Loading [MathJax]/jax/output/SVG/jax.js

Multiple patterns formation for an aggregation/diffusion predator-prey system

  • Received: 01 September 2020 Published: 20 May 2021
  • 35B40, 35B36, 35Q92, 45K05, 92D25

  • We investigate existence of stationary solutions to an aggregation/diffusion system of PDEs, modelling a two species predator-prey interaction. In the model this interaction is described by non-local potentials that are mutually proportional by a negative constant α, with α>0. Each species is also subject to non-local self-attraction forces together with quadratic diffusion effects. The competition between the aforementioned mechanisms produce a rich asymptotic behavior, namely the formation of steady states that are composed of multiple bumps, i.e. sums of Barenblatt-type profiles. The existence of such stationary states, under some conditions on the positions of the bumps and the proportionality constant α, is showed for small diffusion, by using the functional version of the Implicit Function Theorem. We complement our results with some numerical simulations, that suggest a large variety in the possible strategies the two species use in order to interact each other.

    Citation: Simone Fagioli, Yahya Jaafra. Multiple patterns formation for an aggregation/diffusion predator-prey system[J]. Networks and Heterogeneous Media, 2021, 16(3): 377-411. doi: 10.3934/nhm.2021010

    Related Papers:

    [1] Simone Fagioli, Yahya Jaafra . Multiple patterns formation for an aggregation/diffusion predator-prey system. Networks and Heterogeneous Media, 2021, 16(3): 377-411. doi: 10.3934/nhm.2021010
    [2] Rinaldo M. Colombo, Francesca Marcellini, Elena Rossi . Biological and industrial models motivating nonlocal conservation laws: A review of analytic and numerical results. Networks and Heterogeneous Media, 2016, 11(1): 49-67. doi: 10.3934/nhm.2016.11.49
    [3] Mostafa Bendahmane . Analysis of a reaction-diffusion system modeling predator-prey with prey-taxis. Networks and Heterogeneous Media, 2008, 3(4): 863-879. doi: 10.3934/nhm.2008.3.863
    [4] Ivano Primi, Angela Stevens, Juan J. L. Velázquez . Pattern forming instabilities driven by non-diffusive interactions. Networks and Heterogeneous Media, 2013, 8(1): 397-432. doi: 10.3934/nhm.2013.8.397
    [5] Marie Henry . Singular limit of an activator-inhibitor type model. Networks and Heterogeneous Media, 2012, 7(4): 781-803. doi: 10.3934/nhm.2012.7.781
    [6] Stefan Berres, Ricardo Ruiz-Baier, Hartmut Schwandt, Elmer M. Tory . An adaptive finite-volume method for a model of two-phase pedestrian flow. Networks and Heterogeneous Media, 2011, 6(3): 401-423. doi: 10.3934/nhm.2011.6.401
    [7] José Antonio Carrillo, Yingping Peng, Aneta Wróblewska-Kamińska . Relative entropy method for the relaxation limit of hydrodynamic models. Networks and Heterogeneous Media, 2020, 15(3): 369-387. doi: 10.3934/nhm.2020023
    [8] Panpan Xu, Yongbin Ge, Lin Zhang . High-order finite difference approximation of the Keller-Segel model with additional self- and cross-diffusion terms and a logistic source. Networks and Heterogeneous Media, 2023, 18(4): 1471-1492. doi: 10.3934/nhm.2023065
    [9] Julien Barré, Pierre Degond, Diane Peurichard, Ewelina Zatorska . Modelling pattern formation through differential repulsion. Networks and Heterogeneous Media, 2020, 15(3): 307-352. doi: 10.3934/nhm.2020021
    [10] Verónica Anaya, Mostafa Bendahmane, Mauricio Sepúlveda . Mathematical and numerical analysis for Predator-prey system in a polluted environment. Networks and Heterogeneous Media, 2010, 5(4): 813-847. doi: 10.3934/nhm.2010.5.813
  • We investigate existence of stationary solutions to an aggregation/diffusion system of PDEs, modelling a two species predator-prey interaction. In the model this interaction is described by non-local potentials that are mutually proportional by a negative constant α, with α>0. Each species is also subject to non-local self-attraction forces together with quadratic diffusion effects. The competition between the aforementioned mechanisms produce a rich asymptotic behavior, namely the formation of steady states that are composed of multiple bumps, i.e. sums of Barenblatt-type profiles. The existence of such stationary states, under some conditions on the positions of the bumps and the proportionality constant α, is showed for small diffusion, by using the functional version of the Implicit Function Theorem. We complement our results with some numerical simulations, that suggest a large variety in the possible strategies the two species use in order to interact each other.



    The mathematical modelling of the collective motion through aggregation/diffusion phenomena has raised a lot of interest in the recent years and it has been deeply studied for its application in several areas, such as biology [9, 39, 47, 48], ecology [35, 42, 43], animal swarming [3, 4, 40, 46] sociology and economics, [49, 50, 51]. One of the common idea in this modelling approach is that a certain population composed by agents evolves according to long-range attraction and short-range repulsion forces between agents. We are interested in modelling the problem of predator-prey interactions, namely we consider two populations that attract (prey) and repel (predators) each others. The pioneering works for this problem are the ones by Lotka, [36] and Volterra[54], which describe the predator-prey interaction via reaction terms in a set of differential equations, possibly combined with diffusion terms, see [41] and the references therein.

    As in [21], in this paper we model predator-prey interactions via a transport terms rather than a reaction ones as follows: consider N predators located at X1,,XNRn, and M prey at Y1,,YMRn with masses miX>0 and miY>0 respectively. We assume that each agent of the same species interacts under the effect of a radial non-local force that is attractive in the long range and repulsive in the short range. Moreover, predators are attracted by the prey, while the latter are subject to a repulsive force from the predators, that is proportional to the previous one. This set of modelling assumptions leads to the following system of ODEs:

    {˙Xi(t)=Nk=1mkX(Sr1(Xi(t)Xk(t))+Sa1(Xi(t)Xk(t)))Mh=1mhYK(Xi(t)Yh(t)),˙Yj(t)=Mh=1mhY(Sr2(Xi(t)Xk(t))+Sa2(Yj(t)Yh(t)))+αMk=1mkXK(Yj(t)Xk(t)), (1)

    with i=1,,N and j=1,,M. The potentials Sa1 and Sa2 are called self-interaction and model the long-range attraction among agents of the same species. The potential K is responsible for the predator-prey interaction, and it is called cross-interaction potential. The coefficient α>0 models the escape propensity of prey from the predators. The short-range repulsion among particles of the same species is modelled by the non-local forces Sr1 and Sr2, and we can assume that it scales with the number of particles, Sri(z)=NβS(Nβ/nz) for a smooth functional S, see [40].

    The formal limit when the number of particles tends to infinity leads to the following system of partial differential equations

    {tρ=div(ρ(d1ρSa1ρKη)),tη=div(η(d2ηSa2η+αKρ)), (2)

    where ρ and η are the densities of predators and prey respectively. Through this limit the (non-local) short-range repulsion formally turns to a (local) nonlinear diffusion terms, being d1 and d2 positive constants modelling the spreading velocity, while the long-range attraction takes into account the non-local self-interactions. We can therefore lighten the notation by calling Sa1=Sρ and Sa2=Sη.

    The goal of this paper is to show that the model above catches one of the main features that occur in nature, namely the formation of spatial patterns where the predators are surrounded of empty zones and the prey aggregates around, that is usually observed in fish schools or in flock of sheeps, see [32, 38]. In the fully aggregative case, namely system (2) with d1=d2=0, the formation of these types of patterns has been studied in several papers, see [15, 27, 21, 46] and references therein.

    Existence theory for solutions to system of the form (2) can be performed in the spirit of [10, 20]. In particular, system (2) should be framed in the context of non symmetrizable systems, for which the Wasserstein gradient flow theory developed in [1] and adapted to systems in [22] does not work. In [10, 20, 22], the authors consider different choices for the diffusion part (no diffusion in [22], diagonal nonlinear diffusion in [10] and cross-diffusion with dominant diagonal in [20]), and the existence of solutions is proved via an implicit-explicit version of the JKO scheme [33].

    In the following, we reduce our analysis to the one-dimensional setting

    {tρ=x(ρx(d1ρSρρKη))tη=x(ηx(d2ηSηη+αKρ)). (3)

    We are interested in the existence of stationary solutions to (3), which are solutions to the following system

    {0=(ρ(d1ρSρρKη)x)x,0=(η(d2ηSηη+αKρ)x)x, (4)

    as well as their properties, e.g., symmetry, compact support, etc. We shall assume throughout the paper that d1=d2=θ for simplicity. Note that we can always assume this by a simple scaling argument. Indeed, it is enough to multiply the first equation in (4) by d2/d1, and setting d2=θ, ˜Sρ=d2d1Sρ, ˜K=d2d1K and ˜α=d1d2α to get

    {0=(ρ(θρ˜Sρρ˜Kη)x)x,0=(η(θηSηη+˜α˜Kρ)x)x.

    The stationary equation for the one species case, i.e.,

    tρ=x(ρx(θρSρ))

    is studied several papers, see [2, 7, 11, 16] and therein references. In [7] the Krein-Rutman theorem is used in order to characterise the steady states as eigenvectors of a certain non-local operator. The authors prove that a unique steady state with given mass and centre of mass exists provided that θ<KL1, and it exhibits a shape similar to a Barenblatt profile for the porous medium equation; see [52] and [24] for the local stability analysis. Similar techniques are used in [8] in order to partly extend the result to more general nonlinear diffusion, see also [34]. This approach is used in [6] in order to explore the formation of segregated stationary states for a system similar to (3) but in presence of cross-diffusion. Unfortunately, when dealing with systems, it is not possible to reproduce one of the major issues solved in [7], namely the one-to-one correspondence between the diffusion constant (eigenvalue) and the support of the steady state. A support-independent existence result for small diffusion coefficient θ is obtained in [6] by using the generalised version of the implicit function theorem, see also [5] where this approach is used in the one species case.

    In this paper we apply the aforementioned approach in order to show that stationary solutions to (3) are composed of multiple Barenblatt profiles. Let us introduce, for fixed zρ,zη>0, the following space

    M={(ρ,η)(LL1(R))2:ρ,η0,ρL1=zρ,ηL1=zη}.

    Definition 1.1. We say that a pair (ρ,η)M is a multiple bumps steady state to (3) if the pair (ρ,η) solves (4) weakly and there exist two numbers Nρ,NηN, and two families of intervals Iiρ=[liρ,riρ], for i=1,...,Nρ, and Ihη=[lhη,rhη], for h=1,...,Nη such that

    IiρIjρ=, for i,j=1,...,Nρ, ij and IhηIkη=, for h,k=1,...,Nη, hk,

    ρ and η are supported on

    supp(ρ)=Nρi=1Iiρandsupp(η)=Nηi=1Iiη,

    respectively and

    where, for i=1,...,Nρ and h=1,...,Nη, ρi and ηh are even w.r.t the centres of Iiρ and Ihη respectively, non-negative and C1 functions supported on that intervals.

    Example 1.1. A possible example of steady states as defined above is a steady state (ρ,η) consisting of three bumps for each one of the pair (ρ,η) (Nρ=Nη=3), namely, ρ1,ρ2,ρ3 and η1,η2,η3 respectively, with centers of masses {cmiρ:=riρliρ}Nρi=1 and {cmiη:=riηliη}Nηi=1 as solutions to system (17). A plot is carried out in Figure 1 which shows all the properties stated in Definition 1.1.

    Figure 1. 

    A possible example of a stationary solution to (3) with Nρ=Nη=3 is plotted as described in Definition 1.1

    .

    Remark 1.1. We remark that one should be careful with finding a shape of steady state. More precisely, one should choose the numbers Nρ,Nη and the centers of masses {cmiρ}Nρi=1 and {cmiη}Nηi=1 such that all the conditions required for the existence of steady states are satisfied, see Theorem 1.1.

    In order to simplify the notations, in the following we will denote with l{ρ,η} a generic index that recognise one of the two families. Throughout the paper we shall assume that the kernels satisfy the following:

    (A1) Sρ, Sη and K are C2(R) functions.

    (A2) Sρ, Sη and K are radially symmetric and decreasing w.r.t. the radial variable.

    (A3) Sρ, Sη and K are non-negative, with finite L1-norm supported on the whole real line R.

    Note that assumption (A2) together with the sign in front of the nonlocal terms Sρ and K (in the first equation) and Sη in system (3), give the effect of an attractive potential, i.e. a radial interaction potential G(x)=g(|x|) for some g:[0,+)R, such that g(r)>0 for r>0. For K in the second equation we obtain the effect of a repulsive potential, i.e. g(r)<0 for r>0.

    The main result of the paper is the following

    Theorem 1.1. Assume that Sρ, Sη and K are interaction kernels are under the assumptions (A1), (A2) and (A3). Consider Nρ,NηN and let zil be fixed positive numbers for i=1,2,,Nl, and l{ρ,η}. Consider two families of real numbers {cmiρ}Nρi=1 and {cmiη}Nηi=1 such that

    (i) {cmiρ}Nρi=1 and {cmiη}Nηi=1 are stationary solutions of the purely non-local particle system, that is, for i=1,2,,Nl, for l,h{ρ,η} and lh,

    Bil=Nlj=1Sl(cmilcmjl)zjl+αlNhj=1K(cmilcmjh)zjh=0, (5)

    (ii) the following quantities

    Dil=Nlj=1Sl(cmilcmjl)zjlαlNhj=1K(cmilcmjh)zjh, (6)

    are strictly positive, for all i=1,2,,Nl, l,h{ρ,η} and lh.

    where αρ=1 and αη=α. Then, there exists a constant θ0 such that for all θ(0,θ0) the stationary equation (4) admits a unique solution in the sense of Definition 1.1 of the form

    where

    each interval Iil is symmetric around cmil for all i=1,2,,Nl, l{ρ,η},

    ρi and ηj are C1, non-negative and even w.r.t the centres of Iiρ and Ijη respectively, with masses ziρ and zjη, for i=1,...,Nρ and j=1,...,Nη,

    the solutions ρ and η have fixed masses

    zρ=Nρi=1ziρandzη=Nηi=1ziη,

    respectively.

    The paper is structured as follows. In Section 2 we recall the basics notions on optimal transport and we introduce the p-Wasserstein distances in spaces of probability measures. Then, we recall the strategy for proving existence to systems of the form (2). The remaining part of the Section is devoted to a preliminary and partial existence analysis of steady states via the Krein-Rutman theorem of a particular type of stationary solutions that we call mixed steady state. Section 3 is devoted to the proof of Theorem 1.1 in which existence and uniqueness results for multiple bumps stationary solutions are proved in case of small diffusion coefficient using the Implicit Function Theorem. We conclude the paper with Section 4, complementing our results with numerical simulations that also show interesting stability issues of the stationary states, namely transitions between states and others effects such as traveling waves profiles.

    We start collecting preliminary concepts on the Wasserstein distance. Let P(Rn) be the space of probability measures on Rn and fix p[1,+). The space of probability measures with finite p-moment is defined by

    Pp(Rn)={μP(Rn):mp(μ)=Rn|x|pdμ(x)<}.

    For a measure μP(Rn) and a Borel map T:RnRk, denote with T#μP(Rn) the push-forward of μ through T, defined by

    Rkf(y)dT#μ(y)=Rnf(T(x))dμ(x)forallfBorelfunctionsonRk.

    We endow the space Pp(Rn) with the Wasserstein distance, see for instance [1, 45, 53]

    Wpp(μ,ν)=infγΓ(μ,ν){Rn×Rn|xy|pdγ(x,y)}, (7)

    where Γ(μ1,μ2) is the class of transport plans between μ and ν, that is the class of measures γP(Rn×Rn) such that, denoting by πi the projection operator on the i-th component of the product space, the marginality condition πi#γ=μi i=1,2 is satisfied.

    Since we are working in a 'multi-species' structure, we consider the product space Pp(Rn)×Pp(Rn) endowed with a product structure. In the following we shall use bold symbols to denote elements in a product space. For a p[1,+], we use the notation

    Wpp(μ,ν)=Wpp(μ1,ν1)+Wpp(μ2,ν2),

    with μ=(μ1,μ2),ν=(ν1,ν2)Pp(Rn)×Pp(Rn). In the one-dimensional case, given μP(R), we introduce the pseudo-inverse variable uμL1([0,1];R) as

    uμ(z)inf{xR:μ((,x])>z},z[0,1], (8)

    see [14].

    In the Section 1 we mentioned that the well-posedness of (3) can be stated according to the results in [10, 20] in an arbitrary space dimension n. In these papers, the existence of weak solutions is provided using an implicit-explicit version of the Jordan-Kinderlehrer-Otto (JKO) scheme [33, 22], that we will sketch in the following. A key point in this approach is to associate to (3) a relative energy functional

    F[μ,ν](ρ,η)=θ2Rnρ2+η2dx12RnρSρρdx12RnηSηηdxRnρKμdx+αRnηKνdx,

    for a fixed reference couple of measures (μ,ν). We state our definition of weak measure solution for (3), in the space P2(Rn)2:=Pp(Rn)×Pp(Rn).

    Definition 2.1. A curve μ=(ρ(),η()):[0,+)P2(Rn)2 is a weak solution to (3) if

    (i) ρ,ηL2([0,T]×Rn) for all T>0, and ρ,ηL2([0,+)×Rn) for i=1,2,

    (ii) for almost every t[0,+) and for all ϕ,φCc(Rn), we have

    ddtRnϕρdx=θRnρρϕdx+Rnρ(Sρρ+Kη)ϕdx,ddtRnφηdx=θRnηηφdx+Rnη(SηηαKρ)ϕdx.

    Theorem 2.1. Assume that (A1)-(A3) are satisfied. Let μ0=(ρ1,0,ρ2,0)P2(Rn)2 such that

    F[μ0](μ0)<+.

    Then, there exists a weak solution to (3) in the sense of Definition 2.1 with μ(0)=μ0.

    As already mentioned the proof of Theorem 2.1 is a special case of the results in [10, 20] and consists in the following main steps:

    1. Let τ>0 be a fixed time step and consider the initial datum μ0P2(Rn)2, such that F[μ0](μ0)<+. Define a sequence {μkτ}kN recursively: μ0τ=μ0 and, for a given μkτP2(Rn)2 with k0, we choose μk+1τ as follows:

    μk+1τargminμP2(Rn)2{12τW22(μkτ,μ)+F[μkτ](μ)}. (9)

    For a fixed T>0, set N:=[Tτ], and

    μτ(t)=(ρτ(t),ητ(t))=μkτt((k1)τ,kτ],kN,

    with μkτ defined in (9).

    2. There exists an absolutely continuous curve ˜μ:[0,T]P2(Rn)2 such that the piecewise constant interpolation μτ admits a sub-sequence μτh narrowly converging to ˜μ uniformly in t[0,T] as h+. This is a standard result coming from the minimising condition (9).

    3. There exist a constant C>0 such that

    T0[||ρτ(t,)||2H1(Rn)+||ητ(t,)||2H1(Rn)]dtC(T,μ0), (10)

    and the sequence μτh:[0,+)P2(Rn)2 converges to ˜μ strongly in

    L2((0,T)×Rn)×L2((0,T)×Rn).

    The estimate in (10) can be deduced by using the so called Flow-interchange Lemma introduced in [37], see also [25]. In order to deduce the strong convergence we use the extended version of the Aubin-Lions Lemma in [44].

    4. The approximating sequence μτh converges to a weak solution ˜μ to (3). This can be showed considering two consecutive steps in the semi-implicit JKO scheme (9), i.e. μkτ, μk+1τ, and perturbing in the following way

    μϵ=(ρϵ,ηϵ)=(Pϵ#ρk+1τ,ηkτ), (11)

    where Pϵ=id+ϵζ, for some ζCc(Rn;Rn) and ϵ0. From the minimizing property of μk+1τ we have

    012τ[W22(μk+1τ,μϵ)W22(μkτ,μk+1τ)]+F[μkτ](μϵ)F[μkτ](μϵ). (12)

    After some manipulations, sending first ϵ0 and then τ0 the inequality (12) leads to the first weak formulation in Definition 2.1. Perturbing now on η and repeating the same procedure we get the required convergence.

    The existence of weak solutions to the purely non-local systems, i.e.,

    {tρ=div(ρ(Sρρ+Kρη)),tη=div(η(Sηη+Kηρ)), (13)

    with generic cross-interaction kernels Kρ and Kη is investigated in [22], whereas studies on the shape of stationary states can be found in [17, 27]. Concerning the predator-prey modelling and patterns formation, in [15, 46] a minimal version of (1) has been considered with only one predator and arbitrarily many prey subject to (different) singular potentials. This model induces the formation of nontrivial patterns in some way to prevent the action of the predators. In [21] the authors study existence and stability of stationary states for the purely aggregative version of system (3), namely equation (3) with θ=0,

    {tρ=div(ρ(Sρρ+Kη)),tη=div(η(SηηαKρ)), (14)

    and its relation with the particle system

    {˙Xi(t)=Nk=1mkXSρ(Xi(t)Xk(t))Mk=1mkYK(Xi(t)Yk(t)),˙Yj(t)=Mk=1mkYSη(Yj(t)Yk(t))+αNk=1mkXK(Yj(t)Xk(t)). (15)

    It is proved that stationary states of system (14) are linear combinations of Dirac's deltas, namely ˉρ,ˉηP(Rn), with

    (ˉρ,ˉη)=(Nk=1mkXδˉXk(x),Mh=1mhYδˉYh(x)). (16)

    where {ˉXk}k, {ˉYh}h are stationary solutions of system (15), i.e.,

    {0=Nk=1Sρ(ˉXkˉXi)mkX+Mh=1K(ˉYhˉXi)mhY0=Mh=1Sη(ˉYhˉYj)mhYαNk=1K(ˉXkˉYj)mkX. (17)

    for i=1,...,N and j=1,...,M, see also [29, 30] for a symilar result in the one-species case. As pointed out in [21], system (17) is not enough to determine a unique steady state, since the linear combination of the first N equations, weighted with αmiX, and the final M equations weighted with coefficients mjY get the trivial identity 0=0. System (17) should be coupled with the quantity

    Cα=αNi=1miXXiMj=1mjYYj (18)

    that is a conserved quantities, and therefore one would like to produce a unique steady state once the quantity Cα is prescribed. Solutions to system (17) will play a crucial role in the proof of the main Theorem 1.1.

    As a preliminary result, we now prove the existence of one possible shape of steady state, that will be a prototype example for the general case. The steady state is what we can call a mixed steady state, that identifies the case in which the predators can catch the prey, see Figure 2.

    Figure 2. 

    Example of mixed stationary state. Note that by symmetry Lρ=Rρ and Lη=Rη

    .

    The proof of the existence of such steady state follows by using the strong version of the Krein-Rutman theorem, see [26].

    Theorem 2.2 (Krein-Rutman). Let X be a Banach space, KX be a solid cone, such that λKK for all λ0 and K has a nonempty interior Ko. Let T be a compact linear operator on X, which is strongly positive with respect to K, i.e. T[u]Ko if uK{0}. Then,

    (i) the spectral radius r(T) is strictly positive and r(T) is a simple eigenvalue with an eigenvector vKo. There is no other eigenvalue with a corresponding eigenvector vK.

    (ii) |λ|<r(T) for all other eigenvalues λr(T).

    As pointed out in [6], using this strategy we can only obtain existence of stationary states for a diffusion coefficient that depends on the support, without providing any one-to-one correspondence between the diffusion constant (eigenvalue) and the support. Even if non completely satisfactory, the following results give a useful insight on the possible conditions we can expect in order to get existence of steady states, see Remark 3.1.

    Let us first introduce a proper definition for mixed steady states as in Figure 2.

    Definition 2.2. Let 0<Rρ<Rη be fixed. We call a pair (ρ,η) a mixed steady state a solution to system (3) with ρ and η in L1L(R), non-negative, symmetric and radially decreasing functions with supports

    Iρ:=supp(ρ)=[Rρ,Rρ],andIη:=supp(η)=[Rη,Rη].

    Let us now assume that (ρ,η) is a steady state to system (3) as in Definition 2.2, then (4) can be rephrased as

    {θρ(x)Sρρ(x)Kη(x)=CρxIρθη(x)Sηη(x)+αKρ(x)=CηxIη. (19)

    where Cρ,Cη>0 are two constants. Differentiating the two equations in (19) w.r.t. xsupp(ρ) and xsupp(η) respectively, we obtain

    {θρx=RρRρSρ(xy)ρy(y)dy+RηRηK(xy)ηy(y)dyxIρθηx=RηRηSη(xy)ηy(y)dyαRρRρK(xy)ρy(y)dyxIη. (20)

    By symmetry properties of the kernels Sρ, Sη and K and the steady states ρ and η, for x>0, we get

    θρx=Rρ0(Sρ(xy)Sρ(x+y))ρy(y)dy+Rη0(K(xy)K(x+y))ηy(y)dy,θηx=Rη0(Sη(xy)Sη(x+y))ηy(y)dyαRρ0(K(xy)K(x+y))ρy(y)dy. (21)

    In order to simplify the notations, we set

    ˜G(x,y):=G(xy)G(x+y),forG=Sρ,Sη,K.

    Notice that ˜G, under assumptions (A1)-(A3), is a nonnegative function for x,y>0. We also set p(x)=ρx(x) for x(Rρ,Rρ) and q(x)=ηx(x) for x(Rη,Rη). Hence, (21) is rewritten simply as

    {θp(x)=Rρ0˜Sρ(x,y)p(y)dy+Rη0˜K(x,y)q(y)dyθq(x)=Rη0˜Sη(x,y)q(y)dyαRρ0˜K(x,y)p(y)dy. (22)

    Proposition 2.1. Assume that Sρ,Sη,K satisfy (A1), (A2) and (A3) and fix 0<Rρ<Rη and 0<zρ, zη. Assume that Sη and K are strictly concave on [Rη,Rη] and [Rρ,Rρ] respectively. Assume that there exists a constant C such that

    C<Rη0˜Sη(x,y)h(y)dyRρ0˜K(x,y)k(y)dy, (23)

    for any (k,h)C1([0,Rρ])×C1([0,Rη]) with k(0)=h(0)=0 and k(0)>0, k(x)>0 for all x(0,Rρ), and h(0)>0, h(x)>0 for all x(0,Rη). Then, there exists a unique mixed steady state (ρ,η) in the sense of Definition 2.2 to system (3) with θ=θ(Rρ,Rη)>0, provided that

    α<min{C,Sη(Rη)zηR2ηK(0)zρ},

    where zρ and zη are masses of ρ and η respectively.

    Proof. Let us first introduce the following Banach space

    XRρ,Rη={(p,q)C1([0,Rρ])×C1([0,Rη]):p(0)=q(0)=0},

    endowed with the W1,-norm for the two components p and q. Define the operator TRρ,Rη[P] on the Banach space XRρ,Rη as

    TRρ,Rη[P]:=(f,g)C1([0,Rρ])×C1([0,Rη]),

    where P denotes the elements P=(p,q)XRρ,Rη, and (f,g) are given by

    f(x)=Rρ0˜Sρ(x,y)p(y)dy+Rη0˜K(x,y)q(y)dyforx[0,Rρ],g(x)=Rη0˜Sη(x,y)q(y)dyαRρ0˜K(x,y)p(y)dyforx[0,Rη].

    The assumptions on the kernels ensure that the operator TRρ,Rη is compact on the Banach space XRρ,Rη. Indeed, as the operator TRρ,Rη is defined on a space of C1 functions defined on compact intervals, and since the kernels Sρ,Sη,K are all in C2, which is from assumption (B1), defined in the operator on compact intervals, then using Arzelá's theorem, it is easy to prove that TRρ,Rη maps bounded sequences in XRρ,Rη into pre-compact ones. Now, consider the subset KRρ,RηXRρ,Rη defined as

    KRρ,Rη={PXRρ,Rη:p0,q0}.

    It can be shown that this set is indeed a solid cone in KRρ,Rη. Moreover, we have that

    HRρ,Rη={PKRρ,Rη:p(0)>0,p(x)>0forallx(0,Rρ),andq(0)>0,q(x)>0forallx(0,Rη)}KRρ,Rη,

    where KRρ,Rη denotes the interior of KRρ,Rη. We now show that the operator TRρ,Rη is strongly positive on the solid cone KRρ,Rη in the sense of Theorem 2.2. Let (p,q)KRρ,Rη with p,q0, then by the definition of the operator TRρ,Rη, it is easy to see that the first component is non-negative. Concerning the second component, we have

    Rη0˜Sη(x,y)q(y)dyαRρ0˜K(x,y)p(y)dy>0, (24)

    if and only if α<C with C as in (23). Next, it is easy to show that the derivative at x=0 of the first component is strictly positive. The derivative of the second component is given by

    ddx|x=0(Rη0˜Sη(x,y)q(y)dyαRρ0˜K(x,y)p(y)dy)=Rη0˜Sη,x(0,y)q(y)dyαRρ0˜Kx(0,y)p(y)dy=Rη0(Sη(y)Sη(y))q(y)dyαRρ0(K(y)K(y))p(y)dy=2Rη0Sη(y)q(y)dy+2αRρ0K(y)p(y)dy:=A.

    Now, we need to find the condition on α such that A>0. Chebyshev's inequality in the first integral of A and the concavity assumption of Sη on the interval [Rη,Rη] yields the bound

    2RηRη0Sη(y)q(y)dy=2RηRη0Sη(y)η(y)dy(1RηRη0Sη(y)dy)(2RηRη0η(y)dy)=Sη(Rη)zηR2η.

    Using the concavity assumption of K on the interval [Rρ,Rρ], the other integral can be easily bounded by

    2Rρ0K(y)p(y)dy=2Rρ0K(y)ρ(y)dy<K(0)zρ.

    Thus, A>0 holds under the condition

    α<Sη(Rη)zηR2ηK(0)zρ. (25)

    As a consequence, TRρ,Rη[P] belongs to HRρ,Rη, which implies that the operator TRρ,Rη is strongly positive on the solid cone KRρ,Rη. Then, the Krein-Rutman theorem applies and guarantees the existence of an eigenvalue θ=θ(Rρ,Rη) such that

    TRρ,Rη[P]=θP,

    with an eigenspace generated by one given nontrivial element (ˉp,ˉq) in the interior of the solid cone KRρ,Rη. Moreover, by (i) of Theorem 2.2, there exists no other eigenvalues to TRρ,Rη with corresponding eigenvectors in KRρ,Rη besides the one with eigenfunction (ˉp,ˉq), and by (ii) of Theorem 2.2 all other eigenvalues ˜θ with eigenfunctions in XRρ,Rη satisfy |˜θ|<θ.

    Remark 2.1. Motivated from the purely aggregative case (13), we actually expect a more rich behavior for the possible steady states configurations, such as the separated stationary state in Figure 3. This is expected as a possible winning strategy for the prey, since it corresponds to prey finding a safe distance from the predators. Unfortunately, the symmetrization argument used in the previous proof seems to fail, since in this case we need to require the symmetry around cm2η for the convolutions. In the next section we prove the existence and uniqueness for the multiple bumps stationary state in the sense of Definition 1.1, that includes the mixed and the separated one, using a completely different approach.

    Figure 3. 

    An example of a separated stationary state

    .

    In this Section we prove the existence and uniqueness of a multiple bumps steady state in the sense of Definition 1 fixing masses and a small diffusion coefficient. Following the approach in [5, 6], we first formulate the problem in terms of the pseudo-inverse functions and then we use the Implicit Function Theorem (cf. [19, Theorem 15.1]).

    We start rewriting our stationary system in terms of pseudo-inverse functions. Let (ρ,η) be a solution to the stationary system

    {0=(ρ(θρSρραρKη)x)x,0=(η(θηSηηαηKρ)x)x. (26)

    where αρ=1 and αη=α. Assume that (ρ,η) have masses zρ and zη respectively and denote by cml,l{ρ,η}, the centres of masses

    Rxρ(x)dx=cmρ,Rxη(x)dx=cmη.

    Remember that the only conserved quantity in the evolution, together with the masses, is the joint centre of mass

    CMα=αcmρcmη, (27)

    that we can consider fixed. Define the cumulative distribution functions of ρ and η as

    Fρ(x)=xρ(x)dx,Fη(x)=xη(x)dx.

    Let ul:[0,zl]R,l{ρ,η}, be the pseudo-inverse of Fl, namely

    ul(z)=inf{xR:Fl(x)z},l{ρ,η},

    supported on

    supp(ul)=[0,zl]:=Jl,l{ρ,η}.

    For ρ and η multiple bumps in the sense of Definition 1.1 we can denote the mass of each bump as

    ρi(x)dx=ziρ,ηj(x)dx=zjη,i=1,2,,Nρ,j=1,2,,Nη,

    and the centres of masses accordingly,

    xρi(x)dx=cmiρ,xηj(x)dx=cmjη,i=1,2,,Nρ,j=1,2,,Nη,

    and we can always assume that the centres of masses are ordered species by species, i.e. cmilcmjl if ij. Let us consider the case of centres of masses that are stationary solutions of the purely non-local particle system (17), that we recall for the reader convenience,

    {Nρj=1Sρ(cmiρcmjρ)zjρ+Nηj=1K(cmiρcmjη)zjη=0,i=1,,Nρ,Nηj=1Sη(cmiηcmjη)zjηαNρj=1K(cmiηcmjρ)zjρ=0,i=1,,Nη, (28)

    coupled with the conservation of the joint centre of mass CMα in (27), see the discussion in Section 2. For such a density the pseudo-inverse ul reads as

    ul(z)=Nli=1uil(z)1Jil(z),l{ρ,η},

    where

    supp(ul)=[0,zl]=Jl=Nli=1[ik=1zk1l,ik=1zkl]:=Nli=1[ˆzil,˜zil]:=Nli=1Jil,l{ρ,η},

    with z0l=0 and zl=Nlk=1zkl. We are now in the position of reformulating (26) in terms of the pseudo-inverse functions as follows:

    {θ2z((zuρ(z))2)=JρSρ(uρ(z)uρ(ξ))dξ+αρJηK(uρ(z)uη(ξ))dξ,zJρ,θ2z((zuη(z))2)=JηSη(uη(z)uη(ξ))dξ+αηJρK(uη(z)uρ(ξ))dξ,zJη. (29)

    The restriction to zJil, i=1,2,,Nl, and l{ρ,η}, allow us to rephrase (29) in the compact form

    θ2z((zuil(z))2)=Nlj=1JjlSl(uil(z)ujl(ξ))dξ+αlNhj=1JjhK(uil(z)ujh(ξ))dξ,zJil. (30)

    Similar to [5, 6], we suggest the linearization

    uil=cmil+δvili=1,2,,Nl,andl{ρ,η},

    with vil, being odd functions defined on Jil. Using this ansatz in (30), with the scaling θ=δ3 we have

    δ2z((zvil(z))2)=Nlj=1JjlSl(cmilcmjl+δ(vil(z)vjl(ξ)))dξ+αlNhj=1JjhK(cmilcmjh+δ(vil(z)vjh(ξ)))dξ. (31)

    Multiplying (31) by δzvil, and taking the primitives w.r.t. z, we obtain

    δ2zvil(z)=Nlj=1JjlSl(cmilcmjl+δ(vil(z)vjl(ξ)))dξ+αlNhj=1JjhK(cmilcmjh+δ(vil(z)vjh(ξ)))dξ+Ail,zJil, (32)

    where Ail are the integration constants. In order to recover the constants Ail, we substitute ˜zil into equation (32). Denoting by vil(˜zil)=λil, we obtain

    Ail=Nlj=1JjlSl(cmilcmjl+δ(λilvjl(ξ)))dξαlNhj=1JjhK(cmilcmjh+δ(λilvjh(ξ)))dξ. (33)

    Next, we set Gl and H such that Gl=Sl and H=K, with Gl,H to be odd and satisfy Gl(0)=H(0)=0. Then, multiplying (32) again by δzvil and taking the primitives w.r.t. zJil, we obtain

    δ3z=Nlj=1JjlGl(cmilcmjl+δ(vil(z)vjl(ξ)))dξ+αlNhj=1JjhH(cmilcmjh+δ(vil(z)vjh(ξ)))dξ+Ailδvil(z)+βil,zJil. (34)

    Let us denote with ˉzil the middle point of each interval Jil. Then, in order to recover the integration constants βil, we substitute ˉzil in (34) which yields

    βil=δ3ˉzilNlj=1JjlGl(cmilcmjlδvjl(ξ))dξαlNhj=1JjhH(cmilcmjhδvjh(ξ))dξ. (35)

    As a consequence of all above computations, we get the following relation for zJil,

    δ3(zˉzil)=Nlj=1JjlGl(cmilcmjl+δ(vil(z)vjl(ξ)))Gl(cmilcmjlδvjl(ξ))dξδvil(z)Nlj=1JjlSl(cmilcmjl+δ(λilvjl(ξ)))dξ+αlNhj=1JjhH(cmilcmjh+δ(vil(z)vjh(ξ)))H(cmilcmjhδvjh(ξ))dξδvil(z)αlNhj=1JjhK(cmilcmjh+δ(λilvjh(ξ)))dξ. (36)

    If we define, for p=(v1ρ,,vNρρ,v1η,,vNηη)

    Fil[p;δ](z)=ˉzilz+δ3[Nlj=1JjlGl(cmilcmjl+δ(vil(z)vjl(ξ)))Gl(cmilcmjlδvjl(ξ))dξδvil(z)Nlj=1JjlSl(cmilcmjl+δ(λilvjl(ξ)))dξ+αlNhj=1JjhH(cmilcmjh+δ(vil(z)vjh(ξ)))H(cmilcmjhδvjh(ξ))dξδvil(z)αlNhj=1JjhK(cmilcmjh+δ(λilvjh(ξ)))dξ],zJil, (37)

    we have that (30) reduces to the equation Fil[p;δ](z)=0. In the following we compute the Taylor expansion of Fil[p;δ](z) around δ=0. Let us begin with the first integral on the r.h.s. of (37), i.e.,

    JjlGl(cmilcmjl+δ(vil(z)vjl(ξ)))Gl(cmilcmjlδvjl(ξ))dξ=[Sl(cmilcmjl)δvil(z)+δ22Sl(cmilcmjl)(vil(z))2+δ36Sl(cmilcmjl)(vil(z))3]|Jjl|+Jjlδ32Sl(cmilcmjl)(vjl(ξ))2vil(z)dξ+R(Sl,δ4), (38)

    where we used the fact that Jilvil(ξ)dξ=0 and R(Sl,δ4) is a remainder term. For the second integral we have

    δvil(z)JjlSl(cmilcmjl+δ(λilvjl(ξ)))dξ=[Sl(cmilcmjl)δvil(z)δ2Sl(cmilcmjl)λilvil(z)δ32Sl(cmilcmjl)(λil)2vil(z)]|Jjl|Jjlδ32Sl(cmilcmjl)(vjl(ξ))2vil(z)dξ+R(Sl,δ4). (39)

    Summing up the contributions in (38) to (39), we get that the self-interaction part in (37) reduces to

    δ3[δ12Sl(cmilcmjl)vil(z)(vil(z)2λil)+16Sl(cmilcmjl)((vil(z))33vil(z)(λil)2)]|Jjl|+R(Sl,δ4). (40)

    Similarly, for the cross-interaction part we obtain

    δ3[δ12K(cmilcmjh)vil(z)(vil(z)2λil)+16K(cmilcmjh)((vil(z))33vil(z)(λil)2)]|Jjh|+R(K,δ4). (41)

    Putting together the contributions of (40) and (41) in the functional equation (37), we get

    Fil[p;δ](z)=(ˉzilz)+Dil6(3vil(z)(λil)2(vil(z))3)+δ1Bil2vil(z)(vil(z)2λil)+R(Sl,K,δ4), (42)

    where we used the notations introduced in (6) and (5), namely

    Dil=Nlj=1Sl(cmilcmjl)|Jjl|αlNhj=1K(cmilcmjh)|Jjh|,

    and

    Bil=Nlj=1Sl(cmilcmjl)|Jjl|+αlNhj=1K(cmilcmjh)|Jjh|.

    Note that since the values cmil satisfy (28) we have that Bil=0. After the manipulations above, equation Fil[p;0](z)=0 reads as

    (ˉzilz)+Dil6(3vil(z)(λil)2(vil(z))3)=0,zJil, (43)

    that gives a unique solution once the value of λil is determined. In order to do that, we evaluate the functional Fil in the end point ˜zil of the corresponding interval

    Λil[p;δ]=Fil[p;δ](˜zil). (44)

    Performing Taylor expansions similar to the ones in (38) and (39) we get that

    Λil[p;0]=(ˉzil˜zil)+Dil3(λil)3, (45)

    and we are now in the position to solve

    {(ˉzilz)+Dil6(3vil(z)(λil)2(vil(z))3)=0,(ˉzil˜zil)+Dil3(λil)3=0. (46)

    The second equation in (46) admits a solution once we have that Dil>0, and λil is uniquely determined by

    λil=(3(˜zilˉzil)Dil)1/3. (47)

    By construction λilλjl if ij, and this implies that equation (46) admits the unique solution ˉvil which can be recovered as the pseudo inverse of the following Barenblatt type profiles

    ˉρi(x)=Diρ2((λiρ)2(xcmiρ)2)1Iiρ(x),i=1,,Nρ,ˉηh(x)=Dhη2((λhη)2(xcmhη)2)1Ihη(x),h=1,,Nη, (48)

    where the intervals Iiρ=[liρ,riρ] and Ihη=[lhη,rhη] are determined imposing

    likk=cmikkλikk,rikk=cmikk+λikk,ik=1,,Nk,k=ρ,η.

    We are now ready to reformulate (36) as a functional equation on a proper Banach space. Consider the spaces

    Ωil={vL([ˉzil,˜zil))|vincreasing,v(ˉzil)=0},i=1,,Nl,l{ρ,η}, (49)

    endowed with the L norm and take the product spaces

    Ωl=Nl×i=1Ωil,forl{ρ,η}.

    We now introduce the space Ω defined by

    Ω=Ωρ×RNρ×Ωη×RNη, (50)

    with elements ω=(v1ρ,,vNρρ,λ1ρ,,λNρρ,v1η,,vNηη,λ1η,,λNηη) endowed with the norm

    |||ω|||=Nρi=1(viρL+|λiρ|)+Nηi=1(viηL+|λiη|). (51)

    For γ>0, calling ˜Jil=[ˉzil,˜zil), we consider the norm

    |||ω|||γ=|||ω|||+l{ρ,η}Nli=1supz˜Jil|λilvil(z)|(˜zilz)γ, (52)

    and set Ωγ:={ωΩ:|||ω|||γ<+}. For a given ωΩ12, we define the operator T:Ω12Ω1

    T[ω;δ](z):=(Fρ[ω;δ](z)Λρ[ω;δ]Fη[ω;δ](z)Λη[ω;δ]), (53)

    where for l{ρ,η} we have shorted the notation introducing

    Fl[ω;δ](z)=(F1l[ω;δ](z),,FNll[ω;δ](z)),Λl[ω;δ](z):=(Λ1l[ω;δ],,ΛNll[ω;δ]). (54)

    The operator T is a bounded operator for any fixed δ0 and can be continuously extended at δ=0 to (43) and (45). In order to prove existence of stationary solutions for small δ>0 using the Implicit Function Theorem, we need to prove that the Jacobian matrix of T is a bounded linear operator form Ω1/2 to Ω1 with bounded inverse. The Jacobian of T has the following structure

    DT[ω;δ]=(DvρFρ(δ)DλρFρ(δ)DvηFρ(δ)DληFρ(δ)DvρΛρ(δ)DλρΛρ(δ)DvηΛρ(δ)DληΛρ(δ)DvρFη(δ)DλρFη(δ)DvηFη(δ)DληFη(δ)DvρΛη(δ)DλρΛη(δ)DvηΛη(δ)DληΛη(δ)), (55)

    where the components are actually matrices defined by

    DvhFl(δ)=(Fil[ω;δ]vjh(νjh))Nl,Nhi,j=1,DλhFl(δ)=(Fil[ω;δ]λjh(ajh))Nl,Nhi,j=1DvhΛl(δ)=(Λil[ω;δ]vjh(νjh))Nl,Nhi,j=1,DλhΛl(δ)=(Λil[ω;δ]λjh(ajh))Nl,Nhi,j=1,

    where νjh and ajh are generic directions. We first compute the diagonal terms in the matrix DvlFl(δ). We have

    Fil[ω;δ]vil(νil)=δ2Jilνil(ξ)[Sl(δ(vil(z)vil(ξ)))δvil(z)Sl(δ(λilvil(ξ)))Sl(δvil(ξ))]dξ+δ2νil(z)Nlj=1Jjl[Sl(cmilcmjl+δ(vil(z)vjl(ξ)))Sl(cmilcmjl+δ(λilvjl(ξ)))]dξ+δ2νil(z)αlNhj=1Jjh[K(cmilcmjh+δ(vil(z)vjh(ξ)))K(cmilcmjh+δ(λilvjh(ξ)))]dξ.

    A Taylor expansion around δ=0 similar to the ones in (38) - (41) easily gives that in the limit δ0 we obtain

    Fil[ω;0]vil(νil)=Dil2((λil)2(vil(z))2)νil(z).

    Concerning the other terms in DvlFl(δ) we get

    Fil[ω;δ]vjl(νjl)=δ2Nlj=1Jjlνjl(ξ)[Sl(cmilcmjl+δ(vil(z)vjl(ξ)))Sl((cmilcmjlδvjl(ξ))δvil(z)Sl(cmilcmjl+δ(λilvjl(ξ)))]dξ,

    that all vanish in the limit δ0. Let us now focus on the matrix DλlFl(δ). By (37) it is easy to see that the only non-zero terms in DλlFl(δ) are the diagonal ones that are given by

    Fil[ω;δ]λil(ail)=δ1vil(z)ail[Nlj=1JjlSl(cmilcmjl+δ(λilvjl(ξ)))dξ+αlNhj=1JjhK(cmilcmjh+δ(λilvjh(ξ)))dξ].

    Then, Taylor expansion w.r.t. δ yields

    Fil[ω;0]λil(ail)=Dilλilvil(z)ail.

    Since all the entrances in the matrix DλhFl(δ) are zero, the last matrix that concerns Fil is DvhFl(δ). The elements of this matrix are given by

    Fil[ω;δ]vjh(νjh)=δ2αlNhj=1Jjhνjh(ξ)[K(cmilcmjh+δ(vil(z)vjh(ξ)))K(cmilcmjhδvjh(ξ))δvil(z)K(cmilcmjh+δ(λilvjh(ξ)))]dξ,

    that vanish in the limit δ0. We now start in computing the functional derivatives for Λil in (44). Again we should consider the four matrices in (55), and we start from DvlΛl(δ). Note that the terms in the diagonal are zero in this case and the others are given by

    Λil[ω;δ]vjl(νjl)=δ2Nlj=1Jjlνjl(ξ)[Sl(cmilcmjl+δ(λil(z)vjl(ξ)))Sl((cmilcmjlδvjl(ξ))δλilSl(cmilcmjl+δ(λilvjl(ξ)))]dξ.

    The terms in DvhΛl(δ) are

    Λil[ω;δ]vjh(νjh)=δ2αlNhj=1Jjhνjh(ξ)[K(cmilcmjh+δ(λilvjh(ξ)))K(cmilcmjhδvjh(ξ))δλilK(cmilcmjh+δ(λilvjh(ξ)))]dξ,

    and the usual Taylor expansions around δ=0, show that both the matrices DvlΛl(0)=DvhΛl(0)=0. Since DλhΛl(δ) is trivially a zero matrix, only remains to compute the diagonal terms in DλlΛl(δ). We have

    Λil[ω;δ]λil(ail)=δ1ailNlj=1JjlλilSl(cmilcmjl+δ(λilvjl(ξ)))dξδ1ailαlNhj=1JjhλilKl(cmilcmjh+δ(λilvjh(ξ)))dξ.

    The last Taylor expansion gives

    Λil[ω;0]λil(ail)=Dil(λil)2ail.

    We have proved that

    DT[ω;0]=(El1dg(Diρλiρviρaiρ)000dg(Diρ(λiρ)2aiρ)0000El2dg(Diηλiηviηaiη)000dg(Diη(λiη)2aiη)), (56)

    where

    El1=dg(Diρ2((λiρ)2(viρ)2)νiρ),El2=dg(Diη2((λiη)2(viη)2)νiη),

    and dg(Ai) is a diagonal matrix with elements Ai. Let us denote by ω0 the unique solution to (46), we have the following lemma:

    Lemma 3.1. For δ>0 small enough, the operator DT[ω0;δ] is a bounded linear operator from Ω1/2 to Ω1.

    Proof. Thanks to the previous computations it is easy to see that DT is a bounded linear operator from Ω into itself and it is continuous at δ=0. The definition of the norm in (52) implies that for z˜Jil we need to control only that

    sup|||ω|||1/211(˜zilz)|Filvil[,δ](νil)Λilvil[;δ](νil)(Filvil[;0](νil)Λilvil[;0](νil))+Filλil[,δ](ail)Λilλil[;δ](ail)(Filλil[;0](ail)Λilλil[;0](ail))+Filvjh[,δ](νjh)Λilvjh[;δ](νjh)(Filvjh[;0](νjh)Λilvjh[;0](νjh))+Filλjh[,δ](ajh)Λilλjh[;δ](ajh)(Filλjh[;0](ajh)Λilλjh[;0](ajh))|0, (57)

    as δ0. We start estimating the third row in (57),

    1(˜zilz)[Filvjh[,δ](νjh)Λilvjh[;δ](νjh)(Filvjh[;0](νjh)Λilvjh[;0](νjh))]=αl(vil(z)λil)2˜zilzNhj=1JjhK(cmilcmjh+δ(λilvjh(ξ)))νjh(ξ)2dξδαl(vil(z)λil)3˜zilzNhj=1JjhK(˜x(ξ))νjh(ξ)6dξ=δαl(vil(z)λil)2˜zilzNhj=1JjhK(ˉx(ξ))(λilvjh(ξ))νjh(ξ)2dξδαl(vil(z)λil)3˜zilzNhj=1JjhK(˜x(ξ))νjh(ξ)6dξ.=αl((vil(z)λil)2˜zilz+(vil(z)λil)3˜zilz)O(δ),

    where in the first equality we did a Taylor expansion around the point x0=cmilcmjh+δ(λilvjh(ξ)) for the kernel K(cmilcmjh+δ(vil(z)vjh(ξ))), while in the second equality we did a Taylor expansion around the point x0=cmilcmjh for the kernel K(cmilcmjh+δ(λilvjh(ξ))). Similarly, we can show that

    1(˜zilz)[Filλjh[,δ](ajh)Λilλjh[;δ](ajh)(Filλjh[;0](ajh)Λilλjh[;0](ajh))]=0.

    The first two rows in (57) can be treated as follows,

    1(˜zilz)[Filvil[,δ](νil)Λilvil[;δ](νil)(Filvil[;0](νil)Λilvil[;0](νil))+Filλil[,δ](ail)Λilλil[;δ](ail)(Filλil[;0](ail)Λilλil[;0](ail))]=δ(˜zilz)[Nlj=1Jjlδ2(vil(z)λil)(νil(z)ail)Sl(cmilcmjl+δ(λilvjl(ξ)))+δ112(vil(z)λil)2(νil(z)νjl(ξ))Sl(cmilcmjl+δ(λilvjl(ξ)))+16(vil(z)λil)3(νil(z)νjl(ξ))Sl(˜x1(ξ))dξ+αlNhj=1Jjhδ2(vil(z)λil)(νil(z)ail)K(cmilcmjh+δ(λilvjh(ξ)))+δ112(vil(z)λil)2(νil(z)νjl(ξ))K(cmilcmjh+δ(λilvjh(ξ)))+16(vil(z)λil)3(νil(z)νjl(ξ))K(˜x2(ξ))dξ]=(2(νil(z)ail)(vil(z)λil)(˜zilz)+(2νil(z)1)((vil(z)λil)3(˜zilz)+(vil(z)λil)2(˜zilz)))O(δ).

    Since the functions vil are components of a vector ω belonging to Ω1/2, the quantities

    λilvil(z)(˜zilz)1/2,

    are uniformly bounded in ˜Jil, that gives (57).

    Lemma 3.2. For any δ>0 small enough, DT[ω0;0]:Ω1/2Ω1 is a linear isomorphism.

    Proof. Given wΩ1, we have to prove that

    DT[ω0;0]ω=w, (58)

    admits a unique solution ωΩ1/2 with the property

    ||ω||1/2C||w||1.

    The determinant of the matrix in (56) is given by

    detDT=(Nρi=1(Diρ)22((λiρ)2(viρ)2)(λiρ)2)(Nηi=1(Diη)22((λiη)2(viη)2)(λiη)2), (59)

    that is always different from zero under the condition Dil>0 and since (vil(z)λil)<0 on z[ˉzil,˜zil). Thanks to the structure in (56) and denoting with νil,ail and σil,kil the generic entrances in ω and w respectively, we easily get that

    νil(z)=2σil(z)Dil((vil(z))2(λil)2)+2λilvil(z)ail(vil(z))2(λil)2,ail=kilDil(λil)2,

    that implies ||νil||1/2C||σil|| for i=1,2,,Nl, and l{ρ,η}. In order to close the argument, it is enough to note that the ratio

    ailνil(z)(˜zilz)1/2,

    is uniformly bounded since (λilvil(z))/(˜zilz) is uniformly bounded, see [5, Lemma 4.4].

    We are now in the position of proving the main result of the paper, namely Theorem 1.1, that we recall below for convenience.

    Theorem 3.1. Assume that Sρ, Sη and K are interaction kernels are under the assumptions (A1), (A2) and (A3). Consider Nρ,NηN and let zil be fixed positive numbers for i=1,2,,Nl, and l{ρ,η}. Consider two families of real numbers {cmiρ}Nρi=1 and {cmiη}Nηi=1 such that

    (i) {cmiρ}Nρi=1 and {cmiη}Nηi=1 are stationary solutions of the purely non-local particle system, that is, for i=1,2,,Nl, for l,h{ρ,η} and lh,

    Bil=Nlj=1Sl(cmilcmjl)zjl+αlNhj=1K(cmilcmjh)zjh=0, (60)

    (ii) the following quantities

    Dil=Nlj=1Sl(cmilcmjl)zjlαlNhj=1K(cmilcmjh)zjh, (61)

    are strictly positive, for all i=1,2,,Nl, l,h{ρ,η} and lh.

    where αρ=1 and αη=α. Then, there exists a constant θ0 such that for all θ(0,θ0) the stationary equation (4) admits a unique solution in the sense of Definition 1.1 of the form

    where

    each interval Iil is symmetric around cmil for all i=1,2,,Nl, l{ρ,η},

    ρi and ηj are C1, non-negative and even w.r.t the centres of Iiρ and Ijη respectively, with masses ziρ and zjη, for i=1,...,Nρ and j=1,...,Nη,

    the solutions ρ and η have fixed masses

    zρ=Nρi=1ziρandzη=Nηi=1ziη,

    respectively.

    Proof. Consider zρ and zη fixed masses and a set of points cmil for i=1,2,,Nl, and l{ρ,η} that satisfy (i) and (ii). The results in Lemma 3.1 and Lemma 3.2 imply that given T defined in (53), the functional equation

    T[ω;δ](z)=0,

    admits a unique solution

    ω=(v1ρ(z),,vNρρ(z),λ1ρ,,λNρρ,v1η(z),,vNηη(z),λ1η,,λNηη),

    for δ>0 small enough. The entrances vil(z) are solutions to (36) for zJil. Consider now uil defined for zJil as uil(z)=cmil+δvil(z). Differentiating (36) twice w.r.t z we get that vil is differentiable and strictly increasing for zJil and that uil is a solution to (30). Moreover uil is also strictly increasing and we can define the inverse Fil that and its spatial derivative ρil=xFil is a solution to (4).

    Remark 3.1 Note that conditions (ii) in Theorem 3.1 turn to be conditions on the positions of the centres of masses and on the value of α. Indeed, as a sufficient condition for Diρ>0 we can assume that all the differences between the centres of masses are in the range of concavity of the kernels. Moreover, Diη>0 is satisfied if

    α<mini=1,,NηNηj=1Sη(cmiηcmjη)|Jjη|Nρj=1K(cmiηcmjρ)|Jjρ|.

    Note that the above conditions are comparable to the one we got in the proof of Section 2 using the Krein-Rutman approach.

    In this section, we study numerically solutions to system (3) using two different methods, the finite volume method introduced in [12, 13] and the particles method studied in [23, 28]. We validate the results about the existence of the mixed steady state and the multiple bumps steady states. Moreover, we perform some examples to show the variation in the behaviour of the solution to system (3) under different choices of initial data and the parameter α, which, in turn, suggests future work on the stability of the solutions to system (3). Finally, travelling waves are detected under a special choice of initial data and value for the parameter α. We begin by sketching the particles method. This method essentially consists in a finite difference discretization in space to the pseudo inverse version of system (3)

    {tuρ(z)=θ2z((zuρ(z))2)+JρSρ(uρ(z)uρ(ξ))dξ+αρJηK(uρ(z)uη(ξ))dξ,zJρtuη(z)=θ2z((zuη(z))2)+JηSη(uη(z)uη(ξ))dξ+αηJρK(uη(z)uρ(ξ))dξ,zJη, (62)

    as the following: Let NN, let {zi}Ni=1 be a sequence of points that partition the interval [0,1] uniformly. Denote by Xil(t):=ul(t,zi) the approximating particles of the pseudo inverse at each point zi of the partition. Assuming that the densities ρ and η are of unit masses, then we have the following approximating system of ODEs

    {tXiρ(t)=θ2N((ρi1(t))2(ρi(t))2)+1NNj=1Sρ(Xiρ(t)Xjρ(t))dξ+αρNNj=1K(Xiρ(t)Xjη(t))dξ,i=1,N,tXiη(t)=θ2N((ηi1(t))2(ηi(t))2)+1NNj=1Sη(Xiη(t)Xjη(t))dξ+αηNNj=1K(Xiη(t)Xjρ(t))dξ,i=1,N, (63)

    where the densities are reconstructed as

    {ρi(t)=1N(Xi+1l(t)Xil(t)),i=1,,N1,ρ0(t)=0,ρN(t)=0, (64)

    and the same for the ηi(t). To this end, we are ready to solve this particle system by applying the Runge-Kutta MATLAB solver ODE23, with initial positions Xl(0)=Xl,0={Xil,0}Ni=1,l=ρ,η determined by solving

    Xi+1ρ,0Xiρ,0ρ(t=0)dX=1N1,i=1,,N1.

    The second method we use is the finite volume method which introduced in [12] and extended to systems in [13], that consists in a 1D positive preserving finite-volume method for system (3). To do so, we first partition the computational domain into finite-volume cells Ui=[xi12,xi+12] of a uniform size Δx with xi=iΔx, i{s,,s}. Define

    ˜ρi(t):=1ΔxUiρ(x,t)dx,˜ηi(t):=1ΔxUiη(x,t)dx,

    the averages of the solutions ρ,η computed at each cell Ui. Then we integrate each equation in system (3) over each cell Ui, and so we obtain a semi-discrete finite-volume scheme described by the following system of ODEs for ¯ρi and ¯ηi

    {d˜ρi(t)dt=Fρi+12(t)Fρi12(t)Δx,d˜ηi(t)dt=Fηi+12(t)Fηi12(t)Δx, (65)

    where the numerical flux Fli+12, l=ρ,η, is considered as an approximation for the continuous fluxes ρ(θρ+Sρρ+αρKη)x and η(θη+Sηη+αηKρ)x respectively. More precisely, the expression for Fρi+12 is given by

    Fρi+12=max(ϑi+1ρ,0)[˜ρi+Δx2(ρx)i]+min(ϑi+1ρ,0)[˜ρiΔx2(ρx)i], (66)

    where

    ϑi+1ρ=θΔx(˜ρi+1˜ρi)j˜ρj(Sρ(xi+1xj)Sρ(xixj))αρj˜ηj(K(xi+1xj)K(xixj)), (67)

    and

    (ρx)i=minmod(2˜ρi+1˜ρiΔx,˜ρi+1˜ρi12Δx,2˜ρi˜ρi1Δx). (68)

    The minmod limiter in (68) has the following definition

    minmod(a1,a2,):={min(a1,a2,),ifai>0imax(a1,a2,),ifai<0i0,otherwise. (69)

    We have the same as the above expresions for η. Finally, we integrate the semi-discrete scheme (65) numerically using the third-order strong preserving Runge-Kutta (SSP-RK) ODE solver used in [31].

    In all the simulations below, we will fix the kernels as a normalised Gaussian potentials

    Sρ(x)=Sη(x)=K(x)=1πex2,

    that are under the assumptions on the kernels (A1), (A2) and (A3). This choice helps us in better understanding the variation in the behavior of the solutions w.r.t. the change in the initial data and the parameter α. In the first five examples, see Figures 4, 5, 6, 7 and 8 we show:

    Figure 4. 

    In this figure, a mixed steady state is plotted by using initial data given by (70), α=0.1, θ=0.4. Number of particles are chosen equal to number of cells in the finite volume method, which is N=71

    .
    Figure 5. 

    A separated steady state is presented in this figure. Initial data are given by (71). The parameters are α=0.2 and θ=0.4 with N=91

    .
    Figure 6. 

    This figure shows how from the initial densities ρ0,η0 and θ, as in Figure 4, a transition between mixed and a sort of separated steady state appears by choosing the value of α=6. This large value of α suggests an unstable behavior in the profile, see Remark 3.1

    .
    Figure 7. 

    A steady state of four bumps is showed in this figure starting from initial data as in (72) with α=0.05 and θ=0.3. The number of particles N=181, which is the same as number of cells

    .
    Figure 8. 

    Starting from initial data as in (73) with α=1 and θ=0.3, we get a five bumps steady state

    .

    ● in the first row steady states are plotted at the level of density, on the l.h.s. we compare the two methods illustrated above, while on the r.h.s. we show the evolution by the finite volume method,

    ● in the second row we plot the particles paths for both species obtained with the particles method,

    ● in the last row we show the pseudo inverse functions corresponding to the steady state densities.

    The last example we present shows an interesting traveling waves-type evolution.

    The first example is devoted to validating existence of mixed steady state and separated steady state. By choosing the initial data (ρ0,η0) as

    (70)

    and fixing α=0.1 and θ=0.4, we obtain a mixed steady state as plotted in Figure 4. Note that, the small value of α allows the predators to dominate the prey which results in the shape shown in Figure 4. Next, we take the initial data as

    (71)

    with the same θ as above and α=0.2. This choice of the initial data, actually, introduce two equal attractive forces on the right and left hand sides of the predators which, in turn, fix the predators at the centre and gives the required shape of the separated steady state as shown in Figure 5. Finally, a sort of separated steady state can also be obtained starting from the same initial data as in (70) and same diffusion parameter θ. If we choose α=6, the prey η will have enough speed to get out of the predators region, producing a transition between the mixed state to the separated one, this is illustrated in Figure 6.

    We then, test two cases where we validate the existence of steady states of more bumps. Let us fix θ=0.3. Then, a four-bumps steady state is performed and plotted in Figure 7, where we consider

    (72)

    as initial data and we take α=0.05. This way of choosing initial data produces a balanced attractive forces which in turn form the required steady state. Similarly, we can obtain a steady state of five bumps by using initial data

    (73)

    and α=1, see Figure 8.

    Finally, in this example, we detect existence of traveling waves. Indeed, by choosing initial data as

    (74)

    α=1 and θ=0.2, we obtain a traveling wave that is shown in Figure 9. Once the initial data and the value of θ are fixed, if α is taken as a small value, then we will come out with a mixed steady state. If we take a larger value for α then the prey will be fast escaping from the predators. Therefore, the proper value for α will produce a situation where the attack speed of the predators is equal to the escape speed of the prey, which results in a traveling wave of both the densities ρ and η. All the simulations above motivate further studies on stability vs. instability for such a system, together with the possible existence on traveling wave type solutions.

    Figure 9. 

    This last figure shows a possible existence of traveling waves by choosing initial data as in (74), α=1 and θ=0.2. The first two plots are performed by particles method, while the third one is done by finite volume method. Here we fix N=101

    .


    [1]

    L. Ambrosio, N. Gigli and G. Savaré, Gradient Flows in Metric Spaces and in the Space of Probability Measures, 2nd edition, Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 2008.

    [2] Global minimizers for free energies of subcritical aggregation equations with degenerate diffusion. Applied Mathematics Letters (2011) 24: 1927-1932.
    [3]

    N. Bellomo and S. -Y Ha, A quest toward a mathematical theory of the dynamics of swarms, Math. Models Methods Appl. Sci., 27 (2017), 745-770.

    [4]

    N. Bellomo and J. Soler, On the mathematical theory of the dynamics of swarms viewed as complex systems, Math. Models Methods Appl. Sci., 22 (2012), 1140006.

    [5] Large time behavior of nonlocal aggregation models with nonlinear diffusion. Netw. Heterog. Media (2008) 3: 749-785.
    [6] Sorting phenomena in a mathematical model for two mutually attracting/repelling species. SIAM J. Math. Anal. (2018) 50: 3210-3250.
    [7]

    M. Burger, M. Di Francesco and M. Franek, Stationary states of quadratic diffusion equations with long-range attraction, Commun. Math. Sci., 11 (2013), 709-738.

    [8]

    M. Burger, R. Fetecau and Y. Huang, Stationary states and asymptotic behavior of aggregation models with nonlinear local repulsion, SIAM J. Appl. Dyn. Syst., 13 (2014), 397-424.

    [9] Modeling the aggregative behavior of ants of the species polyergus rufescens}. Nonlinear Anal. Real World Appl. (2000) 1: 163-176.
    [10] Remarks on continuity equations with nonlinear diffusion and nonlocal drifts. J. Math. Anal. Appl. (2016) 444: 1690-1702.
    [11] Equilibria of homogeneous functionals in the fair-competition regime. Nonlinear Anal. (2017) 159: 85-128.
    [12] A finite-volume method for nonlinear nonlocal equations with a gradient flow structure. Commun. Comput. Phys. (2015) 17: 233-258.
    [13] Zoology of a nonlocal cross-diffusion model for two species. SIAM J. Appl. Math. (2018) 78: 1078-1104.
    [14]

    J. A. Carrillo and G. Toscani, Wasserstein metric and large-time asymptotics of nonlinear diffusion equations., New Trends in Mathematical Physics, World Sci. Publ., Hackensack, NJ, 2004, 234-244.

    [15]

    Y. Chen and T. Kolokolnikov, A minimal model of predator-swarm interactions, J. R. Soc. Interface, 11 (2014).

    [16] On minimizers of interaction functionals with competing attractive and repulsive potentials. Ann. Inst. H. Poincaré Anal. Non Linéaire (2015) 32: 1283-1305.
    [17] Ground states of a two phase model with cross and self attractive interactions. SIAM J. Math. Anal. (2016) 48: 3412-3443.
    [18]

    B. Düring, P. Markowich, J. -F. Pietschmann and M. -T. Wolfram, Boltzmann and Fokker-Planck equations modelling opinion formation in the presence of strong leaders, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 465 (2009), 3687-3708.

    [19]

    K. Deimling, Nonlinear Functional Analysis, Springer-Verlag, Berlin, 1985.

    [20] Nonlinear degenerate cross-diffusion systems with nonlocal interaction. Nonlinear Anal. (2018) 169: 94-117.
    [21] A nonlocal swarm model for predators-prey interactions. Math. Models Methods Appl. Sci. (2016) 26: 319-355.
    [22] Measure solutions for non-local interaction PDEs with two species. Nonlinearity (2013) 26: 2777-2808.
    [23]

    M. Di Francesco, S. Fagioli, M. D. Rosini and G. Russo, Follow-the-Leader approximations of macroscopic models for vehicular and pedestrian flows, Active Particles, Vol. 1, Birkhäuser/Springer, Cham, 2017, 333-378.

    [24] Multiple large-time behavior of nonlocal interaction equations with quadratic diffusion. Kinet. Relat. Models (2019) 12: 303-322.
    [25] Curves of steepest descent are entropy solutions for a class of degenerate convection-diffusion equations. Calc. Var. Partial Differential Equations (2014) 50: 199-230.
    [26]

    Y. Du, Order Structure and Topological Methods in Nonlinear Partial Differential Equations, Vol. 1, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2006.

    [27]

    J. Evers, R. Fetecau and T. Kolokolnikov, Equilibria for an aggregation model with two species, SIAM J. Appl. Dyn. Syst., 16 (2017), 2287-2338.

    [28] Solutions to aggregation-diffusion equations with nonlinear mobility constructed via a deterministic particle approximation. Math. Models Methods Appl. Sci. (2018) 28: 1801-1829.
    [29] Stable stationary states of non-local interaction equations. Math. Models Methods Appl. Sci. (2010) 20: 2267-2291.
    [30] Stability of stationary states of non-local equations with singular interaction potentials. Math. Comput. Modelling (2011) 53: 1436-1450.
    [31] Strong stability-preserving high-order time discretization methods. SIAM Rev. (2001) 43: 89-112.
    [32]

    S. Guo, Bifurcation and spatio-temporal patterns in a diffusive predator-prey system, Nonlinear Anal. Real World Appl., 42 (2018), 448-477.

    [33] The variational formulation of the Fokker-Planck equation. SIAM J. Math. Anal. (1998) 29: 1-17.
    [34] Stationary states of an aggregation equation with degenerate diffusion and bounded attractive potential. SIAM J. Math. Anal. (2017) 49: 272-296.
    [35] (2002) Living in Groups.Oxford University Press.
    [36] Contribution to the theory of periodic reaction. J. Phys. Chem. (1910) 14: 271-274.
    [37] A family of nonlinear fourth order equations of gradient flow type. Comm. Partial Differential Equations (2009) 34: 1352-1397.
    [38]

    M. Mimura and M. Yamaguti, Pattern formation in interacting and diffusing systems in population biology, Adv. Biophys. 15, 19-65, 1982.

    [39] A non-local model for a swarm. J. Math. Biol. (1999) 38: 534-570.
    [40]

    D. Morale, V. Capasso and K. Oelschläger, An interacting particle system modelling aggregation behavior: From individuals to populations, J. Math. Biol., 50 (2005), 49-66.

    [41]

    J. D. Murray, Mathematical Biology. I, 3rd edition, Interdisciplinary Applied Mathematics, Vol. 17, Springer-Verlag, New York, 2002.

    [42]

    A. Okubo and S. A. Levin, Diffusion and Ecological Problems: Modern Perspectives, 2nd edition, Interdisciplinary Applied Mathematics, Vol. 14, Springer-Verlag, New York, 2001.

    [43] Complexity, patterns and evolutionary trade-offs in animal aggregation. Science (1999) 254: 99-101.
    [44] Tightness, integral equicontinuity and compactness for evolution problems in Banach spaces. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) (2003) 2: 395-431.
    [45]

    F. Santambrogio, Optimal Transport for Applied Mathematicians, Progress in Nonlinear Differential Equations and Their Applications, Vol. 87, Birkhäuser/Springer, Cham, 2015.

    [46]

    H. Tomkins and T. Kolokolnikov, Swarm shape and its dynamics in a predator-swarm model, preprint, 2014.

    [47] Swarming patterns in a two-dimensional kinematic model for biological groups. SIAM J. Appl. Math. (2004) 65: 152-174.
    [48] A nonlocal continuum model for biological aggregation. Bull. Math. Biol. (2006) 68: 1601-1623.
    [49]

    M. Torregrossa and G. Toscani, On a Fokker-Planck equation for wealth distribution, Kinet. Relat. Models, 11 (2018), 337-355.

    [50]

    M. Torregrossa and G. Toscani, Wealth distribution in presence of debts. A Fokker-Planck description, Commun. Math. Sci., 16 (2018), 537-560.

    [51]

    G. Toscani, Kinetic models of opinion formation, Commun. Math. Sci., 4 (2006), 481-496.

    [52] (2007) The Porous Medium Equation. Mathematical Theory. Oxford University Press, Oxford: Oxford Mathematical Monographs, The Clarendon Press.
    [53]

    C. Villani, Topics in Optimal Transportation, Graduate Studies in Mathematics, Vol. 58, American Mathematical Society, Providence, RI, 2003.

    [54] Variazioni e fluttuazioni del numero d'individui in specie animali conviventi. Mem. Accad. Lincei Roma (1926) 2: 31-113.
  • This article has been cited by:

    1. Erin Ellefsen, Nancy Rodríguez, Nonlocal Mechanistic Models in Ecology: Numerical Methods and Parameter Inferences, 2023, 13, 2076-3417, 10598, 10.3390/app131910598
    2. Swadesh Pal, Roderick Melnik, Nonlocal Models in Biology and Life Sciences: Sources, Developments, and Applications, 2025, 15710645, 10.1016/j.plrev.2025.02.005
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1961) PDF downloads(177) Cited by(2)

Figures and Tables

Figures(9)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog