Transport of measures on networks

  • Received: 01 October 2016 Revised: 01 February 2017
  • 35R02, 35Q35, 28A50

  • In this paper we formulate a theory of measure-valued linear transport equations on networks. The building block of our approach is the initial and boundary-value problem for the measure-valued linear transport equation on a bounded interval, which is the prototype of an arc of the network. For this problem we give an explicit representation formula of the solution, which also considers the total mass flowing out of the interval. Then we construct the global solution on the network by gluing all the measure-valued solutions on the arcs by means of appropriate distribution rules at the vertexes. The measure-valued approach makes our framework suitable to deal with multiscale flows on networks, with the microscopic and macroscopic phases represented by Lebesgue-singular and Lebesgue-absolutely continuous measures, respectively, in time and space.

    Citation: Fabio Camilli, Raul De Maio, Andrea Tosin. Transport of measures on networks[J]. Networks and Heterogeneous Media, 2017, 12(2): 191-215. doi: 10.3934/nhm.2017008

    Related Papers:

    [1] Markus Schlarb . A multi-parameter family of metrics on stiefel manifolds and applications. Journal of Geometric Mechanics, 2023, 15(1): 147-187. doi: 10.3934/jgm.2023008
    [2] Valentin Duruisseaux, Melvin Leok . Time-adaptive Lagrangian variational integrators for accelerated optimization. Journal of Geometric Mechanics, 2023, 15(1): 224-255. doi: 10.3934/jgm.2023010
    [3] Jacob R. Goodman . Local minimizers for variational obstacle avoidance on Riemannian manifolds. Journal of Geometric Mechanics, 2023, 15(1): 59-72. doi: 10.3934/jgm.2023003
    [4] Jordi Gaset, Arnau Mas . A variational derivation of the field equations of an action-dependent Einstein-Hilbert Lagrangian. Journal of Geometric Mechanics, 2023, 15(1): 357-374. doi: 10.3934/jgm.2023014
    [5] Francesco Bonechi, Jian Qiu, Marco Tarlini . Generalised Kähler structure on CP2 and elliptic functions. Journal of Geometric Mechanics, 2023, 15(1): 188-223. doi: 10.3934/jgm.2023009
    [6] William Clark, Anthony Bloch . Existence of invariant volumes in nonholonomic systems subject to nonlinear constraints. Journal of Geometric Mechanics, 2023, 15(1): 256-286. doi: 10.3934/jgm.2023011
    [7] Maulik Bhatt, Amit K. Sanyal, Srikant Sukumar . Asymptotically stable optimal multi-rate rigid body attitude estimation based on lagrange-d'alembert principle. Journal of Geometric Mechanics, 2023, 15(1): 73-97. doi: 10.3934/jgm.2023004
    [8] Marcin Zając . The dressing field method in gauge theories - geometric approach. Journal of Geometric Mechanics, 2023, 15(1): 128-146. doi: 10.3934/jgm.2023007
    [9] Robert I McLachlan, Christian Offen . Backward error analysis for conjugate symplectic methods. Journal of Geometric Mechanics, 2023, 15(1): 98-115. doi: 10.3934/jgm.2023005
    [10] Xavier Rivas, Daniel Torres . Lagrangian–Hamiltonian formalism for cocontact systems. Journal of Geometric Mechanics, 2023, 15(1): 1-26. doi: 10.3934/jgm.2023001
  • In this paper we formulate a theory of measure-valued linear transport equations on networks. The building block of our approach is the initial and boundary-value problem for the measure-valued linear transport equation on a bounded interval, which is the prototype of an arc of the network. For this problem we give an explicit representation formula of the solution, which also considers the total mass flowing out of the interval. Then we construct the global solution on the network by gluing all the measure-valued solutions on the arcs by means of appropriate distribution rules at the vertexes. The measure-valued approach makes our framework suitable to deal with multiscale flows on networks, with the microscopic and macroscopic phases represented by Lebesgue-singular and Lebesgue-absolutely continuous measures, respectively, in time and space.



    The warped-product manifolds are type of manifolds introduced by Bishop and O'Neill [1]. These manifolds have become very important in the context of differential geometry and are also extensively studied in the arena of General Relativity, for instance with respect to generalized Friedmann-Robrtson-Walker spacetimes. Many properties for warped product manifolds and submanifolds were presented by B.-Y. Chen in [2].

    A warped-product manifold can be constructed as follows. Let (B,gB) and (F,gF) be two semi-Riemannian manifolds and τ, σ be the projection of B×F onto B and F, respectively.

    The warped-product M=B×fF is the manifold B×F equipped with the metric tensor g=τgB+f2σgF, where denotes the pullback and f is a positive smooth function on B, the so-called warping function.

    Explicitly, if X is tangent to B×F at (p,q) (where p is a point on B and q is a point on F), then:

    X,X=dτ(X),dτ(X)+f2(p)(dσ(X),dσ(X)).

    B is called the base-manifold of M=B×fF and F is the fiber-manifold. If f=1, then B×fF reduces to a semi-Riemannian product manifold. The leaves B×q=σ1(q) and the fibers p×F=τ1(p) are Riemannian submanifolds of M. Vectors tangent to leaves are called horizontal and those tangent to fibers are called vertical. ByH we denote the orthogonal projection of T(p,q)M onto its horizontal subspace T(p,q)(B×q) and V denotes the projection onto the vertical subspace T(p,q)(p×F), see [3].

    If M is an n-dimensional manifold, and gM is its metric tensor, the Einstein condition means that RicM=λgM for some constant λ, where RicM denotes the Ricci tensor of gM. An Einstein manifold with λ=0 is called Ricci-flat manifolds.

    Then keeping this in mind, we get that a warped-product manifold (M,gM)=(B,gB)×f(F,gF) (where (B,gB) is the base-manifold, (F,gF) is the fiber-manifold), with gM=gB+f2gF, is Einstein if only if (see [2]):

    RicM=λgM{RicBdfHess(f)=λgBRicF=μgFfΔf+(d1)|f|2+λf2=μ (1.1)

    where λ and μ are constants, d is the dimension of F, Hess(f), Δf and f are, respectively, the Hessian, the Laplacian (given by trHess(f)) and the gradient of f for gB, with f:(B)R+ a smooth positive function. Contracting first equation of (1.1) we get:

    RBf2fΔfd=nf2λ (1.2)

    where n and RB is the dimension and the scalar curvature of B respectively. From third equation, considering d0 and d1, we have:

    fΔfd+d(d1)|f|2+λf2d=μd (1.3)

    Now from (1.2) and (1.3) we obtain:

    |f|2+[λ(dn)+RBd(d1)]f2=μ(d1). (1.4)

    In 2017 de Sousa and Pina [4], studied warped-product semi-Riemannian Einstein manifolds in case that base-manifold is conformal to an n-dimensional pseudo-Euclidean space and invariant under the action of an (n1)-dimensional group with Ricci-flat fiber F. In [5] the authors extend the work done for multiply warped space. In [6], the author introduced a new type of warped-products called sequential warped-products, i.e. (M,gM) where M=(B1×hB2)×fF and gM=(gB1+h2gB2)+f2gF, to cover a wider variety of exact solutions to Einstein's field equation.

    Regarding the sequential warped-product manifolds, some works have been published in recent years ([7,8,9,10,11,12]).

    The main aim of the present paper is largely to continue to extend the work done in [4] (as was done for the multiply warped-product manifold in [5]), also for a special case of sequential warped-product manifolds, (i.e. for h=1, with B2 as an Einstein manifold, and flat fiber F, where the base-manifold B=B1×B2 is the product of two manifolds both equipped with a conformal metrics, and the warping function is a smooth positive function f(x,y)=f1(x)+f2(y) where each is a function on its individual manifold). The method will be as follows: first deriving the general formulas to be Einstein and second, providing the existence of solutions that are invariant under the action of a (n11)-dimensional group of transformations to the case of positive constant Ricci curvature. In fact, since in both references, [4] and [5], the authors show solutions for the Ricci-flat case (λ=0), we, following their same construction, show the existence of a family solutions for constant positive Ricci curvature (λ>0). In particular, this proof of the existence of a family of solutions also holds for [4] considering dimF=dimB.

    Definition 1.1. We consider the special case of the Einstein sequential warped-product manifold, that satisfies (1.1). The manifold (M,gM) comprises the base-manifold (B,gB) which is a Riemannian (or pseudo-Riemannian) product-manifold B=B1×B2, with B2 as an Einstein manifold (i.e., RicB2=λgB2, where λ is the same for (1.1) and gB2 is the metric for B2), and dim(B2)=n2, dim(B1)=n1 the dimension of B2 and B1, respectively, so that dim(B)=n=n1+n2. The warping function f:BR+ is a smooth positive function f(x,y)=f1(x)+f2(y) (where each is a function on its individual manifold, i.e., f1:B1R+ and f2:B2R+). The fiber-manifold (F,gF) is the Rd, with orthogonal Cartesian coordinates such that gab=δab.

    Proposition 1.2. If we write the B-product as B=B1×B2, where:

    i) RicBi is the Ricci tensor of Bi referred to gBi, where i=1,2,

    ii) f(x,y)=f1(x)+f2(y), is the smooth warping function, where fi:BiR+,

    iii) Hess(f)=iτiHessi(fi) is the Hessian referred on its individual metric, where τi are the respective pullbacks, (and τ2Hess2(f2)=0 since B2 is Einstein),

    iv) f is the gradient (then |f|2=i|ifi|2), and

    v) Δf=iΔifi is the Laplacian, (from (iii) therefore also Δ2f2=0).

    Then the Ricci curvature tensor will be:

    {RicM(Xi,Xj)=RicB1(Xi,Xj)dfHess1(f1)(Xi,Xj)RicM(Yi,Yj)=RicB2(Yi,Yj)RicM(Ui,Uj)=RicF(Ui,Uj)gF(Ui,Uj)fRicM(Xi,Yj)=0RicM(Xi,Uj)=0,RicM(Yi,Uj)=0, (1.5)

    where f=Δ1f1f+(d1)|f|2f2, and Xi, Xj, Yi, Yj, Ui, Uj are vector fields on B1, B2 and F, respectively.

    Theorem 1.3. A warped-product manifold is a special case of an Einstein sequential warped-product manifold, as defined in Definition 1.1, if and only if:

    RicM=λgM{RicB1dfτ1Hess1(f1)=λgB1τ2Hess2(f2)=0RicB2=λgB2RicF=0fΔ1f1+(d1)|f|2+λf2=0, (1.6)

    (since RicB is the Ricci curvature of B referred to gB, then RicB=RicB1+RicB2=λ(gB1+gB2)+dfτ1Hess1(f1).

    Therefore from (1.2) and (1.3):

    RM=λ(n+d){RB1fΔ1f1d=n1fλΔ2f2=0RB2=λn2RF=0fΔ1f1+(d1)|f|2+λf2=0. (1.7)

    where n1 and R1 are the dimension and the scalar curvature of B1 referred to gB1, respectively.

    Proof. We applied the condition that the warped-product manifold of system (1.5) is Einstein.

    This particular type of Einstein sequential warped-product manifold, as per Definition 1.1, allows to cover a wider variety of exact solutions of Einstein's field equation, without complicating the calculations much, compared to the Einstein warped-product manifolds with Ricci-flat fiber (F,gF), also considered by the authors of [4].

    In this section we will consider a special type of sequential warped-product manifold (M,gM), as described in the previous section, but in which the base-manifold is the product of two manifolds both equipped with a conformal metrics. First we will show the general formulas for which such a manifold M is Einstein, then we will show the same in the case where the conformal metrics are both diagonal, and finally for the case in which the base-manifold is the product of two conformal manifolds to a n1-dimensional and n2-dimensional pseudo-Euclidean space, respectively.

    Theorem 2.1. Let (B,gB), be the base-manifold B=(B1×B2), B1=Rn1, with coordinates (x1,x2,..xn1), B2=Rn2, with coordinates (y1,y2,..yn2), where n1,n23, and let gB=gB1+gB2 be the metrics on B, where gB1=ϵiδij and gB2=ϵlδlr.

    Let f1:Rn1R, f2:Rn2R, ϕ1:Rn1R and ϕ2:Rn2R, be smooth functions, where f1 and f2 are positive functions, such that f=f1+f2 as in Definition 1.1. Finally, let (M,gM) be ((B1×B2)×f=f1+f2F,gM), with gM=ˉgB+(f1+f2)2gF, with conformal metric ˉgB=ˉgB1+ˉgB2, where ˉgB1=1ϕ21gB1, ˉgB2=1ϕ22gB2, and F=Rd with gF=δab.

    Then the warped-product metric gM=ˉgB+(f1+f2)2gF is Einstein with constant Ricci curvature λ if and only if, the functions f1, f2, ϕ1 and ϕ2 satisfy:

    (I) (n12)fϕ1,xixjϕ1f1,xixjdϕ1,xif1,xjdϕ1,xjf1,xid=0 for ij,

    (II) (n22)ϕ2,ylyr=0 for lr,

    (III) ϕ1[(n12)fϕ1,xixiϕ1f1,xixid2ϕ1,xif1,xid]+

    +ϵi[fϕ1n1k=1ϵkϕ1,xkxk(n11)fn1k=1ϵkϕ12,xk+ϕ1dn1k=1ϵkϕ1,xkf1,xk]=ϵiλf,

    (IV) ϕ2(n22)ϕ2,ylyl+ϵlϕ2n2s=1ϵsϕ2,ysys(n21)ϵln2s=1ϵsϕ22,ys=λϵl,

    (V) fϕ12n1k=1ϵkf1,xkxk+(n12)fϕ1n1k=1ϵkϕ1,xkf1,xk+

    (d1)(ϕ12n1k=1ϵkf12,xk+ϕ22n2s=1ϵsf22,ys)=λf2.

    Before proving Theorem 2.1, and showing the existence of a solution for λ>0, we want to deduce the formulas for generic diagonal conformal metrics gB1 and gB2.

    Based on this, we consider (B,gB), the base-manifold B=(B1×B2), with dim(B1)=n1, dim(B2)=n2, and gB=gB1+gB2. We also consider f1:Rn1R, f2:Rn2R, ϕ1:Rn1R and ϕ2:Rn2R, are smooth functions, where f1 and f2 are positive functions, such that f=f1+f2 as in Definition 1.1. And finally, we consider (M,gM) with ((B1×B2)×(f1+f2)F,gM), with gM=ˉgB+(f1+f2)2gF, with conformal metric ˉgB=ˉgB1+ˉgB2, where ˉgB1=1ϕ21gB1, ˉgB2=1ϕ22gB2, and F=Rd with gF=δab.

    From (1.6), considering the conformal metric on B1 and B2, it is easy to deduce that M is Einstein if and only if:

    RicˉB1=λˉgB1+dfHessˉ1(f1), (2.1)

    or equivalently

    RˉB1=λn1+dfΔˉ1(f1), (2.2)
    RicˉB2=λˉgB2, (2.3)

    or equivalently

    RˉB2=λn2, (2.4)
    0=λf2+fΔˉ1f1+(d1)[|ˉ1f1|2+|ˉ2f2|2]. (2.5)

    If we consider a generic diagonal metric, ˉgBij=ˉgB1ij+ˉgB2ij=ηij, and ηij=0 for ij, then M is Einstein if and only if (2.1), (2.3) (or equivalently (2.2), (2.4)), (2.5) and the following, are satisfied:

    RicˉB1=dfHessˉ1(f1),forij, (2.6)
    RicˉB2=0,forij. (2.7)

    Proof. (of Theorem 2.1): At this point we can calculate:

    RicˉB1=1ϕ21{(n12)ϕ1Hess1(ϕ1)+[ϕ1Δ1ϕ1(n11)|1ϕ1|2]gB1}, (2.8)
    RicˉB2=1ϕ22{(n22)ϕ2Hess2(ϕ2)+[ϕ2Δ2ϕ2(n21)|2ϕ22]gB2}, (2.9)

    so we can write:

    RicˉB1(Xi,Xj)=1ϕ21{(n12)ϕ1Hess1(ϕ1)(Xi,Xj)+[ϕ1Δ1ϕ1(n11)|1ϕ1|2]gB1(Xi,Xj)}, (2.10)
    RicˉB2(Yl,Yr)=1ϕ22{(n22)ϕ2Hess2(ϕ2)(Yl,Yr)+[ϕ2Δ2ϕ2(n21)|2ϕ2|2]gB2(Yl,Yr)}, (2.11)
    RicM(Xi,Xj)=RicˉB1(Xi,Xj)dfHessˉ1(f1)(Xi,Xj), (2.12)

    for what was stated in Proposition 1.2 we have:

    RicM(Yl,Yr)=RicˉB2(Yl,Yr), (2.13)

    and in the end

    RicM(Xi,Yj)=0. (2.14)
    RicM(Xi,Uj)=0. (2.15)
    RicM(Yi,Uj)=0. (2.16)

    Since RicF=0 we obtain:

    RicM(Ui,Uj)=gM(Ui,Uj)(Δˉ1f1f+(d1)gM(f,f)f2), (2.17)

    where, analogous to Proposition 1.2, we consider gM(f,f)=ˉgB1(f1,f1)+ˉgB2(f2,f2).

    Let ϕ1,xixj, ϕ1,xi, f1,xixj, f1,xi, ϕ2,ylyr, ϕ2,yl, f2,ylyr and f2,yl, be the second and the first order derivatives of ϕ1, ϕ2, f1 and f2, respectively, with respect to xixj and ylyr. Now we have:

    Hess1(ϕ1)(Xi,Xj)=ϕ1,xixj, (2.18)
    Δ1(ϕ1)=n1k=1ϵkϕ1,xkxk, (2.19)
    |1(ϕ1)|2=n1k=1ϵkϕ21,xk, (2.20)
    Hess2(ϕ2)(Yl,Yr)=ϕ2,ylyr, (2.21)
    Δ2(ϕ2)=n2s=1ϵsϕ2,ylyr (2.22)
    |2(ϕ2)|2=n2s=1ϵsϕ22,ys. (2.23)
    Hessˉ1(f1)(Xi,Xj)=f1,xixjkˉΓkijf1,xk, (2.24)

    where ˉΓkij=0, ˉΓiij=ϕ1,xjϕ1, ˉΓkii=ϵiϵkϕ1,xkϕ1 and ˉΓiii=ϕ1,xjϕ1, so (2.24) becomes:

    Hessˉ1(f1)(Xi,Xj)=f1,xixj+ϕ1,xjϕ1f1,xi+ϕ1,xiϕ1f1,xj, (2.25)

    for ij, and

    Hessˉ1(f1)(Xi,Xi)=f1,xixi+2ϕ1,xiϕ1f1,xiϵin1k=1ϵkϕ1,xkϕ1f1,xk. (2.26)

    Since Hessˉ2(f2)(Yl,Yr)=0, we get:

    Hessˉ2(f2)(Yl,Yr)=f2,ylyr+ϕ2,yrϕ2f2,yl+ϕ2,ylϕ2f2,yr=0, (2.27)

    for lr, and

    Hessˉ2(f2)(Yl,Yl)=f2,ylyl+2ϕ2,ylϕ2f2,ylϵln2s=1ϵsϕ2,ysϕ2f2,ys=0. (2.28)

    Then the Ricci tensors are:

    RicˉB1(Xi,Xj)=(n12)ϕ1,xixjϕ1, (2.29)

    for ij,

    RicˉB1(Xi,Xi)=(n12)ϕ1,xixi+ϵin1k=1ϵkϕ1,xkxkϕ1(n11)ϵin1k=1ϵkϕ21,xkϕ21, (2.30)
    RicˉB2(Yl,Yr)=(n22)ϕ2,ylyrϕ2, (2.31)

    for lr,

    RicˉB2(Yl,Yl)=(n22)ϕ2,ylyl+ϵln2s=1ϵsϕ2,ysysϕ2(n21)ϵln2s=1ϵsϕ22,ysϕ22. (2.32)

    Using (2.29) and (2.25) in the (2.12) and then using (2.30) and (2.26) in the (2.12) we obtain respectively:

    RicM(Xi,Xj)=(n12)ϕ1,xixjϕ1df[f1,xixj+ϕ1,xjϕ1f1,xi+ϕ1,xiϕ1f1,xj], (2.33)

    for ij,

    RicM(Xi,Xi)=(n12)ϕ1,xixi+ϵin1k=1ϵkϕ1,xkxkϕ1(n11)ϵin1k=1ϵkϕ21,xkϕ21+df[f1,xixi+2ϕ1,xiϕ1f1,xiϵin1k=1ϵkϕ1,xkϕ1f1,xk], (2.34)

    while, using (2.31) and (2.27) in the (2.13) and then using (2.32) and (2.28) in the (2.13) we obtain respectively:

    RicM(Yl,Yr)=(n22)ϕ2,ylyrϕ2, (2.35)

    for lr,

    RicM(Yl,Yl)=(n22)ϕ2,ylyl+ϵln2s=1ϵsϕ2,ysysϕ2(n21)ϵln2s=1ϵsϕ2,ysϕ22. (2.36)

    Now considering:

    RicF=0, (2.37)
    gM(Ui,Uj)=f2gF(Ui,Uj), (2.38)

    with f=f1+f2,

    Δˉ2(f2)=0 (2.39)
    Δˉ1(f1)=ϕ21n1k=1ϵkf1,xkxk(n12)ϕ1n1k=1ϵkϕ1,xkf1,xk, (2.40)
    gM(f,f)=ϕ21n1k=1ϵkf21,xk+ϕ22n2s=1ϵsf22,ys, (2.41)

    and by replacing them in (2.17):

    RicM(Ui,Uj)={fϕ21n1k=1ϵkf1,xkxk+(n12)fϕ1n1k=1ϵkϕ1,xkf1,xk+(d1)(ϕ21n1k=1ϵkf21,xk+ϕ22n2s=1ϵsf22,ys)}gF(Ui,Uj). (2.42)

    Using the equations (2.33), (2.34), (2.35), (2.36) and (2.42), it follows that (M,gM) is an Einstein manifold if and only if, the equations (I), (II), (III), (IV), (V) are satisfied.

    In this section we look for the existence of a solution to the positive constant Ricci curvature case (λ>0) when the base-manifold is the product of two conformal manifolds to a n1-dimensional and n2-dimensional pseudo-Euclidean space, respectively, invariant under the action of a (n11)-dimensional group of transformations and that the fiber F is flat.

    Theorem 3.1. Let (B,gB), be the base-manifold B=(B1×B2), B1=Rn1, with coordinates (x1,x2,..xn1), B2=Rn2, with coordinates (y1,y2,..yn2), where n1,n23, and let gB=gB1+gB2 be the metrics on B, where gB1=ϵiδij and gB2=ϵlδlr.

    Let f1:Rn1R, f2:Rn2R, ϕ1:Rn1R and ϕ2:Rn2R, be smooth functions f1(ξ1), f2(ξ2), ϕ1(ξ2) and ϕ2(ξ2), such that f(ξ1,ξ2)=f1(ξ1)+f2(ξ2) be as in Definition 1.1, where ξ1=n1i=1αixi, αiR, and iϵiα2i=ϵi0 or iϵiα2i=0, and by the same token ξ2=n2l=1αlyl, αlR, and lϵlα2l=ϵl0 or lϵlα2l=0.

    Finally, let (M,gM) be ((B1×B2)×f=f1+f2F,gM), with gM=ˉgB+(f1+f2)2gF, with conformal metric ˉgB=ˉgB1+ˉgB2, where ˉgB1=1ϕ21gB1, ˉgB2=1ϕ22gB2, and F=Rd with gF=δab.

    Then, whenever iϵiα2i=ϵi0 (and lϵlα2l=ϵl0), the warped-product metric

    gM=ˉgB+(f1+f2)2gF is Einstein with constant Ricci curvature λ if and only if the functions f1, f2, ϕ1 and ϕ2 satisfy the following conditions:

    (Ia) (n12)fϕ1ϕ1f1d2ϕ1f1d=0, for ij,

    (IIa) ϕ2=0, for lr,

    (IIIa) kϵkα2k[fϕ1ϕ1(n11)fϕ21+ϕ1ϕ1f1d]=λf,

    (IVa) sϵsα2s[(n21)ϕ22]=λ

    (Va) kϵkα2k[fϕ21f1+(n12)fϕ1ϕ1f1(d1)ϕ21f21]+

    sϵsα2s[(d1)ϕ22f22]=λf2.

    Proof. We have:

    ϕ1,xixj=ϕ1αiαj, ϕ1,xi=ϕ1αi, f1,xixj=f1αiαj, f1,xi=f1αi,

    and

    ϕ2,ylyr=ϕ2αlαr, ϕ2,yl=ϕ2αl, f2,ylyr=f2αlαr, f2,yl=f2αl.

    Substituting these in (I) and (II) and if ij and lr such that αiαj0 and αlαr0, we obtain (Ia) and (IIa).

    In the same manner for (III) and (IV), by considering the relation between ϕ1 and f1 from (Ia) and ϕ2=0 from (IIa), we get (IIIa) and (IVa) respectively. Analogously, the equation (V) reduces to (Va).

    Now we are going to look for the existence of a solution to the positive constant Ricci curvature case (λ>0), considering f2(ξ2)=1, and dim(B1)=dim(F), i.e., n1=d. So, whenever n1i=1α2iϵi0, without loss of generality, we may consider n1i=1α2iϵi=1 (the same for n2l=1α2lϵl0, in which we consider n2l=1α2lϵl=1).

    In this way the equations (Ia), (IIa), (IIIa), (IVa) (Va) become:

    (Ib) (n12)(f1+1)ϕ1n1ϕ1f12n1ϕ1f1=0, for ij,

    (IIb) ϕ2=0, for lr,

    (IIIb) (f1+1)ϕ1ϕ1+(n11)(f1+1)ϕ21n1ϕ1ϕ1f1=λ(f1+1),

    (IVb) (n21)ϕ22=λ,

    (Vb) (f1+1)ϕ21f1(n12)(f1+1)ϕ1ϕ1f1+(n11)ϕ21f21=λ(f1+1)2.

    Note that since f2(ξ2)=constant, then the equations (2.27) and (2.28), concerning the condition Hessˉ2(f2)=0, are obviously satisfied.

    It is worth noticing that there is no reason to believe that any nontrivial solutions exist, since the system is overdetermined. One must first check out the compatibility conditions and fortunately this is easy to figure out. Changing the notation: from (ξ1,ϕ1(ξ1),f1(ξ1)), to (t,β(t),γ(t)1) (in order to simplify the writing and avoid confusion with the indexes), and also writing λ=qm2/2>0, where q=n1, i.e. dim(B1), our system of equations then becomes:

    {(q2)γβqβγ2qβγ=0βγβ(q1)γβ2qβγ12qm2γ=0γβ2γ(q2)βγβγ+(q1)β2γ212qm2γ2=0 (3.1)

    So, if we solve the second and third equations for β and γ and substituting them into the first equation, we note that the first equation can be replaced by a first order equation, that is:

    (q2)γ2β22qβγβγ+qβ2γ2qm2γ2=:Z(β,γ,β,γ)=0. (3.2)

    Now, differentiating Z with respect to t and then eliminating β and γ using the second and third equations of (3.1), the resulting expression in (β,γ,β,γ) is a multiple of Z(β,γ,β,γ). This shows us that the combined system of equations (3.1) and (3.2) satisfies the compatibility conditions, so that the system has solutions, specifically, a 3-parameter family of them.

    If we want to describe these solutions more explicitly, we must note that the equations are t-autonomous and have a 2-parameter family of scaling symmetries. In particular, the equations are invariant under the 3-parameter group of transformations of the form:

    Φa,b,c(t,β,γ)=(at+c,aβ,bγ) (3.3)

    where a and b are nonzero constants and c is any constant. In fact, the equation (3.2) implies that there is a function ω(t) such that

    {β=2mqω(ω1)((q2)ω22qω+q)γ=mγ((q2)ω2q)β((q2)ω22qω+q) (3.4)

    and then the second and third equations of (3.1) imply that ω must satisfy

    ω=m(q+2qω(3q2)ω2)β. (3.5)

    Conversely, the combined system of (3.4) and (3.5) gives the general solution of the original system. This latter system is easily integrated by the usual separation of variables method, i.e., by eliminating t yields a system of the form:

    dββ=R(ω)dω (3.6)

    and

    dγγ=S(ω)dω (3.7)

    where R(ω) and S(ω) are rational functions of ω. Writing β and γ as elementary functions of ω, then we can also write:

    dt=βT(ω)dω, (3.8)

    where T is a rational function of ω, so that t can be written as a function of ω by quadrature. Thus, we have the integral curves in (t,β,γ,ω)-space in terms of explicit functions.

    In conclusion (because of the 3-parameter family of equivalences of solutions), we can say that in certain sense, these solutions are all equivalent to a finite number of possibilities.

    Remark 3.2. As is well known, an Einstein warped product manifold with Riemannian-metric and Ricci-flat fiber-manifold can only admit zero or negative Ricci tensor, Ric0. Here we have shown, that a simple pseudo-Riemannian metric construction allows, an Einstein warped product manifold with Ricci-flat fiber-manifold, to obtain Ric>0, and this may find interest, for example, in how to build warped-product spacetime models, with positive curvature, whose fiber is Ricci-flat.

    All sources of funding of the study must be disclosed. The fourth author is thankful to Slovak Science Agency for providing partial financial by VEGA fund under grant number VEGA 2/0076/23.

    The authors declare there is no conflict of interest.

    [1] L. Ambrosio, N. Gigli and G. Savaré, Gradient Flows in Metric Spaces and in the Space of Probability Measures, Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 2008.
    [2] (2007) Measure Theory. Berlin Heidelberg: Springer-Verlag.
    [3] An easy-to-use algorithm for simulating traffic flow on networks: Numerical experiments. Discrete Contin. Dyn. Syst. Ser. S (2014) 7: 379-394.
    [4] An easy-to-use algorithm for simulating traffic flow on networks: Theoretical study. Netw. Heterog. Media (2014) 9: 519-552.
    [5] Stationary mean field games systems defined on networks. SIAM J. Control. Optim. (2016) 54: 1085-1103.
    [6] A well-posedness theory in measures for some kinetic models of collective motion. Math. Models Methods Appl. Sci. (2011) 21: 515-539.
    [7] E. Cristiani, B. Piccoli and A. Tosin, Multiscale Modeling of Pedestrian Dynamics, vol. 12 of MS & A: Modeling, Simulation and Applications, Springer International Publishing, 2014. doi: 10.1007/978-3-319-06620-2
    [8] On the micro-to-macro limit for first-order traffic flow models on networks. Netw. Heterog. Media (2016) 11: 395-413.
    [9] Modelling supply networks with partial differential equations. Quart. Appl. Math. (2009) 67: 419-440.
    [10] Vertex control of flows in networks. Netw. Heterog. Media (2008) 3: 709-722.
    [11] Mild solutions to a measure-valued mass evolution problem with flux boundary conditions. J. Differential Equations (2015) 259: 1068-1097.
    [12] Measure-valued mass evolution problems with flux boundary conditions and solution-dependent velocities. SIAM J. Math. Anal. (2016) 48: 1929-1953.
    [13] A fully-discrete-state kinetic theory approach to traffic flow on road networks. Math. Models Methods Appl. Sci. (2015) 25: 423-461.
    [14] M. Garavello and B. Piccoli, Traffic Flow on Networks -Conservation Laws Models, AIMS Series on Applied Mathematics, American Institute of Mathematical Sciences, Springfield, MO, 2006.
    [15] Models of discrete and continuous cell differentiation in the framework of transport equation. SIAM J. Math. Anal. (2012) 44: 1103-1133.
    [16] D. Mugnolo, Semigroup Methods for Evolution Equations on Networks, Understanding Complex Systems, Springer International Publishing, 2014. doi: 10.1007/978-3-319-04621-1
    [17] Generalized Wasserstein distance and its application to transport equations with source. Arch. Ration. Mech. Anal. (2014) 211: 335-358.
    [18] Differential equations on networks. J. Math. Sci. (N. Y.) (2004) 119: 691-718.
    [19] D. T. H. Worm, Semigroups on Spaces of Measures, PhD thesis, Leiden University, 2010.
  • Reader Comments
  • © 2017 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(4749) PDF downloads(187) Cited by(12)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog