The Lax-Oleinik semigroup on graphs

  • Received: 01 June 2017 Revised: 01 August 2017
  • Primary: 35R02, 35F21; Secondary: 35Q93

  • We consider Tonelli Lagrangians on a graph, define weak KAM solutions, which happen to be the fixed points of the Lax-Oleinik semi-group, and identify their uniqueness set as the Aubry set, giving a representation formula. Our main result is the long time convergence of the Lax Oleinik semi-group. It follows that weak KAM solutions are viscosity solutions of the Hamilton-Jacobi equation [3, 4], and in the case of Hamiltonians called of eikonal type in [3], we prove that the converse holds.

    Citation: Renato Iturriaga, Héctor Sánchez Morgado. The Lax-Oleinik semigroup on graphs[J]. Networks and Heterogeneous Media, 2017, 12(4): 643-662. doi: 10.3934/nhm.2017026

    Related Papers:

    [1] Renato Iturriaga, Héctor Sánchez Morgado . The Lax-Oleinik semigroup on graphs. Networks and Heterogeneous Media, 2017, 12(4): 643-662. doi: 10.3934/nhm.2017026
    [2] Jan Friedrich, Oliver Kolb, Simone Göttlich . A Godunov type scheme for a class of LWR traffic flow models with non-local flux. Networks and Heterogeneous Media, 2018, 13(4): 531-547. doi: 10.3934/nhm.2018024
    [3] John D. Towers . The Lax-Friedrichs scheme for interaction between the inviscid Burgers equation and multiple particles. Networks and Heterogeneous Media, 2020, 15(1): 143-169. doi: 10.3934/nhm.2020007
    [4] Raimund Bürger, Harold Deivi Contreras, Luis Miguel Villada . A Hilliges-Weidlich-type scheme for a one-dimensional scalar conservation law with nonlocal flux. Networks and Heterogeneous Media, 2023, 18(2): 664-693. doi: 10.3934/nhm.2023029
    [5] Anya Désilles . Viability approach to Hamilton-Jacobi-Moskowitz problem involving variable regulation parameters. Networks and Heterogeneous Media, 2013, 8(3): 707-726. doi: 10.3934/nhm.2013.8.707
    [6] Anya Désilles, Hélène Frankowska . Explicit construction of solutions to the Burgers equation with discontinuous initial-boundary conditions. Networks and Heterogeneous Media, 2013, 8(3): 727-744. doi: 10.3934/nhm.2013.8.727
    [7] Giuseppe Maria Coclite, Carlotta Donadello . Vanishing viscosity on a star-shaped graph under general transmission conditions at the node. Networks and Heterogeneous Media, 2020, 15(2): 197-213. doi: 10.3934/nhm.2020009
    [8] Christian Budde, Marjeta Kramar Fijavž . Bi-Continuous semigroups for flows on infinite networks. Networks and Heterogeneous Media, 2021, 16(4): 553-567. doi: 10.3934/nhm.2021017
    [9] Nurehemaiti Yiming . Dynamic analysis of the M/G/1 queueing system with multiple phases of operation. Networks and Heterogeneous Media, 2024, 19(3): 1231-1261. doi: 10.3934/nhm.2024053
    [10] Alberto Bressan, Khai T. Nguyen . Conservation law models for traffic flow on a network of roads. Networks and Heterogeneous Media, 2015, 10(2): 255-293. doi: 10.3934/nhm.2015.10.255
  • We consider Tonelli Lagrangians on a graph, define weak KAM solutions, which happen to be the fixed points of the Lax-Oleinik semi-group, and identify their uniqueness set as the Aubry set, giving a representation formula. Our main result is the long time convergence of the Lax Oleinik semi-group. It follows that weak KAM solutions are viscosity solutions of the Hamilton-Jacobi equation [3, 4], and in the case of Hamiltonians called of eikonal type in [3], we prove that the converse holds.



    There has been an increasing interest in the study of dynamical systems and differential equation on networks. In the same vein, the study of control problems on networks has interesting applications in various fields. A typical optimal control problem is the minimum time problem, which consists of finding the shortest path between an initial position and a given target set. If the cost changes in a continuous way along the edges and the dynamics is continuous in time, the minimum time problem can be seen as a continuous-state continuous-time control problem where the admissible trajectories of the system are constrained to remain on the network. Control problems with state constrained in closures of open sets have been intensively studied but the there is much fewer literature on problems on networks.

    Networks are the simplest examples of ramified spaces. Ramified spaces are the natural settings for problems of interaction between different media which are described by differential equations on the branches and transition conditions on the ramifications. Those interaction problems have different applications in physics, chemistry, and biology. Concerning the analysis of these models, the possibility of the application of several mathematical methods does not only depend on the structure of the differential equations on the branches, but also they depend strongly on the properties of the transition conditions. In fact, their effect on existence, uniqueness, and regularity of solutions is quite considerable. Many well-known results for elliptic and parabolic problems in the classical non-ramified situation have been extended to ramified spaces, providing results for boundary or initial value problems, now stablished for collections of media with transition conditions. In many cases, the core of the difficulties found on ramified spaces are already present in the simple setting of networks.

    Viscosity solutions of the Hamilton Jacobi equation on networks have been studied for several authors, but the problem of the long time behaviour of the solutions to the Cauchy problem for the Hamilton Jacobi equation has not been considered. Upon addressing that problem, we found more convenient to follow the variational approach, considering the viscosity solution given by the Lax Oleinik semigroup.

    In the first part of this article we study the Lax-Oleinik semi-group Lt defined by a Tonelli Lagrangian on a graph and prove that for any continuous function u, Ltu+ct converges as t where c is the critical value of the Lagrangian.

    For Lagrangians on compact manifolds, Fathi [6, 7] proved the convergence using the Euler-Lagrange flow and conservation of energy. In our case we do not have those tools but we can follow ideas of Roquejoffre [11] and Davini-Siconolfi [5]. See also [10] for the convergence when the manifold is the whole euclidian space.

    Camilli and collaborators [1, 4, 3] have studied viscosity solutions of the Hamilton-Jacobi equation, and given sufficient conditions for a set to be a uniqueness set and a representation formula. See also the recent papers [2, 9].

    In the second part of this article we prove that, under the assumption that the Lagrangian is symmetric at the vertices, the sets of weak KAM and viscosity solutions of the Hamilton-Jacobi equation coincide.

    We consider a graph G without boundary consisting of finite sets of unoriented edges I={Ij} and vertices V={el}. The interior of Ij is IjV. Parametrizing each edge by arc length σj:Ij[0,sj] we can write its tangent bundle as TIj=Ij×R and

    TG=j{j}×TIj/

    where (i,x,v)(j,y,w)(i,x,v)=(j,y,w) or x=yIiIj,v=w=0. Thus, a function L:TGR is given by a collection of functions Lj:TIjR such that Li(el,0)=Lj(el,0) for elIiIj. A Lagrangian in G is a function L:TGR such that each Lj is Ck, k2, and Lj(x,) is strictly convex and super-linear for any xIj. We will say that a Lagrangian is it symmetric at the vertices if at each vertex el there is a function λl:{uR:u0}R such that Lj(el,z)=λl(|z|) if elIj. As an example consider the mechanical Lagrangian given by Lj(x,v)=12v2Uj(x), with Uj(el)=al if elIj. For xIjV, we say that (x,v) points towards σ1(sj) if v>0 and points towards σ1(0) if v<0.

    We say that (σ1j(0),v) is an Ij-incoming or outgoing vector according to whether v>0 or v<0, and we say that (σ1j(sj),v) is an Ij-incoming or outgoing vector according to wether v<0 or v>0. We let T+elIj (TelIj) to be the set of Ij-outgoing (incoming or zero) vectors in TelIj.

    We start defining a distance in the most natural way. We say a continuous path α:[a,b]G is a unit speed geodesic (u.s.g.) if there is a partition a=t0<<tm=b such that for each 1im there is j(i) such that

    α([t0,t1])Ij(1),α([ti1,ti])=Ij(i),,α([tm1,tm])Ij(m),

    σj(i)α|[ti1,ti] is differentiable and either (σj(i)α)1 or (σj(i)α)1. We set the length of a u.s.g. to be

    (α)=|σj(1)(α(t1))σj(1)(α(a))|+m1i=2sj(i)+|σj(m)(α(b))σj(m)(α(tm1))|

    and define a distance on G by

    d(x,y)=min{(α):α:[a,b]G is a u.s.g.,α(a)=x,α(b)=y}

    We say a path γ:[a,b]G is absolutely continuous if for any ε>0 there is δ>0 such that for any finite collection of disjoint intervals {[ci,di]} with i(dici)<δ we have id(γ(ci),γ(di))<ε.

    If γ:[a,b]G, γ(t)V, we define ˙γ(t)=0 if for any ε>0 there is δ>0 such that d(γ(s),γ(t))<ε|st| when |st|<δ.

    Let γ:[a,b]G be absolutely continuous and consider the closed set V=γ1(V) so that (a,b)V=i(ai,bi) where the intervals (ai,bi) are disjoint and γ([ai,bi])Ij(i). It is clear that each σj(i)γ:[ai,bi][0,sj(i)] is absolutely continuous. We set ˙γ(t)=(σj(i)γ)(t) whenever is defined.

    Next Proposition will allow us to define the action of an absolutely continuous curve.

    Proposition 1. Let γ:[a,b]G be absolutely continuous and V=γ1(V)

    (a)˙γ=0 Lebesgue almost everywhere in V.

    (b)˙γ is integrable and for any [c,d][a,b] we have d(γ(c),γ(d))dc|˙γ|.

    proof. Write (a,b)V=i(ai,bi) as above with the intervals (ai,bi) disjoint. Since ˙γ=0 on the interior of V and {an,bn} is numerable to stablish item (a) it remains to prove that ˙γ=0 Lebesgue almost everywhere in V{an,bn}.

    Let ˉs=minjsj and take δ>0 such that d(γ(t1),γ(t2))<ˉs if |t1t2|<δ. There is N such that biai<δ for i>N. Since γ(ai),γ(bi)V we have that γ(ai)=γ(bi) for i>N. We change the labeling of the first N terms to have a1<b1a2<<bm and γ(ai)=γ(bi) for i>m. Letting J0=[a,a1], Ji=[bi,ai+1], 1i<m, Jm=[bm,b], and Vi=VJi we have γ(Vi)=eli, 0im. We can forget about the cases bi=ai+1.

    Define the function fi:JiR by fi(t)=d(eli,γ(t)). For t,sJi we have |fi(t)fi(s)|d(γ(t),γ(s)), so fi is absolutely continuous and then fi exists Lebesgue almost everywhere in Ji. Let tVin{an,bn} be a point where fi exists. There is a sequence nk such that ankt and γ(ank)=eli. Thus fi(t)=0, which means that ˙γ(t)=0.

    If (aj,bj)Ji then |˙γ|=|fi| Lebesgue almost everywhere in (aj,bj), so that

    Ji|˙γ|=Ji|fi|,

    and then

    ba|˙γ|=mi=0Ji|˙γ|+mi=1biai|˙γ|<.

    It is also easy to see that for t,sJi we have d(γ(t),γ(s))ts|fi|, and using the partition aa1<b1a2<<bmb to make a partition of [c,d] we get d(γ(c),γ(d))dc|˙γ|.

    In this crucial part of the paper we prove that in the framework of graphs we have the lower semicontinuity of the action and apriori bounds for the Lipschitz norm of minimizers. The proofs have the same spirit as in euclidean space, paying attention to what happens at the vertices. Denote by Cac([a,b]) the set of absolutely continuous functions γ:[a,b]G provided with the topology of uniform convergence. We define the action of γCac([a,b]) as

    A(γ)=baL(γ(t),˙γ(t))dt

    A minimizer is a γCac([a,b]) such that for any αCac([a,b]) with α(a)=γ(a), α(b)=γ(b) we have

    A(γ)A(α)

    The following two properties of the Lagrangian are important to achieve our goal and follow from its strict convexity and super-linearity.

    Proposition 2. If C0, ε>0, there is η>0 such that for x,yIj, d(x,y)<η and v,wR, |v|C, we have

    L(y,w)L(x,v)+Lv(x,v)(wv)ε.

    Proposition 3. If Lvvθ>0, C0, ε>0, there is η>0 such that for x,yIj, d(x,y)<η and v,wR, |v|C, we have

    L(y,w)L(x,v)+Lv(x,v)(wv)+3θ4|wv|2ε.

    Lemma 2.1. Let L be a Lagrangian on G. If a sequence γnCac([a,b]) converges uniformly to the curve γ:[a,b]G and

    lim infnA(γn)<

    then the curve γ is absolutely continuous and

    A(γ)lim infnA(γn).

    proof. By the super-linearity of L we may assume that L0. Let c=lim infnA(γn). Passing to a subsequence we can assume that

    A(γn)<c+1,nN

    Fix ε>0 and take B>2(c+1)/ε. Again by super-linearity there is a positive number C(B) such that

    L(x,v)B|v|C(B),xGV,vR

    From Proposition 1 and L0, for E[a,b] measurable we have

    C(B)Leb(E)+BE|˙γn|EL(γn,˙γn)+[a,b]EL(γn,˙γn)c+1.

    Thus

    E|˙γn|1B(c+1+C(B)Leb(E))ε2+C(B)Leb(E)B.

    Choosing 0<δ<εB2C(B) we have that

    Leb(E)<δnNE|˙γn|<ε.

    Since the sequence ˙γn is uniformly integrable, we have that γ is absolutely continuous and ˙γn converges to ˙γ in the σ(L1,L) weak topology.

    Set V=γ1(V). Let ε>0 and Ek={t:|˙γ(t)|k,d(t,V)1k}.

    By Propositions 2, 1, for n large,

    γ1(el)[L(el,0)+Lv(el,0)˙γn(t)ε]γ1(el)L(γn,˙γn),Ek[L(γ,˙γ)+Lv(γ,˙γ)(˙γn˙γ)ε]EkL(γn,˙γn),EkV[L(γ,˙γ)+Lv(γ,˙γ)(˙γn˙γ)ε]EkVL(γn,˙γn)A(γn).

    Letting n+ we have that

    EkVL(γ,˙γ)c+ε(ba)

    Since Ek[a,b]V when k+ and L0, we have

    A(γ)=limk+EkVL(γ,˙γ)c+ε(ba)

    Now let ε0.

    Lemma 2.1 implies

    Theorem 2.2. Let L be a Lagrangian on G. The action A:Cac([a,b])R{} is lower semicontinuous.

    Proposition 4. Let L be a Lagrangian on G.

    The set {γCab([a,b]):A(γ)K} is compact with the topology of uniform convergence.

    Let Ct=sup{L(x,v):xG,|v|diam(G)t}, then for any minimizer γ:[a,b]G with bat we have

    A(γ)C(ba).

    Proposition 5. Suppose γnCab([a,b]) converge uniformly to γ:[a,b]G and A(γn) converges to A(γ), then ˙γn converges to ˙γ in L1[a,b]

    proof. Let F[a,b] be a finite union of intervals. From Lemma 2.1 we have

    FL(γ,˙γ)lim infnFL(γn,˙γn) and [a,b]FL(γ,˙γ)lim infn[a,b]FL(γn,˙γn)

    Since

    limnFL(γn,˙γn)+[a,b]FL(γn,˙γn)=limnA(γn)=A(γ)

    we have

    limnFL(γn,˙γn)=FL(γ,˙γ). (1)

    If V=γ1(V)F then

    lim supnVL(γn,˙γn)limnFL(γn,˙γn)=FL(γ,˙γ).

    Thus

    lim supnVL(γn,˙γn)VL(γ,˙γ). (2)

    As in Lemma 2.1, the ˙γn are uniformly integrable, so they converge to ˙γ in the σ(L1,L) weak topology and then, for any Borel set B where ˙γ is bounded,

    limnBLv(γ,˙γ)(˙γn˙γ)=0. (3)

    Given ε>0, from Proposition 3 we have that for n large enough

    3θ4γ1(el)|˙γn|2γ1(el)[L(γn,˙γn)L(el,0)Lv(el,0)˙γn+ε].

    Which together with Proposition 1 and equations (2), (3) give lim supnV|˙γn|2ε for any ε>0. So

    limnV|˙γn|2=0. (4)

    For k>0 let Dk:={t[a,b]:|˙γ(t)|>k}, Bk:={t[a,b]:d(t,V)>1k}. Then limkLeb(Dk)=limkLeb([a,b]VBk)=0. Let Fk be a finite union of intervals such that DkBkFkBk and Leb(Fk(DkBk))<1k. Then limk Leb(Fk)=0.

    Given ε>0, from Proposition 3 we have that for n large enough

    3θ4BkFk|˙γn˙γ|2BkFk[L(γn,˙γn)L(γ,˙γ)Lv(γ,˙γ)(˙γn˙γ)+ε].

    From (1), (3) we get that lim supnBkFk|˙γn˙γ|2ε for any ε>0. So

    limnBkFk|˙γn˙γ|2=0. (5)

    Since {˙γn} is uniformly integrable, given ε>0, for k sufficiently large we have

    Fk[a,b]VBk|˙γn˙γ|Fk[a,b]VBk|˙γn|+|˙γ|<ε, (6)

    From (4), (5), (6) and Cauchy-Schwartz inequality, we have that for any ε>0

    lim supnba|˙γn˙γ|limnV|˙γn|+lim supn|˙γn˙γ|Fk[a,b]VBk+limnBkFk|˙γn˙γ|ε

    Lemma 2.3. Let L be a Lagrangian in G. For ε>0 there exists Kε that is a Lipschitz constant for any minimizer γ:[a,b]G with baε.

    proof. Note that if γ is a minimizer and γ(c,d)Ij then γ|(c,d) is a solution of the Euler Lagrange equation for Lj.

    Suppose the Lemma is not true, then by Proposition 1, for any iN there are a minimizer γi:[si,ti]G with tisiε and a set Ei[si,ti]γ1i(V) with Leb(Ei)>0 such that |˙γ|>i on Ei. Let ciEi. Translating [si,ti] we can assume that ci=c for all i and taking a subsequence that there is aR such that γi is defined in [a,a+ε2],c[a,a+ε2]. As A(γi|[a,a+ε2]) is bounded, by Proposition 4 there is subsequence γi|[a,a+ε2] which converges uniformly to γ:[a,a+ε2]G. Since γ is limit of minimizers, it is a minimizer and A(γ)lim infA(γi|[a,a+ε2]). We can not have that A(γ)<lim supA(γi|[a,a+ε2]) because that would contradict that the γi are minimizers. Thus A(γ)=limA(γi|[a,a+ε2]).

    If γ(c)IjV, there is δ>0 such that γ([cδ,c+δ])Ij.

    If γ(c)=el we have 2 possibilities (not mutually exclusive)

    a) There is an edge Ij with elIj and infinitely many i's such that γi(c)Ij and ˙γi(c) points towards el.

    b) There is an edge Ij with elIj and infinitely many i's such that γi(c)Ij and ˙γi(c) points towards the other vertex.

    In case a) there is δ>0 such that γ([cδ,c])Ij.

    In case b) there is δ>0 such that γ([c,c+δ])Ij.

    We have that γ is a solution of the Euler-Lagrange equation for Lj either on [cδ,c] or on [c,c+δ] and then |˙γ(t)|K on [cδ,c] or [c,c+δ]. For some 0<δ1<δ we have that γi are solutions of the Euler-Lagrange equation for Lj on [cδ1,c] or on [c,c+δ1]. For i suficiently large, we have that |˙γi|>2K either on [cδ1,c] or on [c,c+δ1]. This would contradict Proposition 5.

    The content of this section is similar to that for Lagrangians on compact manifolds. We only give the proofs that are different from those in the compact manifold case, which can be found in [7], as well as extensions in [8].

    Given x,yG let Cac(x,y,t) be the set of curves αCac([0,t]) such that α(0)=x and α(t)=y. For a given real number k define

    ht(x,y)=minαCac(x,y,t)A(α)

    and

    hk(x,y)=lim inftht(x,y)+kt

    Lemma 3.1. For ε>0 the function F:[ε,)×G×GR defined by F(t,x,y)=ht(x,y) is Lipschitz.

    Lemma 3.2. There exists a real c independent of x and y such that

    1. For all k>c we have hk(x,y)=.

    2. For all k<c we have hk(x,y)=

    3.hc(x,y) is finite. The function h:=hc is called the Peierls barrier.

    Lemma 3.3. The value c is the infimum of k such that γL+k0 for all closed curves γ.

    Definition 3.4. The Mañé potencial Φ:G×GR is defined by

    Φ(x,y)=inft>0ht(x,y)+ct.

    Clearly we have Φ(x,y)h(x,y) for any x,yG.

    Proposition 6. Functions h and Φ have the following properties.

    1. Φ(x,z)Φ(x,y)+Φ(y,z).

    2. h(x,z)h(x,y)+Φ(y,z), h(x,z)Φ(x,y)+h(y,z).

    3. h and Φ are Lipschitz

    4. If γn:[0,tn]G is a sequence of absolutely continuous curves with tn and γn(0)x, γn(tn)y, then

    h(x,y)lim infnA(γn)+ctn. (7)

    Definition 3.5. A curve γ:JG defined on an interval J is called

    semi-static if

    Φ(γ(t),γ(s))=stL(γ,˙γ)+c(st)

    for any t,sJ, ts.

    static if

    stL(γ,˙γ)+c(st)=Φ(γ(s),γ(t))

    for any t,sJ, ts.

    The Aubry set A is the set of points xG such that h(x,x)=0.

    Notice that by item (2) in Proposition 6, h(x,z)=Φ(x,z) if xA or zA.

    Proposition 7. If η:RG is static then η(s)A for any sR.

    We do not have a Lagrangian flow and therefore we can not speak of conservation of energy. Nevertheless the following Proposition says that semi-static curves have energy c(L).

    Proposition 8. Let η:JG be semi-static. For almost every tJ

    Lv(η(t),˙η(t))˙η(t)=L(η(t),˙η(t))+c

    proof. For λ>0, let ηλ(t):=η(λt) so that ˙ηλ(t)=λ˙η(λt) almost everywhere.

    For r,sJ let

    Ars(λ):=s/λr/λ[L(ηλ(t),˙ηλ(t))+c]dt=sr[L(η(s),λ˙η(s))+c]dsλ.

    Since η is a free-time minimizer, differentiating Ars(λ) at λ=1, we have that

    0=Ars(1)=T0[Lv(η(s),˙η(s))˙η(s)L(˙η(s),˙η(s)c]ds.

    Since this holds for any r,sJ we have

    Lv(η(t),˙η(t))˙η(t)=L(η(t),˙η(t))+c

    for almost every tJ.

    Following Fathi [7], we define weak KAM solutions and give some of their properties

    Definition 3.6. Let c be given by Lemma 3.2.

    A function u:GR is dominated if for any x,yG, we have

    u(y)u(x)ht(x,y)+ctt>0,

    or equivalently

    u(y)u(x)Φ(x,y).

    γ:IG calibrates a dominated function u:GR if

    u(γ(s))u(γ(t))=stL(γ,˙γ)+c(st)s,tI

    A continuous function u:GR is a backward (forward) weak KAM solution if it is dominated and for any xG there is γ:(,0]G (γ:[0,)G) that calibrates u and γ(0)=x

    Corollary 1. Any static curve γ:JG calibrates any dominated function u:GR

    Proposition 9. For any xG, h(x,) is a backward weak KAM solution and h(,x) is a forward weak KAM solution.

    proof. By item (2) of Proposition 6, h(x,) is dominated.

    The standard construction of calibrating curves for compact manifolds involves the Euler Lagange flow that we do not have, so we use a diagonal trick. Let γn:[tn,0]G be a sequence of minimizing curves connecting x to y such that

    h(x,y)=limnA(γn)+ctn

    By Lemma 2.3, {γn} is uniformly Lipschitz and then equicontinuous. It follows from the Arzela Ascoli Theorem that there is a sequence n1j such that γn1j converges uniformly on [1,0]. Again, by the Arzela Ascoli Theorem, there is a subsequence (n2j)j of the sequence (n1j)j such that γn2j converges uniformly on [2,0]. By induction, this procedure gives for each kN a sequence (nkj)j that is a subsequence of the sequence (nk1j)j and such that γnkj converges uniformly on [k,0] as j. Letting mk=nkk, the sequence γmk converges uniformly on each [l,0]. For s<0 define γ(s)=limkγmk(s). Fix t<0, for k large t+tmk0 and

    A(γmk)+ctmk=ttmkL(γmk,˙γmk)+c(t+tmk)+0tL(γmk,˙γmk)ct. (8)

    Since γmk converges to γ uniformly on [t,0], we have

    lim infk0tL(γmk,˙γmk)0tL(γ,˙γ).

    From item (4) of Proposition 6 we have

    h(x,γ(t))lim infkttmkL(γmk,˙γmk)+c(t+tmk).

    Taking lim infk in (8) we get

    h(x,y)h(x,γ(t))+0tL(γ,˙γ)ct.

    So γ calibrates h(x,).

    From Proposition 9 we have

    Corollary 2. If xA there exists a curve γ:RG such that γ(0)=x and for all t0

    h(γ(t),x)=t0L(γ,˙γ)cth(x,γ(t))=0tL(γ,˙γ)ct.

    In particular the curve γ is static and calibrates any dominated function u:GR.

    Theorem 3.7. The function Φ(x,) is a backward weak KAM solution if and only if xA.

    Corollary 3. Let CG and w0:CR be bounded from below. Let

    w(x)=infzCw0(z)+Φ(z,x)

    1. w is the maximal dominated function not exceeding w0 on C.

    2. If CA, w is a backward weak KAM solution.

    3. If for all x,yC

    w0(y)w0(x)Φ(x,y),

    then w coincides with w0 on C.

    For u:GR let I(u) be the set of points xG for which exists γ:RG such that γ(0)=x and γ calibrates u.

    Corollary 4.

    A=udominatedI(u)

    Proposition 10. For each x,yG with xy we can find ε>0 and a curve γ:[ε,0]G such that γ(0)=y and for all t[0,ε]

    Φ(x,γ(0))Φ(x,γ(t))=0tL(γ,˙γ)+ct.

    In particular, for each xG the function G{x}R; yΦ(x,y) is a backward weak KAM solution.

    Theorem 3.8. A is nonempty and if u:GR is a backward weak KAM solution then

    u(x)=minqAu(q)+h(q,x) (9)

    Corollary 5.

    h(x,y)=minqAh(x,q)+h(q,y)=minqAΦ(x,q)+Φ(q,x)

    Let F be the set of real functions on G, bounded from below.

    The backward Lax semigroup Lt:FF, t>0 is defined by

    Ltf(x)=infyGf(y)+ht(y,x).

    It is clear that fF is dominated if and only if fLtf+ct for any t>0.

    It follows at once that LtLs=Lt+s and

    LtfLtgfg (10)

    The proof of the following Lemma is the same as in the compact manifold case.

    Lemma 4.1. Given ε>0 there is Kε>0 such that for each u:GR continuous, tε, we have Ltu:GR is a Lipschitz with constant Kε.

    Theorem 4.2. A continuous function u:GR is a fixed point of the semigroup Lt+ct if and only if it is a backward weak KAM solution

    proof. Suppose u:GR is a fixed point of the semigroup Lt+ct. For each T2 there is a curve αT:[T,0]G such that αT(0)=x and

    u(x)u(αT(T))=A(αT)+cT.

    By Lemma 2.3 {αT} is uniformly Lipschitz. As in Propostion 9 one obtains a sequence tk and γ:(,0]G such that αtk converges to γ, uniformly on each [n,0].

    By Lemma 2.2

    0nL(γ,˙γ)+nclim infk0nL(αtk,˙αtk)+nc=lim infku(x)u(αtk(n))=u(x)u(γ(n))

    Suppose now that u:GR is a backward weak KAM solution. Since u is dominated, uLtu+ct. For xG let γ:(,0]G be such that γ(0)=x and for all t>0

    u(x)u(γ(t))=0tL(γ,˙γ)+ct.

    Thus

    u(x)u(γ(t))+ht(γ(t),x)+ctLtu(x)+ct.

    From Proposition 9 and Theorem 4.2 one obtains

    Corollary 6. The semigroup Lt+ct has fixed points.

    Without loss of generality assume c=0. For uC(G) define

    v(x):=minzGu(z)+h(z,x). (11)

    Proposition 11. Let ψ=limnLtnu for some tn, then

    ψv. (12)

    proof. For xG let γn:[0,tn]G be such that γn(tn)=x and

    Ltnu(x)=u(γn(0))+A(γn). (13)

    Passing to a subsequence if necessary we may assume that γn(0) converges to yG. Taking lim inf in (13), we have from item (4) of Proposition 6

    ψ(x)=u(y)+lim infnA(γn)u(y)+h(y,x).

    Proposition 12. If Ltu converges as t, then the limit is function v defined in.

    proof. For xG let zG be such that v(z)=u(z)+h(z,x). Since Ltu(x)u(z)+ht(z,x), we have

    limtLtu(x)lim inftu(z)+ht(z,x)=v(z)

    which together with Proposition (11) gives limtLtu=v.

    Thus, given uC(G) our goal is to prove that Ltu converges to v defined in (11).

    Remark 1. Using Corollary 5 we can write (11) as

    v(x)=minyAΦ(y,x)+w(y) (14)
    w(y):=infzGu(z)+Φ(z,y) (15)

    Item (1) of Corollary 3 states that w is the maximal dominated function not exceeding u. Items (2), (3) of the same Corollary imply that v is the unique backward weak KAM solution that coincides with w on A.

    Proposition 13. Suppose that u is dominated, then Ltu converges uniformly as t to the function v given by (11).

    proof. Since u is dominated, the function tLtu is nondecreasing. As well, in this case, w given by (15) coincides with u. Items (1) and (3) of Corollary 3 imply that v is the maximal dominated function that coincides with u on A and then uv on G.

    Since the semigroup Lt is monotone and v is a backward weak KAM solution

    LtuLtv=v for any t>0.

    Thus the uniform limit limtu exists.

    We now address the convergence of Lt following the lines in [5] and [11].

    For uC(G) let

    ωL(u):={ψC(G):tn such that ψ=limnLtnu}.u_(x):=sup{ψ(x):ψωL(u)} (16)
    ¯u(x):=inf{ψ(x):ψωL(u)} (17)

    From these and Proposition 11

    Proposition 14. Let uC(G), v be the function given by (11), u_,¯u defined in (16) and (17). Then

    v¯uu_ (18)

    Proposition 15. For uC(G), function u_ given by (16) is dominated.

    proof. Let x,yG. Given ε>0 there is ψ=limnLtnu such that u_(x)ε<ψ(x). For n>N(ε) and a>0

    u_(x)2ε<ψ(x)εLtnu(x)=La(Ltnau)(x)Ltnau(y)+ha(y,x).

    Choose a divergent sequence nj such that (Ltnjau)j converges uniformly. For j>ˉN(ε), Ltnjau(y)<u_(y)+ε, and then

    u_(x)3ε<Ltnjau(y)+ha(y,x)ε<u_(y)+ha(y,x).

    Denote by K the family of static curves η:RG, and for yA denote by K(y) the set of curves ηK with η(0)=y.

    Proposition 16. K is a compact metric space with respect to the uniform convergence on compact intervals.

    proof. Let {ηn} be a sequence in K. By Lemma 2.3, {ηn} is uniformly Lipschitz. As in Proposition 9 we obtain a sequence nk such that ηnk converges to η:RG uniformly on each [a,b] and then η is static.

    Proposition 17. Two dominated functions that coincide on M=ηKω(η) also coincide on A.

    proof. Let φ1, φ2 be two dominated functions coinciding on M. Let yA and ηK(y). Let (tn)n be a diverging sequence such that limnη(tn)=xM. By Corollary 1

    φi(y)=φi(η(0))Φ(y,η(0))=φi(η(tn))Φ(y,η(tn))

    for every nN,i=1,2. Sending n to , we get

    φ1(y)=limnφ1(η(tn))Φ(y,η(tn))=φ1(x)Φ(y,x)=φ2(x)Φ(y,x)=limnφ2(η(tn))Φ(y,η(tn))=φ2(y).

    Proposition 18. Let ηK, ψC(G) and φ be a dominated function. Then the function t(Ltψ)(η(t))φ(η(t)) is nonincreasing on R+.

    proof. From Corollary 1, for t<s we have

    (Lsψ)(η(s))(Ltψ)(η(t))stL(η(τ),˙η(τ))dτ=φ(η(s))φ(η(t))

    Lemma 4.3. There is a M>0 such that, if η is any curve in K and λ is sufficiently close to 1, we have

    t2t1L(ηλ,˙ηλ)Φ(ηλ(t1),ηλ(t2))+M(t2t1)(λ1)2 (19)

    for any t2>t1, where ηλ(t)=η(λt).

    proof. Let K>0 be a Lipschitz constant for any minimizer γ:[a,b]G with ba>1, 2R=sup{|Lvv(x,v)|:|v|K}. For λ(1δ,1+δ) fixed, using Proposition 8

    t2t1L(ηλ(t),˙ηλ(t))dt=t2t1[L(η(λt),˙η(λt))+(λ1)Lv(η(λt),˙η(λt))˙η(λt)+12(λ1)2Lvv(η(λt),μ˙η(λt))(˙η(λt))2]dtλt2t1L(η(λt),˙η(λt))dt+(t2t1)RK2(λ1)2=Φ(η(λt1),η(λt2))+(t2t1)RK2(λ1)2

    Proposition 19. Let ηK, ψC(G) and φ be a dominated function. Assume that D+((ψφ)η)(0){0} where D+ denote the super-differential. Then for all t>0 we have

    (Ltψ)(η(t))φ(η(t))<ψ(η(0))φ(η(0)) (20)

    proof. Fix t>0. By Corollary 1 it is enough to prove (20) for φ=Φ(,η(t)). Since Lt(ψ+a)=Ltψ+a we can assume that ψ(η(0))=φ(η(0)).

    (Ltψ)(η(t))φ(η(t))=(Ltψ)(η(t))t/λ(1/λ1)tL(ηλ,˙ηλ)+ψ(η((1λ)t)),

    thus, by Lemma 4.3

    (Ltψ)(η(t))φ(η(t))ψ(η((1λ)t))φ(η((1λ)t))+Mt(λ1)2.

    If mD+((ψφ)η)(0){0}, we have

    (Ltψ)(η(t))φ(η(t))m((1λ)t)+o((1λ)t))+Mt(λ1)2,

    where limλ1o((1λ)t)1λ=0. Choosing appropriately λ close to 1, we get

    (Ltψ)(η(t))φ(η(t))<0.

    Proposition 20. Suppose φ is dominated and ψωL(u). For any yM there exists γK(y) such that the function tψ(γ(t))φ(γ(t)) is constant.

    proof. Let (sk)k and (tk)k be diverging sequences, η be a curve in K such that y=limkη(sk), and ψ is the uniform limit of Ltku. As in Proposition 9, we can assume that the sequence of functions tη(sk+t) converges uniformly on compact intervals to γ:RG, and so γK. We may assume moreover that tksk, as k, and that Ltksku converges uniformly to ψ1ωL(u). By the semi-group property and (10)

    LtkuLskψ1Ltkskuψ1

    which implies that Lskψ1 converges uniformly to ψ. From Proposition 18, we have that for any τR s(Lsψ1)(η(τ+s))φ(η(τ+s)) is a nonincreasing function in R+, and hence it has a limit l(τ) as s, which is finite since l(τ)¯uφ. Given t>0, we have

    l(τ)=limk(Lsk+tψ1)(η(sk+τ+t))φ(η(sk+τ+t))=(Ltψ)(γ(τ+t))φ(γ(τ+t))

    The function t(Ltψ)(γ(τ+t))φ(γ(τ+t)) is therefore constant on R+. Applying Proposition 19 to the curve γ(τ+)K, we have D+((ψφ)γ)(τ){0}= for any τR. This implies that ψφ is constant on γ.

    Proposition 21. Let ηK, ψωL(u) and v be defined by (11). For any ε>0 there exists τR such that

    ψ(η(τ))v(η(τ))<ε.

    proof. Since the curve η is contained in A, we have

    v(η(0))=minzGu(z)+Φ(z,η(0)),

    and hence v(η(0))=u(z0)+Φ(z0,η(0)), for some z0G. Take a curve γ:[0,T]G such that

    v(η(0))+ε2=u(z0)+Φ(z0,η(0))+ε2>u(z0)+T0L(γ,˙γ)LTu(η(0)).

    Choosing a divergent sequence (tn)n such that Ltnu converges uniformly to ψ we have for n sufficiently large

    Ltnuψ<ε2,tnT>0.

    Take τ=tnT

    ψ(η(τ))ε2<Ltnu(η(τ))=LτLTu(η(τ))=LTu(η(0))+τ0L(η,˙η)<ε2+v(η(0))+τ0L(η,˙η)=ε2+v(η(τ))

    From Propositions 20 and 21 we obtain

    Theorem 4.4. Let ψωL(u) and v be defined by (11). Then ψ=v on M.

    Theorem 4.5. Let uC(G), then Ltu converges uniformly as t to v given by (11).

    proof. The function u_ is dominated and coincides with v on M by Theorem 4.4. Proposition 17 implies that u_ coincide with v on A and so does with w. By item (1) of Corollary 3 we have u_v.

    In this section we compare weak KAM and viscosity solutions.

    Definition 5.1.    A continuous real function φ defined on the neighborhood of el is C1 if for every j with elIj, φ|Ij is C1.

    A continuous real function φ defined on the neighborhood of (el,t) is C1 if for every j with elIj, φ|Ij×(tδ,t+δ) is C1.

    Note that if α:[0,δ]Ij is differentiable and α(0)=el, then α+(0)TelIj and we have

    Djφ(el)z=(φα)+(0).

    We consider the Hamiltonian consisting in functions Hj:Ij×RR given by

    Hj(x,p)=max{pzLj(x,z):zTxIj,xVzTxIj,xIjV}

    and the Hamilton Jacobi equations

    H(x,Du(x))=c, (21)
    ut(x,t)+H(x,Dxu(x,t))=0. (22)

    Note that if L is symmetric at the vertices, then for any vertex el there is a function ha such that Hj(el,p)=ha(|p|) for any j with elIj. This kind of Hamiltonians are called of eikonal type [3].

    The following definition appeared in [3] and [4].

    Definition 5.2. A function u:GR is a

    viscosity subsolution of (21) if satisfies the usual definition in GV and for any C1 function φ on the neighborhood of any el s.t. uφ has a maximum at el we have

    max{Hj(el,Djφ(el)):elIj}c.

    viscosity supersolution of (21) if satisfies the usual definition in GV and for any C1 function φ on the neighborhood of any el s.t. uφ has a minimum at el we have

    max{Hj(el,Djφ(el)):elIj}c

    viscosity solution if it is both, a subsolution and a supersolution.

    A function u:G×[0,)R is a viscosity subsolution of (22) if satisfies the usual definition in GV×[0,) and for any C1 function φ on the neighborhood of any (el,t) s.t. uφ has a maximum at (el,t) we have

    φt(el,t)+max{Hj(el,Djφ(el,t)):elIj}c.

    viscosity supersolution of (21) if satisfies the usual definition in GV×[0,) and for any C1 function φ on the neighborhood of any (el,t) s.t. uφ has a minimum at (el,t) we have

    φt(el,t)+max{Hj(el,Djφ(el,t)):elIj}c

    viscosity solution if it is both, a subsolution and a supersolution.

    Proposition 22. If u:GR is dominated then then it is a viscosity subsolution of (21). If u is a backward weak KAM solution then it is a viscosity solution.

    proof. Suppose u:GR is dominated. Let φ be a C1 function on the neighborhood of el s.t. uφ has a maximum at el, j s.t. elIj, α:[0,δ]Ij differentiable with α(0)=el, z=α(0). Define γ:[δ,0]Ij by γ(s)=α(s).

    φ(el)φ(γ(s))u(el)u(γ(s))0sLj(γ,˙γ)csφ(el)φ(α(t)))t1t0tLj(γ,˙γ)+cDjφ(el)zLj(el,z)+c.

    So u is a subsolution.

    Let φ be a C1 function on the neighborhood of el s.t. uφ has a minimum at el. Let γ:(,0]G be such that γ(0)=el and for t<0

    u(el)u(γ(t))=0tLj(γ,˙γ)ct

    Let δ>0,j be such that γ([δ,0])Ij.

    φ(el)φ(γ(s))0sLj(γ,˙γ)cs

    Define α:[0,δ]Ij by α(t)=γ(t), z=α(0),

    φ(el)φ(α(t))t1t0tLj(γ,˙γ)+cDjφ(el)zLj(el,z)+c.

    So u is a supersolution.

    Our approach to get a converse to Proposition 22 is to prove uniqueness of solutions to the Cauchy problem for (22), using a comparison principle. For that purpose the symmetry Lagrangian is a sufficient condition. It may be possible that other assumptions imply the required uniqueness or that a different approach gives a converse to Proposition 22.

    Proposition 23. Let f:GR be continuous and define u:G×[0,)R by u(x,t)=Ltf(x), then u is a viscosity solution of (22)

    proof. Since Ltf=Lts(Lsf) if 0s<t, for any γ:[s,t]G

    u(γ(t),t)u(γ(s),s)tsL(γ,˙γ) (23)

    and for any xG there is γ:[s,t]G with γ(t)=x such that equality in (23) holds.

    Let φ be a C1 function on the neighborhood of (el,t) s.t. uφ has a maximum at (el,t), j s.t. elIj, α:[0,δ]Ij differentiable with α(0)=el, z=α(0). Define γ:[tδ,t]Ij by γ(s)=α(ts).

    φ(el,t)φ(γ(s),s)u(el,t)u(γ(s),s)tsLj(γ,˙γ)φ(el,t)φ(α(ts),s))ts1tstsLj(γ,˙γ)φt(el,t)Djxφ(el,t)zLj(el,z).

    So u is subsolution.

    Let φ be a C1 function on the neighborhood of (el,t) s.t. uφ has a minimum at (el,t). Let γ:[t1,t]G be such that γ(t)=el and

    u(el,t)u(γ(t1),t1)=tt1L(γ,˙γ)

    Let δ>0,j be such that γ([tδ,t])Ij. For s[tδ,t]

    φ(el,t)φ(γ(s),s)tsLj(γ,˙γ)

    Define α:[0,δ]Ij by α(s)=γ(ts), z=α(0),

    φ(el,t)φ(α(ts),s)ts1tstsLj(γ,˙γ)φt(el,t)Djxφ(el,t)zLj(el,z).

    So u is supersolution.

    Proposition 24. Suppose the Lagrangian is symmetric at the vertices. Let u,v:G×[0,T]R be respectively a Lipschitz viscosity sub, supersolution of (22) such that u(x,0)v(x,0), for any xG. Then uv.

    proof. Suppose that there are x,t such that δ=u(x,t)v(x,t)>0. Let 0<ρδ4t and define Φ:G2×[0,T]2 by

    Φ(x,y,t,s)=u(x,t)v(y,s)d(x,y)2+|ts|22ερ(t+s).

    From the previous definitions we have

    δ2δ2ρt=Φ(x,x,t,t)supG2×[0,T]2Φ=Φ(xε,yε,tε,sε). (24)

    It follows from Φ(xε,xε,tε,tε)+Φ(yε,yε,sε,sε)2Φ(xε,yε,tε,sε) that

    d(xε,yε)2+|tεsε|22εu(xε,tε)u(yε,sε)+v(xε,tε)v(yε,sε)C(d(xε,yε)2+|tεsε|2)1/2

    Thus, there is a sequence ε0 such that xε,yε converge to ¯xG and tε,sε converge to ¯t[0,T] and (24) gives

    δ2Φ(¯x,¯x,¯t,¯t)u(¯x,¯t)v(¯x,¯t),

    and so ¯t0.

    Define the test functions

    φ(x,t)=v(yε,sε)+d(x,yε)2+|tsε|22ε+ρ(t+sε)ψ(y,s)=u(xε,tε)d(xε,y)2+|tεs|22ερ(tε+s).
    φt(xε,tε)=tεsεε+ρ,ψs(yε,sε)=tεsεερ

    Since uφ has maximum at (xε,tε), vψ has minimum at (yε,sε), u is subsolution and v is supersolution,

    2ρ=φt(xε,tε)ψs(yε,sε)max{Hj(yε,Djy(d(xε,y)22ε)(yε)):yεIj}max{Hj(xε,Djx(d(x,yε)22ε)(xε)):xεIj} (25)

    Since ρ>0 we can not have xε=yε.

    If ¯x is not a vertex, ¯xIj, for ε>0 small we have

    Djx(d(x,yε)22ε)(xε)=±d(xε,yε)ε=Djy(d(xε,y)22ε)(yε).

    If we denote by a(xε,yε) this common value, then (25) becomes

    2ρHj(yε,a(xε,yε))Hj(xε,a(xε,yε))

    with a(xε,yε) bounded as ε0, giving a contradiction.

    Suppose now that ¯x=el. For ε>0 small we distinguish the following cases

    1. Neither xε nor yε is a vertex. If xε,yεIj, d(xε,yε)=|σj(xε)σj(yε)|. If xεIi, yεIj, and elIiIj, then d(xε,yε)=d(xε,el)+d(el,yε). In both subcases

    |Dix(d(x,yε)22ε)(xε)|=d(xε,yε)ε|Djy(d(xε,y)22ε)(yε)|=d(xε,yε)ε

    Then (25) becomes

    2ρHj(yε,±d(xε,yε)ε)Hi(xε,±d(xε,yε)ε).

    2. Suppose xε=el, yεIjV.

    |Djx(d(x,yε)22ε)(el)|=d(el,yε)ε|Djy(d(el,y)22ε)(yε)|=d(el,yε)ε

    Since

    Hj(el,±d(el,yε)ε)=ha(el,d(el,yε)ε),

    we have that (25) becomes

    2ρHj(yε,±d(el,yε)ε)=ha(el,d(el,yε)ε).

    3. If yε=el, xεIjV we get in the same way that (25) becomes

    2ρha(el,d(xε,el)ε)Hj(xε,±d(xε,el)ε).

    Since d(xε,yε)ε remains bounded as ε0, we get a contradiction.

    Corollary 7. Suppose the Lagrangian is symmetric at the vertices. Let u,v:G×[0,T]R be viscosity solutions of (22) such that u(x,0)=v(x,0) for any xG. Then u=v.

    Corollary 8. Suppose the Lagrangian is symmetric at the vertices. Let f:GR be a viscosity solution of (21), then f is a fixed point of the Lax semigroup Lt+ct.

    proof. We next show that u(x,t)=f(x)ct is a viscosity solution of (21). Proposition 23 and Corollary 7 then imply that fct=Ltf.

    Let φ be a C1 function on the neighborhood of (el,t) s.t. uφ has a maximum at (el,t). Then scsφ(el,s) has a maximum at t and so φt(el,t)=c. Since fφ(,t) has a maximum at el we have

    max{Hj(el,Djφ(el,t)):xIj}c=φt(el,t),

    so u is a subsolution of (22). Similarly u is a supersolution of (22).

    Corollary 9. Suppose the Lagrangian is symmetric at the vertices. Let u:GR be a viscosity solution of (21) then the representation formula (9) holds.

    proof. By Proposition 22 and Corollary 9, u is a backward weak KAM solution and by Theorem (3.8), formula (9) holds.

    [1] Hamilton-Jacobi equations constrained on networks. Nonlinear Differ. Equ. Appl. (2013) 20: 413-445.
    [2] A Bellman approach for two-domains optimal control problems in \begin{document}${\mathbb{R}^n}$\end{document}
    . ESAIM: Control, Optimisation and Calculus of Variations (2013) 19: 710-739.
    [3] Viscosity solutions of Eikonal equations on topological networks. Calc. Var. Partial Differential Equatons (2013) 46: 671-686.
    [4] A comparison among various notions of viscosity solution for Hamilton-Jacobi equations on networks. J. Math. Anal. Appl. (2013) 407: 112-118.
    [5] A generalized dynamical approach to the large time behavior of solutions of Hamilton-Jacobi equations. SIAM J. Math. Anal. (2006) 38: 478-502.
    [6] Sur la convergence du semi-groupe de Lax-Oleinik. C. R. Acad. Sci. Paris Sr. I Math. (1998) 327: 267-270.
    [7] A. Fathi, Weak KAM Theorem in Lagrangian Dynamics, To appear in Cambridge Studies in Advanced Mathematics.
    [8] PDE aspects of Aubry-Mather theory for quasi-convex Hamiltonians. Calc. Var. Partial Differential Equations (2005) 22: 185-228.
    [9] C. Imbert and R. Monneau, Flux-limited Solutions for Quasi-Convex Hamilton-Jacobi Equations on Networks, arXiv: 1306.2428
    [10] Asymptotic solutions for large time of Hamilton-Jacobi equations in Euclidean nspace. Anal. Non Linéaire (2008) 25: 231-266.
    [11] Convergence to steady states or periodic solutions in a class of HamiltonJacobi equations. J. Math. Pures Appl. (2001) 80: 85-104.
  • This article has been cited by:

    1. 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
    2. Héctor Sánchez Morgado, Homogenization for sub-riemannian Lagrangians, 2023, 36, 0951-7715, 3043, 10.1088/1361-6544/accdae
    3. Marco Pozza, Antonio Siconolfi, Lax–Oleinik Formula on Networks, 2023, 55, 0036-1410, 2211, 10.1137/21M1448677
  • 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(4557) PDF downloads(68) Cited by(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog