Loading [MathJax]/jax/element/mml/optable/MathOperators.js

Kinetic modelling of multiple interactions in socio-economic systems

  • Received: 01 October 2019 Revised: 01 March 2020 Published: 09 September 2020
  • Primary: 35Q20; Secondary: 35Q84, 82B21, 91D10

  • Unlike the classical kinetic theory of rarefied gases, where microscopic interactions among gas molecules are described as binary collisions, the modelling of socio-economic phenomena in a multi-agent system naturally requires to consider, in various situations, multiple interactions among the individuals. In this paper, we collect and discuss some examples related to economic and gambling activities. In particular, we focus on a linearisation strategy of the multiple interactions, which greatly simplifies the kinetic description of such systems while maintaining all their essential aggregate features, including the equilibrium distributions.

    Citation: Giuseppe Toscani, Andrea Tosin, Mattia Zanella. Kinetic modelling of multiple interactions in socio-economic systems[J]. Networks and Heterogeneous Media, 2020, 15(3): 519-542. doi: 10.3934/nhm.2020029

    Related Papers:

    [1] Fabio Camilli, Raul De Maio, Andrea Tosin . Transport of measures on networks. Networks and Heterogeneous Media, 2017, 12(2): 191-215. doi: 10.3934/nhm.2017008
    [2] Mapundi K. Banda, Michael Herty, Axel Klar . Coupling conditions for gas networks governed by the isothermal Euler equations. Networks and Heterogeneous Media, 2006, 1(2): 295-314. doi: 10.3934/nhm.2006.1.295
    [3] Yacine Chitour, Guilherme Mazanti, Mario Sigalotti . Stability of non-autonomous difference equations with applications to transport and wave propagation on networks. Networks and Heterogeneous Media, 2016, 11(4): 563-601. doi: 10.3934/nhm.2016010
    [4] Yaojun Liu, Li Jia, Ping Wang, Xiaolin Song . Joint optimization of location and allocation for spare parts depots under ($ r, Q $) inventory policy. Networks and Heterogeneous Media, 2024, 19(3): 1038-1057. doi: 10.3934/nhm.20240046
    [5] Fatih Bayazit, Britta Dorn, Marjeta Kramar Fijavž . Asymptotic periodicity of flows in time-depending networks. Networks and Heterogeneous Media, 2013, 8(4): 843-855. doi: 10.3934/nhm.2013.8.843
    [6] Eva Barrena, Alicia De-Los-Santos, Gilbert Laporte, Juan A. Mesa . Transferability of collective transportation line networks from a topological and passenger demand perspective. Networks and Heterogeneous Media, 2015, 10(1): 1-16. doi: 10.3934/nhm.2015.10.1
    [7] Mary Luz Mouronte, Rosa María Benito . Structural analysis and traffic flow in the transport networks of Madrid. Networks and Heterogeneous Media, 2015, 10(1): 127-148. doi: 10.3934/nhm.2015.10.127
    [8] Nastassia Pouradier Duteil . Mean-field limit of collective dynamics with time-varying weights. Networks and Heterogeneous Media, 2022, 17(2): 129-161. doi: 10.3934/nhm.2022001
    [9] Qinglan Xia, Shaofeng Xu . On the ramified optimal allocation problem. Networks and Heterogeneous Media, 2013, 8(2): 591-624. doi: 10.3934/nhm.2013.8.591
    [10] Klaus-Jochen Engel, Marjeta Kramar Fijavž, Rainer Nagel, Eszter Sikolya . Vertex control of flows in networks. Networks and Heterogeneous Media, 2008, 3(4): 709-722. doi: 10.3934/nhm.2008.3.709
  • Unlike the classical kinetic theory of rarefied gases, where microscopic interactions among gas molecules are described as binary collisions, the modelling of socio-economic phenomena in a multi-agent system naturally requires to consider, in various situations, multiple interactions among the individuals. In this paper, we collect and discuss some examples related to economic and gambling activities. In particular, we focus on a linearisation strategy of the multiple interactions, which greatly simplifies the kinetic description of such systems while maintaining all their essential aggregate features, including the equilibrium distributions.



    In recent times there has been an increasing interest in the notion of measure-valued solutions to evolution equations. Compared to standard approaches based on classical and weak solutions, the measure-theoretic setting allows one to better describe some interesting phenomena such as aggregation, congestion and pattern formation in a multiscale perspective. Several of these phenomena occur in applications such as vehicular traffic, data transmission, crowd motion, supply chains, where the state of the system evolves on a network, see e.g. [5,9,13,14,16].

    In order to extend the measure-valued approach to these irregular geometric structures, in this paper we study measure-valued solutions to a linear transport process defined on a network. For classical and weak solutions to transport equations on networks we refer the reader for example to [10,14,18].

    The measure-valued approach in Euclidean spaces relies on the notion of push-forward of measures along the trajectories of a vector field describing the transport paths [1,6,7,17]. The study of these problems in bounded domains poses additional difficulties, especially concerning the behaviour at the boundaries of the transported measure. For problems on networks similar difficulties arise at the vertexes.

    Our analysis is inspired by the results in [11,12], where measure-valued transport equations are studied in a bounded interval. We also refer to [15], where the authors consider instead measure-valued solutions to non-linear transport problems with measure transmission conditions at nodal points, i.e. points where the velocity vanishes.

    Consider a network Γ=(V,E), where V={Vi}iI is the set of vertexes and E={Ej}jJ is the set of arcs. We assume that the network is oriented and that a strictly positive, autonomous and Lipschitz continuous velocity field vj is defined on each arc Ej. Our aim is to describe the evolution of a mass distribution on the network Γ transported by the velocity field v(x)=jJvj(x)χEj(x). For this we will make extensive use of the fundamental fact that a generic measure μ can be written as the superposition of elementary Dirac masses, i.e.

    μ=suppμδxdμ(x), (1)

    where suppμ denotes the support of μ belonging to an appropriate σ-algebra. This representation formula has to be understood in the sense of Bochner integrals.

    From (1) it follows that if we are able to define the transport of an atomic measure δx on the network then by linearity we can transport the whole distribution μ. Hence, let us assume that the mass distribution μ0 at the initial time t=0 is given by a Dirac measure δx0, with x0Ej for some jJ. If we postulate the conservation of the mass then in the time interval (0,τ) where the mass remains inside the arc Ej the evolution of μ0 is governed by the continuity equation

    tμjt+x(vj(x)μjt)=0, (2)

    μjt being a spatial measure denoting the mass distribution along the arc Ej at time t.

    For t<τ the solution to (2) is given by the push-forward of μ0 by means of the flow map

    Φjt(x0,0):=x0+t0vj(Φjs(x0,0))ds,

    which describes the trajectory issuing from the point x0 at time t=0 and arriving at the point Φjt(x0,0)Ej at time t. Consequently, μjt is characterised as μjt(A)=μ0((Φjt)1(A)) for any measurable set AEj. Hence if μ0=δx0 then μjt=δΦjt(x0,0) for t(0,τ).

    At t=τ the trajectory tΦjt(x0,0) hits the final vertex Vi of the arc Ej. Assuming that mass concentration at the vertexes of the network is not admitted, fractions pijk of the mass carried by δΦjτ(x0,0) have then to be distributed on each outgoing arc Ek which originates from Vi.

    This preliminary discussion sketches the main ideas that we intend to follow in order to tackle the global problem on the network. We first consider a local problem, namely a transport equation on each single arc with a measure acting as a source term (boundary condition) at the initial vertex. For this local problem we formulate an appropriate notion of measure-valued solution, for which we give a representation formula taking into account also the mass which flows out of the arc. Then we glue all the solutions on the single arcs by means of appropriate mass distribution rules at the vertexes, thereby constructing the global solution on the network.

    In more detail, the paper is organised as follows. In Section 2 we introduce some notations and assumptions for the problem, while in Section 3 we review some basic facts about the measure-theoretic setting in which we will frame our analysis. In Section 4 we study the initial/boundary-value problem for the transport equation on a single bounded interval, which is the prototype of an arc of the network, then in Section 5 we move to the problem on networks. Finally, in Section 6 we construct explicit measure-valued solutions on simple networks, which constitute preliminary examples of the application of our theory to vehicular traffic.

    We start by describing the constitutive elements of the problem.

    Definition 2.1. (Network). A network Γ is a pair (V,E) where V:={Vi}iI is a finite collection of vertexes and E:={Ej}jJ is a finite collection of continuous non-self-intersecting oriented arcs whose endpoints belong to V. Each arc Ej is parameterised by a smooth function πj:[0,1]Rn. We assume that the network is connected and equipped with the topology induced by the minimum path distance.

    Given a vertex ViV, we say that an arc EjE is outgoing (respectively, incoming) if Vi=πj(0) (respectively, if Vi=πj(1)). We denote by diO (respectively, by diI) the number of outgoing (respectively, incoming) arcs in Vi and by di:=diI+diO the degree of Vi. We say that a vertex Vi is internal if diIdiO>0, that it is a source if diO=di and finally that it is a well if diI=di.

    We denote by I the set of indexes iI corresponding to the internal vertexes, by S the one corresponding to the sources and by W the one corresponding to the wells.

    Definition 2.2. (Distribution matrices). For an internal vertex Vi, iI, and for t>0 we consider a distribution (or transition) matrix {pikj(t)}diI,diOk,j=1 such that

    pikj(t)0diOj=1pikj(t)=j:Vi=πj(0)pikj(t)=1. (3)

    Here pikj(t) represents the fraction of mass which at time t flows from the incoming arc Ek to the outgoing arc Ej through the vertex Vi. Condition (3) corresponds to the fact that, unlike [11,12,15], the mass cannot concentrate at the vertexes of the network.

    For a source vertex Vi, iS, we consider instead a distribution vector {pij(t)}diOj=1 such that

    pij(t)0diOj=1pij(t)=j:Vi=πj(0)pij(t)=1. (4)

    Definition 2.3. (Velocity field). On each arc EjE we assume that a strictly positive, bounded and Lipschitz continuous velocity vj:[0,1](0,vmax] is defined, with 0<vmax<+. We denote by v=jJvjχEj the velocity field on the network (χEj being the characteristic function of the arc Ej).

    Definition 2.4. (Initial and boundary data). We prescribe the initial mass distribution over Γ as a positive measure μ0=jJμj0 with suppμj0Ej for all j. Furthermore, at all the source vertexes Vi, iS, we prescribe an inflow measure ςi with suppςi[0,T], T>0 being a certain final time.

    To define the transport of the initial measure μ0 and of the inflow measures {ςi}iS on the network Γ we describe their evolution inside an arc. On each arc Ej we take into account the inflow mass coming from the initial vertex πj(0) and we describe how the outflow mass leaving from the final vertex πj(1) is distributed to the corresponding outgoing arcs. In detail, we fix a final time T>0 and we consider the following system of measure-valued differential equations on Γ×[0,T]:

    {tμj+x(vj(x)μj)=0xEj,t(0,T],jJμjt=0=μj0xEj,jJμjVi=πj(0)={dIik=1pikj(t)μkVi=πk(1)if iIpij(t)ςiif iS, (5)

    where by μjVi=πj(0) we mean the measure flowing into the arc Ej from its initial vertex Vi=πj(0) while by μkVi=πk(1) we mean the measure flowing out of the arc Ek from its final vertex Vi=πk(1). Moreover, by pikj(t)μkVi=πk(1) we mean a measure (in time) which is absolutely continuous with respect to μkVi=πk(1) with density pikj(t) (and analogously for pij(t)ςi).

    For an internal vertex, the inflow measure is given by the mass flowing in Ej from the arcs incident to Vi=πj(0) according to the distribution rule given by the distribution matrix {pikj(t)}. For a source vertex, the inflow measure is the fraction pij() of the prescribed datum ςi entering Ej. The outflow measure, i.e. the part of the mass leaving the arc from the final vertex πj(1), is not given a priori but depends on the evolution of the measure μ inside the arc.

    The detailed study of problem (5) is postponed to Section 5. Before that, we introduce an appropriate measure theoretic setting, see Section 3, and consider preliminarily the problem on a single arc, see Section 4.

    We introduce a space of measures with an appropriate norm where we consider the solutions to our measure-valued transport equations. Moreover, since the notion of solution is based on the superposition principle (1), we briefly describe the measure-theoretic setting which guarantees the validity of this formula. We refer for details to [1,2,11,19].

    Let T be a topological space with B(T) the Borel σ-algebra in T. We denote by M(T) the space of finite Borel measures on T and by M+(T) the convex cone of the positive measures in M(T). For μM(T) and a bounded measurable function φ:TR we write

    μ,φ:=Tφdμ.

    Given a Borel measurable vector field Φ:TT, the push-forward of the measure μ under the action of Φ is an operation on μ which produces the new measure Φ#μM(T) defined by

    (Φ#μ)(E):=μ(Φ1(E)),EB(T).

    We immediately observe that Φ#μ,φ=μ,φΦ.

    Given a metric d:T×TR+ in T, we denote by BL(T) the Banach space of the bounded and Lipschitz continuous functions φ:TR equipped with the norm

    ||φ||BL:=||ϕ||+|ϕ|L,

    where the semi-norm ||L is defined by

    |φ|L:=supx,yTxy|φ(y)φ(x)|d(x,y).

    Furthermore, we introduce a norm in M(T) by taking the dual norm of ||||BL:

    μBL:=supφBL(T)||φ||BL1μ,φ.

    It is easy to see that if μM+(T) then μBL=μ(T).

    The space (M(T),BL) is in general not complete, hence it is customary to consider its completion ¯M(T)BL with respect to the dual norm. However, the cone M+(T), which is a closed subset of M(T) in the weak topology, is complete, although it is not a Banach space because it is not a vector space. Since in our model we will consider only positive measures, we restrict our attention to the complete metric space (M+(T),BL) with the corresponding distance induced by the norm.

    Remark 1. If T is bounded the Kantorovich-Rubinstein's duality theorem implies that the norm BL induces the 1-Wasserstein distance in M+(T).

    Remark 2. The distance induced in M(T) by the total variation norm:

    ||μ||TV:=supφCb(T)||φ||1μ,φ,

    where Cb(T) is the space of bounded continuous function on T, is another metric frequently used for measures. However, we observe that it may not be fully suited to transport problems where one wants to measure the distance between flowing mass distributions. Indeed, if we consider two points x,yT, xy, and the corresponding Dirac mass distributions δx,δyM+(T) centred at them we see that

    δyδxBLd(x,y),||δyδx||TV=2.

    Hence the two measures are closer and closer in the norm BL as the points x,y approach, which is consistent with the intuitive idea of transport of mass distributions; while they are not in the total variation norm, no matter how close the points x,y are.

    As alredy anticipated in Section 1, for the subsequent development of the theory we will extensively use the following fact linked to the concept of Bochner integral [2,19]: any μM+(T) can be represented as a (continuous) sum of elementary masses in the form

    μ=Tδxdμ(x)

    as a Bochner integral in (¯M(T)BL,BL).

    We now specialise the previous definitions to the case T=Γ×[0,T], where ΓRn is a network. In particular, we will call x the variable in each arc of Γ and t the variable in the interval [0,T]. We equip Γ×[0,T] with the distance

    d(x,y)+|ts|,(x,t),(y,s)Γ×[0,T],

    d being the shortest path distance on Γ.

    We consider the Borel σ-algebra B(Γ×[0,T]) given by the union of the Borel σ-algebras B([0,1]×[0,T]) corresponding to each arc Ej of Γ. Thus AB(Γ×[0,T]) if (π1j,Id)(A(Ej×[0,T]))B([0,1]×[0,T]) for all jJ, where Id denotes the identity mapping.

    A measure μ belongs to M(Γ×[0,T]) if each of its restrictions μj:=μ(Ej×[0,T]), jJ, is a finite Borel measure on Ej×[0,T]. We define the cone M+(Γ×[0,T]) analogously.

    For μM+(Γ×[0,T]) and a bounded measurable function φ:Γ×[0,T]R we write

    μ,φ:=jJEj×[0,T]φdμj. (6)

    For a function φ:Γ×[0,T]R, we denote by φj:[0,1]×[0,T]R its restriction to Ej×[0,T], i.e.:

    φ(x,t)=φj(y,t)for xEj, y=π1j(x), t[0,T].

    A function φ belongs to BL(Γ×[0,T]) if it is continuous on Γ and φjBL([0,1]×[0,T]) for every jJ. For φBL(Γ×[0,T]) the norm ||φ||BL(Γ×[0,T]) is defined by

    ||φ||BL(Γ×[0,T]):=supjJ||φj||BL([0,1]×[0,T]).

    The corresponding dual norm BL of a measure μM(Γ×[0,T]) is given by

    μBL:=supφBL(Γ×[0,T])||φ||BL(Γ×[0,T])1μ,φ.

    In this section we study the transport equation in a bounded interval. Actually, we start by focusing on the problem of prescribing appropriate initial and boundary conditions to the differential equation in R+×R+, which is an unbounded domain with boundary; then we will restrict the results to a truly bounded domain.

    Consider the conservation law

    tμ+x(v(x)μ)=0,(x,t)R+×R+, (7)

    where v:R+R is a strictly positive, bounded and Lipschitz continuous velocity field, so that the flow is one-directional and depends only on the space variable x. Given μM+(R+0×R+0), where R+0:=[0,+), owing to the disintegration theorem [1,Section 5.3] we can decompose this measure by means of its projection maps on the coordinate axes:

    ● using the projection with respect to the space variable we can write

    μ(dxdt)=μt(dx)dt, (8)

    where dt is the Lebesgue measure in time in R+0 and μtM+(R+0×{t})M+(R+0) for a.e. tR+0. The measure μt is called the conditional measure, or trace, of μ with respect to t on the fibre R+0×{t};

    ● similarly, projecting with respect to the time variable we can write

    μ(dxdt)=νx(dt)v(x)dx, (9)

    where dx is the Lebesgue measure in space in R+0 and νxM+({x}×R+0)M+(R+0) for a.e. xR+0. The measure νx is called the conditional measure, or trace, of μ with respect to x on the fibre {x}×R+0.

    Remark 3. The coefficient 1v(x) in the decomposition (9) is considered for dimensional reasons, so that νx represents actually the mass distributed on the fibre {x}×R+0.

    We incidentally notice that if μ solves (7) then the mapping xνx solves the equation xνx+ˉtνx=0, where ˉt:=1v(x)t. As far as the decomposition (8) is concerned, the mapping tμt solves instead the equation tμt+x(v(x)μt)=0.

    Relying on the concept of conditional measures, we formulate the following initial/boundary-value problem for (7):

    {tμ+x(v(x)μ)=0(x,t)R+×R+μt=0=μ0M+(R+0×{0})νx=0=ν0M+({0}×R+0) (10)

    with μM+(R+0×R+0), where:

    ● assigning an initial condition at t=0 amounts to prescribing the trace of μ on the fibre R+0×{0} according to the decomposition (8);

    ● assigning a boundary condition at x=0 amounts to prescribing the trace of μ on the fibre {0}×R+0 according to the decomposition (9).

    In order to give a suitable notion of measure-valued solution to (10), we preliminarily introduce integration-by-parts formulas useful to deal with the initial and boundary data. Let C10(R+0×R+0) be the space of continuous functions in R+0×R+0 which are differentiable in R+×R+ and vanish for x,t+. For μM+(R+0×R+0) and φC10(R+0×R+0) we set:

    tμ,φ:=μ,tφR+0φ(x,0)dμ0(x),x(v(x)μ),φ:=μ,v(x)xφR+0φ(0,t)dν0(t),

    where , denotes the duality pairing between measures and test functions in R+0×R+0, i.e. μ,φ=R+0×R+0φ(x,t)dμ(x,t). Notice that if φ is compactly supported in R+×R+ then the previous formulas agree with the usual definition of the distributional derivatives of μ.

    Remark 4. With a slight abuse of notation, in the following we will denote

    R+0φ(x,0)dμ0(x)=:μ0,φ,R+0φ(0,t)dν0(t)=:ν0,φ,

    the difference between duality pairings in R+0×R+0 and in R+0×{0} or {0}×R+0 being clear from the measures used.

    Thanks to these formulas, we are in a position to introduce the following notion of measure-valued solution to (10):

    Definition 4.1. Given μ0M+(R+0×{0}) and ν0M+({0}×R+0), a measure-valued solution to (10) is a finite measure μM+(R+0×R+0) such that

    μ,tφ+v(x)xφ=μ0,φν0,φ,φC10(R+0×R+0). (11)

    Since (10) is a linear problem, its solution can be obtained from the superposition of two measures μ1,μ2M+(R+0×R+0), where μ1 is the solution to (10) with data μt=0=μ0 and νx=0=0 while μ2 is the solution to (10) with data μt=0=0 and νx=0=ν0. This is doable in a standard way in terms of the push-forward of the initial and boundary data by means of appropriate vector fields in R+0×R+0, cf. [1]. With this approach time and space play the same role, the former being understood in particular as an additional state variable of the system.

    However, for the next purposes it is convenient to characterise the solution μ to (10) by means of the traces of μ1 and μ2 over the fibres R+0×{t}, t>0; i.e.

    μ(dxdt)=(μ1t(dx)+μ2t(dx))dt,

    where μ1t, μ2t are given by the transport of μ0, ν0, respectively, along the characteristics generated in R+×R+ by the velocity field v.

    In order to obtain a formula for μ1t, let Φt=Φt(x,0) be the position at time t>0 of the particle which is in xR+0 at time t=0 and which moves following the velocity field v=v(x):

    {ddtΦt(x,0)=v(Φt(x,0)),t>0Φ0(x,0)=x. (12)

    By standard results, it is well known that

    μ1t=Φt#μ0=R+0δΦt(x,0)dμ0(x)M+(R+0×{t}),

    where # is the push-forward operator, δ is the Dirac delta measure, and the integral at the right-hand side is understood in the sense of Bochner.

    Likewise, to obtain a formula for μ2t we consider the characteristic lines issuing from the t axis. In particular, we denote now by Φt(0,s) the position at time t>0 of the particle which is in x=0 at time sR+0 and which moves following the velocity field v=v(x):

    {ddtΦt(0,s)=v(Φt(0,s)),t>sΦs(0,s)=0. (13)

    By transporting the mass ν0 along these characteristics we can write

    μ2t=[0,t]δΦt(0,s)dν0(s)M+(R+0×{t}),

    where the integral is again meant in the sense of Bochner.

    Summing up, we consider the following representation formula for μ:

    μ(dxdt)=(R+0δΦt(ξ,0)(dx)dμ0(ξ)+[0,t]δΦt(0,s)(dx)dν0(s))dt (14)

    and we check that it actually defines a solution to (10) in the sense of Definition 4.1. To this purpose we preliminarily observe that, since μ1t=Φt#μ0, for every (bounded and measurable) function f:R+0R it results

    R+0f(x)dμ1t(x)=R+0f(Φt(x,0))dμ0(x). (15)

    We can obtain a similar formula for μ2t by observing that, given a simple function f:R+0R, f(x)=Nk=1αkχAk(x), where {Ak}Nk=1 is a measurable finite disjoint partition of R+0, it results

    R+0f(x)dμ2t(x)=Nk=1αkμ2t(Ak)=Nk=1αk[0,t]δΦt(0,s)(Ak)dν0(s)=Nk=1αk[0,t]χAk(Φt(0,s))dν0(s)=[0,t]Nk=1αkχAk(Φt(0,s))dν0(s)=[0,t]f(Φt(0,s))dν0(s).

    Approximating a measurable function f with a sequence of simple functions we get in general

    R+0f(x)dμ2t(x)=[0,t]f(Φt(0,s))dν0(s). (16)

    Interestingly, an integral with respect to the x variable is converted into one with respect to the t variable.

    Plugging (14) into the left-hand side of (11) and using (15), (16) we discover:

    μ,tφ+v(x)xφ=R+0R+0(tφ(Φt(x,0),t)+v(Φt(x,0))xφ(Φt(x,0),t))dμ0(x)dt=+R+0[0,t](tφ(Φt(0,s),t)+v(Φt(0,s))xφ(Φt(0,s),t))dν0(s)dt=R+0R+0ddtφ(Φt(x,0),t)dμ0(x)dt+R+0[0,t]ddtφ(Φt(0,s),t)dν0(s)dt,

    where in the last passage we have invoked (12), (13). By switching the order of integration in view of Fubini-Tonelli's Theorem we further obtain

    =R+0R+0ddtφ(Φt(x,0),t)dtdμ0(x)+R+0[s,+)ddtφ(Φt(0,s),t)dtdν0(s)=R+0[φ(Φt(x,0),t)]t=+t=0dμ0(x)+R+0[φ(Φt(0,s),t)]t=+t=sdν0(s)=R+0φ(x,0)dμ0(x)R+0φ(0,s)dν0(s)=μ0,φν0,φ,

    which confirms that (14) is indeed a measure-valued solution to (10). Uniqueness of such a solution is a consequence of continuous dependence estimates on the initial and boundary data, which can be proved by standard arguments in literature, cf. [1]. In conclusion, for the transport problem in R+×R+ we have the following well-posedness result:

    Theorem 4.2. For μ0M+(R+0×{0}), ν0M+({0}×R+0) there exists a unique measure-valued solution to (10) in the sense of Definition 4.1, which can be represented by (14).

    We now pass to consider the transport problem on the bounded domain Q:=(0,1)×(0,T), T>0, i.e.

    {tμ+x(v(x)μ)=0,(x,t)Qμt=0=μ0M+([0,1]×{0})νx=0=ν0M+({0}×[0,T]) (17)

    for a given bounded, strictly positive and Lipschitz continuous velocity field v:[0,1](0,vmax]. The solution to this problem can be obtained by restricting to Q the measure μ solving (10) (with the velocity field v possibly extended to the whole R+0 as, e.g. v(x)=v(1) for x1). Therefore we are going to consider the restriction of μ to Q defined as the measure μQ such that

    μQ(E):=μ(EQ)

    for every measurable set ER+0×R+0.

    In particular, in view of the application of this problem to a network, it is important to characterise the traces of μQ on the fibres [0,1]×{T} and {1}×[0,T], which depend on the transport of μ0 and ν0 inside Q.

    Let us introduce the following quantities:

    τ(x):=inf{t0:Φt(x,0)=1},x[0,1] (18)
    σ(s):=inf{ts:Φt(0,s)=1},s[0,T] (19)

    corresponding to the time needed to the characteristic line issuing from either (x,0), in case of τ(x), or (0,s), in case of σ(s), to hit the boundary x=1. Since the considered transport problem is linear, in particular the velocity field v does not depend on the measure μ itself, both τ and σ are one-to-one, thus invertible. Moreover τ decreases with x while σ increases with s and, in particular, σ(s)=τ(0)+s because v is autonomous.

    Recalling (14) and using τ, σ we write the trace of μQ on the fibre [0,1]×{T} as (cf. Figure 1)

    Figure 1.  Sketch of the characteristics of problem (17) in the two cases τ(0)=σ(0)<T (left) and τ(0)=σ(0)>T (right). For pictorial purposes we imagine a constant velocity field, so that the characteristics are straight lines in the space-time.
    μT:=[0,max{0,τ1(T)}]δΦT(x,0)dμ0(x)+[max{0,σ1(T)},T]δΦT(0,s)dν0(s) (20)

    whereas, following the characteristics, we construct the trace on the fibre {1}×[0,T] as

    ν1:=(max{0,τ1(T)},1]δτ(x)dμ0(x)+[0,max{0,σ1(T)})δσ(s)dν0(s). (21)

    We incidentally notice that the first term at the right-hand side of (20) is the push-forward of μ0 by the flow map ΦT then restricted to x[0,1] while the second term at the right-hand side of (21) is the push-forward of ν0 by the mapping σ then restricted to t[0,T].

    The relationship between these traces and the transport of μ0, ν0 inside Q is rigorously stated by the following theorem, which represents our main result on problem (17):

    Theorem 4.3. Given μ0M+([0,1]×{0}), ν0M+({0}×[0,T]), the measure μQM+(ˉQ), μM+(R+0×R+0) being the solution to (10), is the unique measure which satisfies the balance

    μQ,tφ+v(x)xφ=μTμ0,φ+ν1ν0,φ,φC1(ˉQ), (22)

    where μTM+([0,1]×{T}), ν1M+({1}×[0,T]) are the traces defined in (20), (21), respectively.

    Moreover, for μk0M+([0,1]×{0}), νk0M+({0}×[0,T]), k=1,2, there exists a constant C=C(T)>0 such that

    μ2Tμ1TBL+ν21ν11BLC(μ20μ10BL+ν20ν10BL). (23)

    Proof. See Appendix A.

    We also give a result about the dependence on time.

    Theorem 4.4. Given μ0M+([0,1]×{0}), ν0M+({0}×[0,T]), there exists a constant C=C(T)>0 such that

    μtμtBL+ν1[0,t]ν1[0,t]BLC|tt|+ν0([t,t]) (24)

    for all t,t[0,T] with t<t.

    Proof. See Appendix A.

    Remark 5. Theorem 4.4 states virtually that the traces μt and ν1[0,t] of μQ are Lipschitz continuous in time, a part from the presence of the term ν0([t,t]) in the estimate (24).

    If the boundary datum ν0 is absolutely continuous with respect to the Lebesgue measure in the interval [t,t] then for tt we get actually μtμtBL+ν1[0,t]ν1[0,t]BL0. If instead ν0 contains singularities in [t,t] then the distances μtμtBL, ν1[0,t]ν1[0,t]BL between two traces on horizontal and vertical fibres are in general not proportional to the time gap |tt|.

    In the applications, a Lebesgue-absolutely continuous ν0 corresponds to a macroscopic inflow mass provided with density. A Lebesgue-singular ν0 corresponds instead to microscopic point masses flowing from the boundary x=0 during the time interval [t,t] and then propagating as singularities across Q.

    In this section we go back to the study of problem (5). In order to make the notation consistent with the one introduced in Section 4, we set

    νj0:=μjVi=πj(0),νj1:=μjVi=πj(1)

    and we rewrite (5) as

    {tμj+x(vj(x)μj)=0xEj,t(0,T],jJμjt=0=μj0xEj,jJνj0={k:Vi=πk(1)pikj(t)νk1if iIpij(t)ςiif iS. (25)

    Let φC1(Γ×[0,T]). Given μj0M+(Ej×{0}), νj0M+({0}×[0,T]), owing to Theorem 4.3 there exists μjM+(Ej×[0,T]) such that

    μj,tφ+vj(x)xφ=μjTμj0,φ+νj1νj0,φ (26)

    for every jJ. Similarly to (20), (21), the traces μjT, νj1 are

    μjT=[0,max{0,τ1j(T)}]δΦjT(x,0)dμj0(x)+[max{0,σ1j(T)},T]δΦjT(0,s)dνj0(s) (27)
    νj1=(max{0,τ1j(T)},1]δτj(x)dμj0(x)+[0,max{0,σ1j(T)})δσj(s)dνj0(s), (28)

    where the flow maps Φjt(x,0) and Φjt(0,s) are defined like in (12), (13), respectively, using the velocity field vj(x) on the arc Ej, jJ. Likewise, τj and σj are defined like in (18), (19).

    Summing (26) over j and recalling (6) we deduce

    μ,tφ+v(x)xφ=μTμ0,φ+jJνj1νj0,φ, (29)

    where

    μ0=jJμj0,μT=jJμjT. (30)

    In particular, the last term at the right-hand side in (29) can be rewritten in more detail by summing on the vertexes of the network:

    jJνj1νj0,φ=iI(j:Vi=πj(1)νj1,φj:Vi=πj(0)νj0,φ)=iI(j:Vi=πj(1)νj1,φj:Vi=πj(0)νj0,φ)=+iWj:Vi=πj(1)νj1,φiSj:Vi=πj(0)νj0,φ.

    For an internal vertex Vi, iI, using the corresponding boundary condition prescribed in (25) we obtain:

    j:Vi=πj(1)νj1,φj:Vi=πj(0)νj0,φ=j:Vi=πj(1)νj1,φ=j:Vi=πj(0)k:Vi=πk(1)pikj(t)νk1,φ=j:Vi=πj(1)νj1,φ=k:Vi=πk(1)j:Vi=πj(0)pikj(t)νk1,φ

    whence, taking (3) into account in the second term at the right-hand side,

    =j:Vi=πj(1)νj1,φk:Vi=πk(1)νk1,φ=0.

    This is the conservation of the mass through the internal vertexes of the network.

    For a source vertex Vi, iS, we use the corresponding boundary condition prescribed in (25) to find:

    iSj:Vi=πj(0)νj0,φ=iSj:Vi=πj(0)pij(t)ςi,φ=iS(j:Vi=πj(0)pij(t))ςi,φ

    whence, in view of (4),

    =iSςi,φ=ς,φ

    where we have defined the measure ς:=iSςiM+(iS{Vi}×[0,T]). This is the total mass flowing into the network from the source vertexes up to the time T.

    Finally, for a well vertex Vi, iW, we define

    ωi:=j:Vi=πj(1)νj1M+({Vi}×[0,T]),ω:=iWωiM+(iW{Vi}×[0,T]), (31)

    which represents the total mass flowing out of the network up to the time T.

    Equation (29) takes then the form

    μ,tφ+v(x)xφ=μTμ0,φ+ως,φ,φC1(Γ×[0,T]), (32)

    thereby expressing the counterpart of (22) on the network.

    Using the formulation just obtained, we are in a position to establish the well-posedness of the transport problem over networks.

    Theorem 5.1. Given μ0M+(Γ×{0}) and ςM+(iS{Vi}×[0,T]), there exists a unique measure μM+(Γ×[0,T]) which satisfies the balance (32) with μTM+(Γ×{T}) defined in (27)-(30) and ωM+(iW{Vi}×[0,T]) defined in (28)-(31).

    Moreover, for μ0,kM+(Γ×{0}), ςkM+(iS{Vi}×[0,T]), k=1,2, there exists a constant C=C(T)>0 such that

    μT,2μT,1BL+ω2ω1BLC(μ0,2μ0,1BL+ς2ς1BL). (33)

    Proof. We treat separately the cases in which the set of the source vertexes is or is not empty.

    (ⅰ) Assume S. We introduce a partition of the set E={Ej}jJ based on the distance from the source set:

    E0={Ej:Vi=πj(0) is a source}Em={Ej:EkEm1 s.t. Vi=πj(0)=πk(1)},m=1,2,

    We first apply Theorem 4.3 to the problem defined on each arc in E0, i.e for each EjE0 such that Vi=πj(0), iS, we consider

    {tμj+x(vj(x)μj)=0in Ej×(0,T]μjt=0=μj0M+(Ej×{0})νj0=pij(t)ςiM+({Vi}×[0,T]).

    Since νj0 is prescribed, we obtain the existence of μjM+(Ej×[0,T]), μjTM+(Ej×{T}) and νj1M+({πj(1)}×[0,T]) satisfying the balance (22). Next we proceed by induction on m=1,2, considering the problem on EjEm with Vi=πj(0):

    {tμj+x(vj(x)μj)=0in Ej×(0,T]μjt=0=μj0M+(Ej×{0})νj0=diIk=1pikj(t)νk1M+({Vi}×[0,T]).

    Since the arcs Ek, k=1,,diI, belong to Em1, the solution to the transport equation on them is known by the inductive step (using the case m=0 as basis), hence the boundary measure νj0 is well defined because so are the outflow measures νk1. Therefore we can apply again Theorem 4.3 to fulfil the balance (22) on EjEm.

    In this way, after a finite number of steps we build arc by arc the measures μM+(Γ×[0,T]), μTM+(Γ×{T}) and ωM+(iW{Vi}×[0,T]) which globally satisfy the balance (32).

    (ⅱ) Assume now S=. Fix an arbitrary internal vertex Vi, iI, and choose

    t0<minjJ:Vi=πj(1)τj(0).

    From (28) we see that, up to the time t0, on all the arcs Ej such that Vi=πj(1) the outflow measure νj1 is given by

    νj1=(τ1j(t0),1]δτj(x)dμj0(x),

    because τ1j(t0)>0 while σ1j(t0)<0 (cf. Figure 1, left). Hence νj1 depends only on the initial datum μj0 and not on the inflow measure νj0.

    Let us consider the initial/boundary-value problem (25) for t(0,t0] with Vi as source vertex and corresponding source measure

    ςi=j:Vi=πj(1)νj1=j:Vi=πj(1)(τ1j(t0),1]δτj(x)dμj0(x).

    From the case S we know that we can construct μM+(Ej×[0,t0]), μt0M(Ej×{t0}) and ωM+(iW{Vi}×[0,t0]) which satisfy the balance (22). Moreover, the inflow measures νj0 of all the arcs Ej such that Vi=πj(0) coincide with those of the original problem without sources, because they are actually determined only by the initial datum. Hence μ is also a solution of the original problem in [0,t0]. By repeating this argument on the intervals (t0,2t0], (2t0,3t0], …, with initial data μt0, μ2t0, …, after a finite number of steps we obtain the solution of the problem without source in any interval [0,T], T>0.

    Finally, the estimate (33) is in both cases an immediate consequence of the corresponding estimate (23) holding on each arc.

    In this section we write explicitly the solution to problem (25) for two typical junctions which occur frequently for instance in traffic flow on road networks. It is worth pointing out that, since in our linear equation the velocity depends only on the space variable but not on the measure μ itself, the transport model that we are considering may provide an acceptable description of the flow of vehicles at most in the so-called free flow regime. In fact, in such a case the number of vehicles is sufficiently small that their speed is almost independent of the presence of other vehicles on the road.

    Let Γ be the road network shown in Figure 2 formed by 3 arcs, viz. roads, E1,E2,E3 and 4 vertexes V1,,V4 such that connects the source vertex to the internal vertex while and connect the internal vertex to the well vertexes and . This gives also the orientation of the arcs. In practice, beyond the junction the road splits in the two roads , . We assume that the network is initially empty. At some time a microscopic vehicle enters the network from the vertex and then travels across it. At the junction we prescribe a flux distribution rule stating that a time-dependent fraction of the incoming mass flows to the road while the complementary fraction flows to the road . Taking , the problem can be formalised as:

    Figure 2.  The 1-2 junction with a sketch of the characteristics along which the solution to the example of Section 6.1 propagates in the space-time of the network.

    where the velocity fields , , are given Lipschitz continuous functions of .

    The solution on each road has the form , where is the trace of on the fibre . Using (27), (28) we determine explicitly the expression of for all and that of the outflow masses on the fibres (notice that , , ). We find (cf. Figure 2):

    Furthermore, using Bochner integrals in the product space we can possibly write the solution on each road as

    Remark 6. By carefully inspecting the expressions of , , we see that the unit-mass Dirac delta prescribed at the source vertex splits in two Dirac deltas beyond the junction , cf. also Figure 2, each of which carries a fraction, and , respectively, of the initial mass.

    Unlike the Dirac delta entering the road from , the two Dirac deltas propagating in the roads , do not represent physical microscopic vehicles. Rather, each of them is the same microscopic vehicle coming from the road and the coefficients , have to be understood as the probabilities that such a vehicle takes either outgoing road beyond the junction .

    This approach differs from the one proposed in [8], which instead assigns a path to each microscopic vehicle through the network in the spirit of the multipath traffic model introduced in [3,4].

    We now consider the same network as in the previous Section 6.1 but we prescribe an inflow measure which is absolutely continuous with respect to the Lebesgue measure:

    where with is the density of the vehicles entering the network from the vertex .

    Recalling that the network is initially empty and using (27), we obtain that for each the trace of the solution in the road is

    where in the last passage we have set after observing from (13) that for all . Likewise, recalling (28) we find that the outflow mass at the vertex is

    where in the second passage we have set . In view of the Bochner representation (1) and considering that , we deduce in particular

    According to our transmission conditions, this mass is distributed to the outgoing roads , as

    which, owing to (27), implies that the traces , of the solutions , in the outgoing roads are respectively given by

    and by

    It is interesting to note that, since in general the density is split asymmetrically in the roads and (unless ), the corresponding measure solution, even if possibly continuous inside the arcs of the network, is discontinuous across the vertex .

    Finally, the outflow masses and are recovered from (28) as

    and

    Observing that for , from the Bochner representation (1) of a measure we further deduce

    Remark 7. The transport problem being linear, the case of an inflow measure carrying both an atomic and a Lebesgue-absolutely continuous part can be addressed by simply superimposing the solutions obtained in Sections 6.1 and 6.2.

    We consider now the road network illustrated in Figure 3 with again 3 arcs, viz. roads, and vertexes . However, in this case both vertexes , are sources and are connected by roads , to the internal vertex . The latter is finally connected to the well vertex by road . In practice, beyond the junction the incoming roads , merge into the outgoing road .

    Figure 3.  The 2-1 junction with a sketch of the characteristics along which the solution to the example of Section 6.3 propagates in the space-time of the network.

    Like in Sections 6.1, 6.2, we assume that the network is initially empty. At two successive time instants two microscopic vehicles enter the network from the sources , , respectively. Their propagation across the network for is then described by the problem:

    where the velocity fields , , are as usual given Lipschitz continuous functions of . Notice that, for mass conservation purposes, the flux distribution coefficients at the junction are necessarily for all .

    Relying again on (27), (28) we write explicitly the solution on each road as well as the outflow measures , with and . We find (cf. Figure 3):

    whence, using Bochner integrals in the product spaces , ,

    Proof of Theorem 4.3 We observe that can be obtained, by linearity, as the sum of the solutions of two transport problems with and , respectively.

    We begin by considering the case and assume, without loss of generality, that . Then whence, recalling (20), (21), we obtain

    (34)

    and we have to show that

    (35)

    where is the measure (14). Following the characteristics, its restriction to writes as

    thus for we discover:

    where in the last passage we have used (12). Switching the order of integration, we continue the calculation as:

    From (34) we recognise that the term (ⅰ) is indeed and that the term (ⅱ) is , while the term (ⅲ) is clearly . Consequently (35) follows.

    We consider now the case and assume, without loss of generality, that . Then whence, recalling again (20), (21), we find

    (36)

    and we have to show that

    (37)

    where is again the measure (14). Following the characteristics we see that is now expressed as

    hence for we obtain:

    where in the last passage we have used (13). We now switch the order of integration to discover:

    Thanks to (36) we recognise that the term (ⅰ) is and that the term (ⅱ) is , while the term (ⅲ) is clearly . Hence (37) follows.

    To conclude the proof, we show the continuous dependence estimate (23). We consider two problems of the type (17) with respective initial data and source data .

    We begin by estimating the term . Let with . Recalling (20) we have:

    where here stands for the total variation of a measure. Thus

    and consequently, taking the supremum over at both sides,

    Proceeding in a similar way for , from (21) we have:

    hence, taking the supremum over at both sides,

    Summing the two estimates just obtained yields finally (23).

    Moreover, for , the estimate (23) implies , , hence the uniqueness of (20) and (21).

    Proof of Theorem 4.4. We begin with the estimate of . Let be such that . By (20), since

    we can write:

    Likewise, assuming for simplicity that ,

    Hence

    and finally, taking the supremum over at both sides,

    We now consider the estimate on the outflow measures. Taking again with , we compute:

    We point out that if then the interval is actually understood as and, in this case, . Moreover, since we have , which implies . Then

    Moreover,

    which gives

    Therefore

    whence, taking the supremum over at both sides,

    Summing the estimates obtained so far for , we finally get (24).

    A.T. is member of GNFM (Gruppo Nazionale per la Fisica Matematica) of INdAM (Istituto Nazionale di Alta Matematica), Italy.

    A.T. acknowledges that this work has been written within the activities of a research project funded by "Compagnia di San Paolo" (Turin, Italy).



    [1] Binary interaction algorithms for the simulation of flocking and swarming dynamics. Multiscale Model. Simul. (2013) 11: 1-29.
    [2] Some kinetic models for a market economy. Boll. Unione Mat. Ital. (2017) 10: 143-158.
    [3] Kinetic models of conservative economies with wealth redistribution. Commun. Math. Sci. (2009) 7: 901-916.
    [4] On the self-similar asymptotics for generalized nonlinear kinetic Maxwell models. Comm. Math. Phys. (2009) 291: 599-644.
    [5] Kinetic modeling of economic games with large number of participants. Kinet. Relat. Models (2011) 4: 169-185.
    [6] Kinetic models for goods exchange in a multi-agent market. Phys. A (2018) 499: 362-375.
    [7] Particle based gPC methods for mean-field models of swarming with uncertainty. Commun. Comput. Phys. (2019) 25: 508-531.
    [8] On a kinetic model for a simple market economy. J. Stat. Phys. (2005) 120: 253-277.
    [9] Numerical methods for kinetic equations. Acta Numer. (2014) 23: 369-520.
    [10] Kinetic modeling of alcohol consumption. J. Stat. Phys. (2019) 177: 1022-1042.
    [11]

    B. Düring, L. Pareschi and G. Toscani, Kinetic models for optimal control of wealth inequalities, Eur. Phys. J. B, 91 (2018), 12pp.

    [12] Scaling solutions of inelastic Boltzmann equations with over-populated high energy tails. J. Statist. Phys. (2002) 109: 407-432.
    [13] Statistical equilibrium in simple exchange games. Ⅱ. The redistribution game. Eur. Phys. J. B (2007) 60: 241-246.
    [14] Taxes in a wealth distribution model by inelastically scattering of particles. Interdisciplinary Description Complex Syst. (2009) 7: 1-7.
    [15] Pareto tails in socio-economic phenomena: A kinetic description. Economics (2018) 12: 1-17.
    [16] Human behavior and lognormal distribution. A kinetic description. Math. Models Methods Appl. Sci. (2019) 29: 717-753.
    [17] Online social networking and addiction – A review of the physchological literature. Int. J. Environ. Res. Public Health (2011) 8: 3528-3552.
    [18] (2013) Interacting Multiagent Systems: Kinetic Equations and Monte Carlo Methods.Oxford University Press.
    [19] Structure preserving schemes for nonlinear Fokker-Planck equations and applications. J. Sci. Comput. (2018) 74: 1575-1600.
    [20] Inelastically scattering particles and wealth distribution in an open economy. Phys. Rev. E (2004) 69: 1-7.
    [21] Kinetic models of opinion formation. Commun. Math. Sci. (2006) 4: 481-496.
    [22]

    G. Toscani, Wealth redistribution in conservative linear kinetic models, Europhys. Lett. (EPL), 88 (2009).

    [23] Kinetic models for the trading of goods. J. Stat. Phys. (2013) 151: 549-566.
    [24] Opinion modeling on social media and marketing aspects. Phys. Rev. E (2018) 98: 1-15.
    [25] Multiple-interaction kinetic modeling of a virtual-item gambling economy. Phys. Rev. E (2019) 100: 1-16.
    [26] Behavior analysis of virtual-item gambling. Phys. Rev. E (2018) 98: 1-12.
    [27]

    X. Wang and M. Pleimling, Online gambling of pure chance: Wager distribution, risk attitude, and anomalous diffusion, Sci. Rep., 9 (2019).

  • This article has been cited by:

    1. Giulia Cavagnari, Antonio Marigonda, Benedetto Piccoli, Generalized dynamic programming principle and sparse mean-field control problems, 2020, 481, 0022247X, 123437, 10.1016/j.jmaa.2019.123437
    2. Fabio Camilli, Raul De Maio, Andrea Tosin, Measure-valued solutions to nonlocal transport equations on networks, 2018, 264, 00220396, 7213, 10.1016/j.jde.2018.02.015
    3. Fabio Bagagiolo, Rosario Maggistro, Raffaele Pesenti, Origin-to-destination network flow with path preferences and velocity controls: A mean field game-like approach, 2021, 8, 2164-6066, 359, 10.3934/jdg.2021007
    4. Joseph Andria, Rosario Maggistro, Raffaele Pesenti, Sustainable Management of Tourist Flow Networks: A Mean Field Model, 2023, 196, 0022-3239, 730, 10.1007/s10957-023-02158-8
    5. Fabio Bagagiolo, Silvia Faggian, Rosario Maggistro, Raffaele Pesenti, Optimal Control of the Mean Field Equilibrium for a Pedestrian Tourists’ Flow Model, 2022, 22, 1566-113X, 243, 10.1007/s11067-019-09475-4
    6. SIMONE CACACE, FABIO CAMILLI, RAUL DE MAIO, ANDREA TOSIN, A measure theoretic approach to traffic flow optimisation on networks, 2019, 30, 0956-7925, 1187, 10.1017/S0956792518000621
    7. Diogo A. Gomes, Diego Marcon, Fatimah Al Saleh, 2019, The current method for stationary mean-field games on networks, 978-1-7281-1398-2, 305, 10.1109/CDC40024.2019.9029982
    8. Zlatinka Dimitrova, Flows of Substances in Networks and Network Channels: Selected Results and Applications, 2022, 24, 1099-4300, 1485, 10.3390/e24101485
    9. Fabio Bagagiolo, Raffaele Pesenti, 2017, Chapter 1, 978-3-319-70618-4, 3, 10.1007/978-3-319-70619-1_1
    10. Fabio Camilli, Giulia Cavagnari, Raul De Maio, Benedetto Piccoli, Superposition principle and schemes for measure differential equations, 2021, 14, 1937-5077, 89, 10.3934/krm.2020050
    11. Christian Düll, Piotr Gwiazda, Anna Marciniak-Czochra, Jakub Skrzeczkowski, Measure differential equation with a nonlinear growth/decay term, 2023, 73, 14681218, 103917, 10.1016/j.nonrwa.2023.103917
    12. Antonio Siconolfi, Alfonso Sorrentino, Aubry–Mather theory on graphs, 2023, 36, 0951-7715, 5819, 10.1088/1361-6544/acf6ef
  • Reader Comments
  • © 2020 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(1925) PDF downloads(301) Cited by(12)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog