, leur fragmentation et leur coagulation. Nous étudions le comportement asymptotique des moments des solutions et expliquons pourquoi le phytoplancton a tendance à créer de grands agrégats.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crvi.2004.03.013
Keywords: phytoplankton, dynamics, growth, formation of aggregates
Mot clés : phytoplancton, dynamique, croissance, formation d'agrégats
Ovide Arino 1 ; Ryszard Rudnicki 2

1 UR GEODES IRD–Bondy, 32, av. Henri-Varagnat, 93143 Bondy, France
2 Institute of Mathematics, Polish Academy of Sciences and Institute of Mathematics, Silesian University, Bankowa 14, 40-007 Katowice, Poland
@article{CRBIOL_2004__327_11_961_0,
     author = {Ovide Arino and Ryszard Rudnicki},
     title = {Phytoplankton dynamics},
     journal = {Comptes Rendus. Biologies},
     pages = {961--969},
     publisher = {Elsevier},
     volume = {327},
     number = {11},
     year = {2004},
     doi = {10.1016/j.crvi.2004.03.013},
     language = {en},
}
TY  - JOUR
AU  - Ovide Arino
AU  - Ryszard Rudnicki
TI  - Phytoplankton dynamics
JO  - Comptes Rendus. Biologies
PY  - 2004
SP  - 961
EP  - 969
VL  - 327
IS  - 11
PB  - Elsevier
DO  - 10.1016/j.crvi.2004.03.013
LA  - en
ID  - CRBIOL_2004__327_11_961_0
ER  - 
%0 Journal Article
%A Ovide Arino
%A Ryszard Rudnicki
%T Phytoplankton dynamics
%J Comptes Rendus. Biologies
%D 2004
%P 961-969
%V 327
%N 11
%I Elsevier
%R 10.1016/j.crvi.2004.03.013
%G en
%F CRBIOL_2004__327_11_961_0
Ovide Arino; Ryszard Rudnicki. Phytoplankton dynamics. Comptes Rendus. Biologies, Volume 327 (2004) no. 11, pp. 961-969. doi : 10.1016/j.crvi.2004.03.013. https://comptes-rendus.academie-sciences.fr/biologies/articles/10.1016/j.crvi.2004.03.013/

Version originale du texte intégral

1 Introduction

Phytoplankton cells have the ability of forming aggregates which are dispersed in the water column as a result of currents and turbulence, leading to a patchy distribution of phytoplankton. Phytoplankton is the first level of food accessible to animals. It is, in particular, the main food available to the early larval stages of many fish species, including the anchovy. At such stages, larvae are passive and can only eat the prey passing in a very close vicinity. The best situation is when the larva is near a phytoplankton aggregate, while on the other hand larvae that stay far from aggregates are not likely to survive. Thus, being able to describe the distribution in numbers of phytoplankton aggregates of different sizes as well as locating them in the space turn out to be of utmost importance in connection with the study of fish recruitment. Recently, several authors have addressed the issue of modelling the dynamics of phytoplankton in such a way as to exhibit such structure. Using the approach of particles moving randomly under the action of currents and having at random times the ability of dividing into two new particles leads to the so-called superprocesses, for which we may refer for example to [1]. One is led to stochastic partial differential equations, whose treatment is still out of reach. Another seemingly easier approach works with approximations of densities by empirical concentrations of particles, these are models known to ecologists as advection–diffusion–reaction (ADR) models [2] and heavily used in simulations [3]. Here, results abound, unfortunately, they are first unpredictable and second unjustifiable.

The approach followed in this work is, in contrast to the above two, rather elementary. In a first study of the problem, we take the view of phenomenology: we are not introducing the specific action of the environment, we are not either describing the individual processes undergone by phytoplankton cells. We consider that the individual unit is an aggregate, aggregates are structured by their size (a definition of which will be given later), and in fact our view is that of a population (of aggregates) with some specific birth, death and growth processes. The population changes with time, the cohorts of a certain size grow or on the contrary lose some members: the various actions of currents on the individual cells are modelled phenomenologically as actions on aggregates.

Apart from growth due to cell division within an aggregate, two main mechanisms are at work: splitting of a given aggregate into parts, which is called fragmentation process, and coagulation (aggregation), by which two distinct aggregates join together to form a single one. We consider here only splitting into two parts. One could consider generally the fission into several or even the complete disassembling of an aggregate. In order to simplify its representation, we assume that if an aggregate has been fragmented into a number of pieces during some time interval, one can subdivide the time into small-enough intervals for only one binary fission to take place during each one of these time intervals.

The main role in the process of coagulation of phytoplankton play TEP (Transparent Exopolymer Particles). TEP are by-product of the growth of phytoplankton and their stickiness cause that cells will remain together upon contact [4–6]. On the other hand, the low level of concentration of TEP leads to fragmentation of phytoplankton aggregates. Again, we assume that within small-enough time intervals, coagulation is a binary process. It should be mentioned here that our description of the coagulation process is rather simple. We assume only that two distinct aggregates join together with some probability, which depends only on the size of aggregates. The coagulation is a complex physical process [5] including turbulent shear, particle settling and Brownian motion. Also porosity of aggregates and their stickiness play an important role in this process [7]. In our model, all above-mentioned factors are hidden in the probability of aggregation, which makes mathematics much simpler.

The view we just briefly described is saving us from the tedious alternate way that would consist in modelling first and cumulating the various forces entailed by currents and the turbulence, on the one hand, as well those forces of a biotic nature, which altogether would make up the state of an aggregate. While we are not aware of another comparable approach for the modelling of phytoplankton aggregates, it has been used and is still being used in the completely different context of polymerisation/depolymerisation of chemical or biochemical species [8–11]. What we will show here is that, under a number of assumptions that we will briefly discuss further on, the higher moments of the distribution of the population of aggregates tend to infinity. It means that phytoplankton tends to create large aggregates.

2 Description of the model and assumptions

The first step is to describe the state variable of the problem. The state at a given time t is the distribution at that time of all the aggregates according to their size. What we call the size of an aggregate is either the number of cells forming the aggregate or the total mass of those cells. It could be also the sum of the lengths of the cells, in the case length is a relevant parameter. Weight and length, as structuring variables, are impaired by the fact that there is a significant heterogeneity of these parameters. We denote x the generic size. In terms of x, the state of the system is characterized at any moment t by the density u(x,t). We will assume that the state can be represented by a function, or rather a class of functions, that is, the map tu(.,t) is continuous from the set of times into a space of Lebesgue measurable functions. In fact, the choice of the right space is easy to make: the total mass of cells (or equivalently, the number of cells in all the aggregates) should be finite at all time, that is:

0xu(x,t)dx<
In this way we obtain the space of work, namely,
X={ϕ:0x|ϕ(x)|dx=defϕX<}
We will also use the cone X+, which consists of all non-negative functions from X.

2.1 Growth and mortality

Here, we consider the processes at the level of a single aggregate. Aggregates grow as a result of divisions of phytoplankton cells and may just die, for example, by sinking to the seabed, or whatever cause. We assume that both processes depend on the actual size of the aggregate.

Definition 1

We assume that the growth rate is a function b(x), smooth enough, such that b(x)>0 for all x>0, b(0)=0, b(0)>0 and that there exists some constant b¯ such that b(x)b¯x. The mortality rate is a function d(x), which we assume continuous and bounded.

If the dynamics were just the result of growth and death, the equation would read:

ut(x,t)=x[b(x)u(x,t)]d(x)u(x,t)(1)

2.2 Fragmentation

Fragmentation involves (at least) two concepts.

Definition 2

(1) The ability of aggregates of a certain size to break. This ability is modelled by a function p(x). During a small time interval Δt, a fraction p(x)Δt of the aggregates of size x are undertaking breakup. We assume that p is a continuous, bounded and non-negative function.

(2) Once an aggregate breaks (into two pieces, as already mentioned), the size of the two pieces is described in terms of a conditional density K(x,y), that is, a non-negative measurable function defined in the positive quadrant, with support in the set {(x,y):x<y}, such that:

  • (i) 0yK(x,y)dx=1, for all x>0
  • (ii) K(x,y)=K(yx,y), for all x, y, y>x

Part (ii) of the definition of K has the following straightforward consequence:

0yxK(x,y)dx=y2for ally>0(2)
According to Eq. (2), the expected size of fragments of aggregates of size y is just y2. If fragmentation were the only process at work, the equation for the dynamics would read:
ut(x,t)=p(x)u(x,t)+2xK(x,y)p(y)u(y,t)dy(3)
It is just a matter of standard computation to check that multiplying the equation on both sides by x and integrating the result from 0 to ∞ will give that the right-hand side is 0, that is to say, the total number of cells remains constant under a pure fragmentation process.

2.3 Coagulation

Until now, we have considered linear processes only. Coagulation of pairs of aggregates is, by the very fact, non-linear. It should normally depend on the space. In this work, space is not explicitly considered, so we are assuming that aggregates of any size are somehow uniformly distributed. As for the fragmentation, we also assume that only part of the aggregates has the competence to join. This could for example be due to the fact that only several species have the necessary devices to glue or to attach to others. The coefficient of competence is a function g(x). We assume that g is a positive, continuous, and bounded function. The population of cells that, at time t, are implicated in the coagulation process is given by:

J(t)=def0zg(z)u(z,t)dz
and
j(x,t)=defxg(x)u(x,t)J(t)
is the fraction of cells in size-x aggregates competent for the coagulation process with respect to the total population of cells in aggregates prone to join. In terms of the quantities introduced so far, we can express the time rate of cells forming aggregates of size x:
J(t)0xj(xy,t)j(y,t)dy
Again, if coagulation were the only process, the equation would read:
(xu)t(x,t)=J(t)0xj(xy,t)j(y,t)dyxg(x)u(x,t)
which, after obvious algebra, leads to:
ut(x,t)=0xu(xy,t)u(y,t)(xy)yg(xy)g(y)dyx0zg(z)u(z,t)dzg(x)u(x,t)(4)

2.4 The full equation

Taking the sums of the variations due to growth and mortality, fragmentation and coagulation, we arrive at the full equation:

ut(x,t)=x[b(x)u(x,t)]a(x)u(x,t)+2xK(x,y)p(y)u(y,t)dy+0xu(xy,t)u(y,t)(xy)yg(xy)g(y)dyx0zg(z)u(z,t)dz(5)
where we use the notation:
a(x)=d(x)+p(x)+g(x)
Eq. (5) can be written in the following abstract form:
u(t)=γu(t)+A1u(t)+A2u(t)+Bu(t)(6)
where
γ=sup{a(x):x0}(7)
(A1ϕ)(x)=ddx(b(x)ϕ(x))(8)
(A2ϕ)(x)=(γa(x))ϕ(x)+2xK(x,y)p(y)ϕ(y)dy(9)
(Bϕ)(x)=0xϕ(xy)ϕ(y)(xy)yg(xy)g(y)dyx0zg(z)ϕ(z)dz(10)
for a non-zero ϕ0 and B0=0.

Theorem 1

For each u 0 X + , there exists a unique solution u : [ 0 , [ X + of Eq. (6) such that u ( 0 ) = u 0 .

The proof of Theorem 1 can be found in Appendix A.

3 Long-term behaviour

In this section we will study the behaviour of the solution of Eq. (5) when time goes to infinity. Now assume that b(x)=bx, d(x)=d, p(x)=p, g(x)=g and that there exists a function ψ:[0,1][0,[ such that K(x,y)=1yψ(xy). Then the last assumption is very natural because:

0ryK(x,y)dx=0rψ(z)dz
which means that the size of aggregates after fragmentation is proportional to the size of the aggregate before fragmentation. From the assumptions concerning K it follows that 01ψ(x)dx=1 and ψ(x)=ψ(1x) for x[0,1]. We assume that, for each non-negative integer n, we have:
01xn+1ψ(x)dx=defcn<(11)
From properties of ψ it follows easily that c0=12. Moreover, the sequence (cn) is decreasing.

It will be a little easier to study the behaviour of the function v(x,t)=xu(x,t) instead of u. Recall that x1x2v(x,t)dx is the number of cells in all aggregates with size between x1 and x2. We will write v(t)(x)=v(x,t) and, for each t0, the function v(t) is an element of the space L1+[0,[ of all integrable functions ϕ:[0,[[0,[. The function v satisfies the following equation:

v(t)=Av(t)(d+p+g)v(t)+pKv(t)+gJv(t)(12)
where
(Aϕ)(x)=bxϕ(x)(13)
(Kϕ)(x)=x2xy−2ψ(x/y)ϕ(y)dy(14)
(Jϕ)(x)=0xϕ(xy)ϕ(y)dy0ϕ(z)dz(15)
for a non-zero ϕX+ and J0=0.

We can assume that b=d. If bd, then we can substitute v(t)=eλtv¯(t), where λ=bd. Then from homogeneity of the operator J and linearity of others operators in Eq. (12), it follows that v¯ satisfies Eq. (12), with b=d¯=d+λ.

For each non-negative integer n, we consider the space Yn of all measurable functions ϕ from [0,[ to R such that the function (1+x++xn)ϕ(x) is integrable. Let

ϕn=0(1+x++xn)|ϕ(x)|dx
be the norm in Yn. We will also use the cone Yn+, which consists of all non-negative functions from Yn. For ϕYn we denote by Mn(ϕ) the nth moment of ϕ, i.e. Mn(ϕ)=0xnϕ(x)dx. First we formulate a result similar to Theorem 1.

Theorem 2

For eachv0Yn+there exists a unique solutionv:[0,[Yn+of Eq.(12)such thatv(0)=v0. Moreover, for each non-negative integer n, we have:

ddtMn(v(t))=βnMn(v(t))+(M0(v(0)))−1gk=0n(nk)Mk(v(t))Mnk(v(t))(16)
whereβn=bn+2cnppg.

The proof of Theorem 2 can be found in Appendix B. Now we study the long-term behaviour of the solutions of Eq. (16). For n=0 Eq. (16) reduces to ddtM0(v(t))=0. This implies that M0(v(t))=M0(v(0)). We simplify the notation by setting wn(t)=Mn(v(t))/M0(v(0)). Then Eq. (16) takes the form:

wn(t)={β¯nwn(t)forn=1β¯nwn(t)+gk=1n1(nk)wk(t)wnk(t)forn2(17)
where β¯n=βn+2g=bn+2cnpp+g.

4 Discussion

To our knowledge, little is known about the solution behaviour of equations like (17). The precise analysis of Eqs. (5) and (17) is difficult and we omit it here. We note that if β¯1=b+(2c11)p+g<0, then w1(t)0 as t. Consequently, if the fragmentation rate p is large in comparison with birth and coagulation rates b and g, then β¯1<0 and then the average size of aggregates tends to zero.

The case β¯1>0 is more interesting. Then the average size of aggregates tends to infinity. Roughly, it means that aggregates with larger size make up an essential part of the whole population of the phytoplankton. We can say more about long-term behaviour of the distribution of aggregates if we assume a stronger inequality (2c11)p+g>0. Consider a stochastic process Xt such that the nth moments of Xt is wn(t). Let Yt=eβ¯1tXt and let mn(t) denotes the nth moments of Yt. Then m1(t)=const, and the function mn(t) satisfies equation:

mn(t)=(β¯nnβ¯1)mn(t)+gk=1n1(nk)mk(t)mnk(t)(18)
for n2. Since β¯nnβ¯1<0 for all n2, one can check that limtmn(t)=mn*, where (mn*) is a sequence of positive constants. It means that the distribution of random variables Yt is weakly convergent to the distribution of a random variable Y with the nth moments mn*. Consequently, the random variable Xt has the distribution like eβ¯1tY as t. In this case, the size of almost all aggregates tends to infinity, which is rather unrealistic, and it is connected with the assumption that the behaviour of aggregates does not depend on their size.

In order to control the growth of the size, we should assume, for example, that the ability of aggregates to break up p(x) depends on the size x and that it is an increasing function. One can check that if p(x)=px and other coefficients are the same, i.e. b(x)=bx, d(x)=d, and g(x)=g, then there exists a stationary distribution of the size of aggregates. We suppose that in this case the distribution of the size of aggregates converges to a stationary distribution when time goes to infinity.

Acknowledgments

This research was partially supported by the State Committee for Scientific Research (Poland) Grant No. 2 P03A 031 25 and by the EC programme Centres of Excellence for States in phase of pre-accession, No. ICA1-CT-2000-70024.

Appendix A Proof of Theorem 1

First observe that A1 is the infinitesimal generator of a C0 semigroup of positive-bounded linear operators on X. Indeed, let πtx0 denote the solution of the equation x(t)=b(x(t)) with x(0)=x0, i.e. πtx0=x(t). If ϕ is a differentiable function, then the initial value problem:

ut(x,t)=x[b(x)u(x,t)],u(0,x)=ϕ(x)(A.1)
has a unique classical solution of the form:
u(t,x)=ϕ(πtx)x(πtx)(A.2)
Let Ptϕ(x)=ϕ(πtx)x(πtx) for ϕX. Observe that {Pt}t0 is a C0 semigroup of linear positive bounded operators on X. Indeed, for ϕX, we have:
0x|Ptϕ(x)|dx=0x|ϕ(πtx)|x(πtx)dx=0πty|ϕ|(y)dy(A.3)
Since b(x)b¯x, we have πtyeb¯ty for t,y0 and we finally obtain:
Ptϕ0eb¯ty|ϕ(y)|dy=eb¯tϕ(A.4)
It is easy to check that the operator A1 with domain D(A1)={ϕX:A1ϕX} is the infinitesimal generator of the semigroup {Pt}t0. One can easily check that boundedness of functions a and p and condition (2) imply that A2 is a bounded positive linear operator on Y. From the Phillips perturbation theorem [12], it follows that the operator A=γI+A1+A2 with the domain D(A)=D(A1) is the infinitesimal generator of a C0 semigroup {St}t0 of linear bounded and positive operators on X.

Now we check that the operator B satisfies a global Lipschitz condition on the set X+={ϕX:ϕ0}. In the proof we use the following notation: Gϕ(x)=xg(x)ϕ(x) and α(ϕ)=0Gϕ(x)dx. Then:

Bϕ(x)=(Gϕ*Gϕ)(x)xα(ϕ)
where * denote the convolution on the positive halfline. Fix a function ϕ0X+{0}. Let c=sup{g(x):x0} and ɛ=α(ϕ0)c−1. Let ϕ be any function from X+{0} such that ϕϕ0ɛ. Then α(ϕ)2α(ϕ0). We have:
BϕBϕ0=(Gϕ*Gϕ)α(ϕ0ϕ)xα(ϕ0)α(ϕ)+G(ϕ+ϕ0)*G(ϕϕ0)xα(ϕ0)(A.5)
This implies that:
BϕBϕ00(Gϕ*Gϕ)(x)dxα(|ϕ0ϕ|)α(ϕ0)α(ϕ)+0[G(ϕ+ϕ0)*|G(ϕϕ0)|](x)dxα(ϕ0)(A.6)
Since
0(Gϕ*Gϕ)(x)dx=[0(Gϕ)(x)dx]2=[α(ϕ)]2(A.7)
0[G(ϕ+ϕ0)*|G(ϕϕ0)|](x)dx=α(ϕ+ϕ0)α(|ϕϕ0|)(A.8)
and α(ϕ)2α(ϕ0), from (A.6) it follows that:
BϕBϕ0α(ϕ)α(|ϕ0ϕ|)α(ϕ0)+α(ϕ+ϕ0)α(|ϕϕ0|)α(ϕ0)(A.9)
5α(|ϕϕ0|)5cϕϕ0(A.10)
Now we check this inequality for all ϕ,ψX+{0}. Fix ϕ,ψX+{0} and let ϕt=(1t)ϕ+tψ for t[0,1]. Since the function tα(ϕt) is continuous and α(ϕt)>0 for each t[0,1] we have inftα(ϕt)>0. Let ɛ¯=c−1inftα(ϕt). Then from (A.10) it follows that BϕsBϕt5cϕsϕt if ϕsϕtɛ¯. Let n be an integer such that nϕψ/ɛ¯ and let ti=i/n for i=0,1,,n. Then ϕtiϕti1ɛ¯ and consequently:
BϕBψi=1nBϕtiBϕti15ci=1nϕtiϕti1=5cϕψ(A.11)
By continuity the inequality passes to the limit at ϕ=0 or ψ=0. The rest of the proof of the existence and uniqueness of solutions of Eq. (6) is a simple consequence of the method of variation of parameters (see, e.g., [13]).

Remark 1

An anonymous referee pointed us out that the proof of the Lipschitz condition for B could be improved. Indeed, from (A.5) it follows that the operator B has at the point ϕ the Fréchet derivative DϕB of the form:

(DϕB(ψ))(x)=(Gϕ*Gϕ)(x)xα2(ϕ)α(ψ)+2(Gϕ*Gψ)(x)xα(ϕ)(A.12)
and therefore
DϕB(ψ)3|α(ψ)|3cψ(A.13)
Since X+{0} is a convex subset of the Banach space X we have B(ψ)B(ψ0)3cψψ0 for ψ,ψ0X+{0}.

Appendix B Proof of Theorem 2

First, let us note that Eq. (16) can be obtained by multiplying both sides of (12) by xn and integration with respect to x in the interval [0,[. But then we should a priori know that the corresponding integral exists and that:

0xn+1vx(x,t)dx=0(n+1)xnv(x,t)dx
which is not obvious.

We start with the definition of the semigroup generated by the operator A. Let us define Qtϕ(x)=ϕ(ebtx) for ϕYn. Then {Qt}t0 is a C0 semigroup of linear positive bounded operators on Yn with the infinitesimal generator A. Moreover, for ϕYn, we have:

Mn(Qtϕ)=0xnϕ(ebtx)dx=0eb(n+1)tynϕ(y)dy=eb(n+1)tMn(ϕ)(B.1)
From (B.1) it follows that the operator Gn=Ab(n+1)I generates a semigroup {Q˜t}t0 such that Mn(Q˜tϕ)=Mn(ϕ).

We also have:

Mn(Kϕ)=0x2xn+1y−2ψ(x/y)ϕ(y)dydx=0(0y2xn+1ψ(x/y)dx)y−2ϕ(y)dy=02cnynϕ(y)dy=2cnMn(ϕ)(B.2)
for ϕYn, which implies that K is a linear, bounded and positive operator on Yn. Let Hn=(bnpg)I+pK. Then from (B.2) it follows that Mn(Hnϕ)=βnMn(ϕ) for ϕYn. This implies that the semigroup {St}t0 generated by the operator Gn+Hn has the property:
Mn(Stϕ)=eβntMn(ϕ)(B.3)
for ϕYn.

Now, we check some properties of the operator J. First observe that for ϕ,ψYn we have:

Mn(ϕ*ψ)=00xxnϕ(xy)ψ(y)dydx=00(y+z)nϕ(z)ψ(y)dydz=k=0n(nk)Mk(ϕ)Mnk(ψ)(B.4)
and therefore
Mn(Jϕ)=(M0(ϕ))−1k=0n(nk)Mk(ϕ)Mnk(ϕ)(B.5)
for ϕYn+{0}. Define α(ϕ)=0ϕ(x)dx. If ϕYn+ and α(ϕ)=1 then Mkn(ϕ)Mnk(ϕ) for 1kn. From this inequality and from (B.4) it follows that:
Mn(ϕ*ψ)=k=0n(nk)Mk(ϕ)Mnk(ψ)k=0n(nk)Mnk/n(ϕ)Mn(nk)/n(ψ)(Mn1/n(ϕ)+Mn1/n(ψ))n2n1(Mn(ϕ)+Mn(ψ))(B.6)
for ϕ,ψYn+ such that α(ϕ)=α(ψ)=1. If ϕYn+ and α(ϕ)=1 then Mn(J(ϕ))=Mn(ϕ*ϕ)2nMn(ϕ). From homogeneity of J it follows that Mn(J(ϕ))2nMn(ϕ) for all ϕYn+{0} and, consequently,
J(ϕ)n2nϕn(B.7)
Next we check that the operator J satisfies a local Lipschitz condition on the set Yn+. It can be done in a similar way as for the operator B in Appendix A but now we apply the method described in Remark 1. The operator J has at the point ϕYn+{0} the Fréchet derivative DϕJ of the form:
DϕJ(ψ)=ϕ*ϕα2(ϕ)α(ψ)+2ϕ*ψα(ϕ)(B.8)
for ψYn+ and from (B.6) it follows that:
Mn(|DϕJ(ψ)|)2n+1Mn(ϕ)α(|ψ|)α(ϕ)+2nMn(|ψ|)(B.9)
Thus
DϕJ(ψ)n2n+1ϕnα(|ψ|)α(ϕ)+2nψn2nψn(1+2ϕn/α(ϕ))(B.10)
and therefore J is a locally Lipschitz operator. Let v(0)Yn+. A continuous function v:[0,T[Yn+ is a solution of (12) if and only if
v(t)=Stv(0)+g0tStrJv(r)drfort[0,T[(B.11)
Using standard arguments based on the Banach principle one can check that Eq. (B.11) has a unique solution defined in some interval [0,T[.

Now we prove (16). From (B.3) and (B.5) it follows that:

Mn(v(t))=eβntMn(v(0))+g0teβn(tr)Mn(Jv(r))drfort[0,T[(B.12)
Since integral equation (B.12) is equivalent to the differential equation:
ddtMn(v(t))=βnMn(v(t))+gMn(Jv(t))(B.13)
it follows from (B.5) that:
ddtMn(v(t))=βnMn(v(t))+(M0(v(t)))−1gk=0n(nk)Mk(v(t))Mnk(v(t))(B.14)
For n=0, Eq. (B.14) reduces to ddtM0(v(t))=0. This implies that M0(v(t))=M0(v(0)) and, consequently, Eq. (16) holds for t[0,T[.

Finally, we check that Eq. (12) has a unique solution v:[0,[Xn+ for every v(0)Xn+. Since operators A and K are linear and the operator J is homogeneous it is enough to consider the case M0(v(0))=1. Contrary to our claim let assume that the solution v is only defined on a bounded interval [0,T[. Since {St}t0 is a C0 semigroup there exists a positive constant c1 such that Stnc1 for t[0,T[. From (B.7) and (B.11) it follows that:

v(t)nc1v(0)n+2nc1g0tv(r)ndr(B.15)
From the above integral inequality it follows that there exists a positive constant c2, which depends only on v(0)n, such that:
sup0t<Tv(t)nc2(B.16)
Let V(c)={ϕYn+:M0(ϕ)=1,ϕnc} for c>0. Then v(t)V(c2) for each t[0,T[. From (B.10) it follows that DϕJ(ψ)n2n(1+2c)ψn for ϕV(c) and ψYn. Since the set V(c) is convex the operator J is Lipschitzean on V(c) for each c>0. It is easy to check that there exists a constant δ>0 such that for each ϕV(c2) Eq. (12) has a solution v¯:[0,δ]V(2c2) such that v¯(0)=ϕ. This contradicts the assumption that the solution v was only defined on a bounded interval.


Bibliographie

[1] R. Adler Superprocesses and plankton dynamics, Proc. ‘Aha Huliko'a Hawaiian Winter Workshop (1997)

[2] W.R. Young Brownian bugs and superprocesses, Proc. 12th ’Aha Huliko'a Hawaiian Winter Workshop (2001)

[3] W.R. Young; A.J. Roberts; G. Stuhne Reproductive pair correlations and the clustering of organisms, Nature, Volume 412 (2001) no. 6844, pp. 328-331

[4] H.E. Dam; D.T. Drapeau Coagulation efficiency, organic-matter glues and the dynamics of particle during a phytoplankton bloom in a mesocosm study, Deep-Sea Res. II, Volume 42 (1995) no. 1, pp. 111-123

[5] G.A. Jackson A model of the formation of marine algal flocs by physical coagulation processes, Deep-Sea Res., Volume 37 (1990), pp. 1197-1211

[6] U. Passow; A.L. Alldredge Aggregation of a diatom bloom in a mesocosm: The role of transparent exopolymer particles (TEP), Deep-Sea Res. II, Volume 42 (1995) no. 1, pp. 99-109

[7] G.A. Jackson Comparing observed changes in particle size spectra with those predicted using coagulation theory, Deep-Sea Res. II, Volume 42 (1995) no. 1, pp. 159-184

[8] A.S. Ackleh; K. Deng On the first-order hyperbolic coagulation model, Math. Models Appl. Sci., Volume 26 (2003), pp. 703-715

[9] J. Banasiak On a non-uniqueness in fragmentation models, Math. Models Appl. Sci., Volume 25 (2002), pp. 541-556

[10] F.E. Boyd; M. Cai; H. Han Rate equation and scaling for fragmentation with mass loss, Phys. Rev. A, Volume 41 (1990) no. 10, pp. 5755-5757

[11] R.M. Ziff; E.D. McGrady Kinetics of polymer degradation, Macromolecules, Volume 19 (1986), pp. 2513-2519

[12] N. Dunford; J.T. Schwartz Linear Operators, Part I, Interscience Publ., New York, 1968

[13] A. Pazy Semigroups of Linear Operators and Applications to Partial Differential Equations, Appl. Math. Sci., vol. 44, Springer-Verlag, New York, 1983


Commentaires - Politique


Ces articles pourraient vous intéresser

Analysis of the effects of fragmentation-coagulation in planktology

Suares Clovis Oukouomi Noutchie

C. R. Biol (2010)


On conservativity and shattering for an equation of phytoplankton dynamics

Jacek Banasiak

C. R. Biol (2004)


On phytoplankton aggregation: a view from an IBM approach

Nadjia El Saadi; Alassane Bah

C. R. Biol (2006)