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 [
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
[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 [
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
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
TG=⋃j{j}×TIj/∼ |
where
We say that
We start defining a distance in the most natural way. We say a continuous path
α([t0,t1])⊂Ij(1),α([ti−1,ti])=Ij(i),…,α([tm−1,tm])⊂Ij(m), |
ℓ(α)=|σj(1)(α(t1))−σj(1)(α(a))|+m−1∑i=2sj(i)+|σj(m)(α(b))−σj(m)(α(tm−1))| |
and define a distance on
d(x,y)=min{ℓ(α):α:[a,b]→G is a u.s.g.,α(a)=x,α(b)=y} |
We say a path
If
Let
Next Proposition will allow us to define the action of an absolutely continuous curve.
Proposition 1. Let
(a)
(b)
proof. Write
Let
Define the function
If
∫Ji|˙γ|=∫Ji|f′i|, |
and then
∫ba|˙γ|=m∑i=0∫Ji|˙γ|+m∑i=1∫biai|˙γ|<∞. |
It is also easy to see that for
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
A(γ)=∫baL(γ(t),˙γ(t))dt |
A minimizer is a
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
L(y,w)≥L(x,v)+Lv(x,v)(w−v)−ε. |
Proposition 3. If
L(y,w)≥L(x,v)+Lv(x,v)(w−v)+3θ4|w−v|2−ε. |
Lemma 2.1. Let
lim infn→∞A(γn)<∞ |
then the curve
A(γ)≤lim infn→∞A(γn). |
proof. By the super-linearity of
A(γn)<c+1,∀n∈N |
Fix
L(x,v)≥B|v|−C(B),x∈G∖V,v∈R |
From Proposition 1 and
−C(B)Leb(E)+B∫E|˙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
Leb(E)<δ→∀n∈N∫E|˙γn|<ε. |
Since the sequence
Set
By Propositions 2, 1, for
∫γ−1(el)[L(el,0)+Lv(el,0)˙γn(t)−ε]≤∫γ−1(el)L(γn,˙γn),∫Ek[L(γ,˙γ)+Lv(γ,˙γ)(˙γn−˙γ)−ε]≤∫EkL(γn,˙γn),∫Ek∪V[L(γ,˙γ)+Lv(γ,˙γ)(˙γn−˙γ)−ε]≤∫Ek∪VL(γn,˙γn)≤A(γn). |
Letting
∫Ek∪VL(γ,˙γ)≤c+ε(b−a) |
Since
A(γ)=limk→+∞∫Ek∪VL(γ,˙γ)≤c+ε(b−a) |
Now let
Lemma 2.1 implies
Theorem 2.2. Let
Proposition 4. Let
The set
Let
A(γ)≤C(b−a). |
Proposition 5. Suppose
proof. Let
∫FL(γ,˙γ)≤lim infn∫FL(γn,˙γn) and ∫[a,b]∖FL(γ,˙γ)≤lim infn∫[a,b]∖FL(γn,˙γn) |
Since
limn∫FL(γn,˙γn)+∫[a,b]∖FL(γn,˙γn)=limnA(γn)=A(γ) |
we have
limn∫FL(γn,˙γn)=∫FL(γ,˙γ). | (1) |
If
lim supn∫VL(γn,˙γn)≤limn∫FL(γn,˙γn)=∫FL(γ,˙γ). |
Thus
lim supn∫VL(γn,˙γn)≤∫VL(γ,˙γ). | (2) |
As in Lemma 2.1, the
limn∫BLv(γ,˙γ)(˙γn−˙γ)=0. | (3) |
Given
3θ4∫γ−1(el)|˙γn|2≤∫γ−1(el)[L(γn,˙γn)−L(el,0)−Lv(el,0)˙γn+ε]. |
Which together with Proposition 1 and equations (2), (3) give
limn∫V|˙γn|2=0. | (4) |
For
Given
3θ4∫Bk∖Fk|˙γn−˙γ|2≤∫Bk∖Fk[L(γn,˙γn)−L(γ,˙γ)−Lv(γ,˙γ)(˙γn−˙γ)+ε]. |
From (1), (3) we get that
limn∫Bk∖Fk|˙γn−˙γ|2=0. | (5) |
Since
∫Fk∪[a,b]∖V∖Bk|˙γn−˙γ|≤∫Fk∪[a,b]∖V∖Bk|˙γn|+|˙γ|<ε, | (6) |
From (4), (5), (6) and Cauchy-Schwartz inequality, we have that for any
lim supn∫ba|˙γn−˙γ|≤limn∫V|˙γn|+lim supn∫|˙γn−˙γ|Fk∪[a,b]∖V∖Bk+limn∫Bk∖Fk|˙γn−˙γ|≤ε |
Lemma 2.3. Let
proof. Note that if
Suppose the Lemma is not true, then by Proposition 1, for any
If
If
a) There is an edge
b) There is an edge
In case a) there is
In case b) there is
We have that
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
ht(x,y)=minα∈Cac(x,y,t)A(α) |
and
hk(x,y)=lim inft→∞ht(x,y)+kt |
Lemma 3.1. For
Lemma 3.2. There exists a real
1. For all
2. For all
3.
Lemma 3.3. The value
Definition 3.4. The Mañé potencial
Φ(x,y)=inft>0ht(x,y)+ct. |
Clearly we have
Proposition 6. Functions
1.
2.
3.
4. If
h(x,y)≤lim infn→∞A(γn)+ctn. | (7) |
Definition 3.5. A curve
Φ(γ(t),γ(s))=∫stL(γ,˙γ)+c(s−t) |
for any
∫stL(γ,˙γ)+c(s−t)=−Φ(γ(s),γ(t)) |
for any
Notice that by item (2) in Proposition 6,
Proposition 7. If
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
Proposition 8. Let
Lv(η(t),˙η(t))˙η(t)=L(η(t),˙η(t))+c |
proof. For
For
Ars(λ):=∫s/λr/λ[L(ηλ(t),˙ηλ(t))+c]dt=∫sr[L(η(s),λ˙η(s))+c]dsλ. |
Since
0=A′rs(1)=∫T0[Lv(η(s),˙η(s))˙η(s)−L(˙η(s),˙η(s)−c]ds. |
Since this holds for any
Lv(η(t),˙η(t))˙η(t)=L(η(t),˙η(t))+c |
for almost every
Following Fathi [7], we define weak KAM solutions and give some of their properties
Definition 3.6. Let
u(y)−u(x)≤ht(x,y)+ct∀t>0, |
or equivalently
u(y)−u(x)≤Φ(x,y). |
u(γ(s))−u(γ(t))=∫stL(γ,˙γ)+c(s−t)∀s,t∈I |
Corollary 1. Any static curve
Proposition 9. For any
proof. By item (2) of Proposition 6,
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
h(x,y)=limn→∞A(γn)+ctn |
By Lemma 2.3,
A(γmk)+ctmk=t∫−tmkL(γmk,˙γmk)+c(t+tmk)+∫0tL(γmk,˙γmk)−ct. | (8) |
Since
lim infk→∞∫0tL(γmk,˙γmk)≥∫0tL(γ,˙γ). |
From item (4) of Proposition 6 we have
h(x,γ(t))≤lim infk→∞∫t−tmkL(γmk,˙γmk)+c(t+tmk). |
Taking
h(x,y)≥h(x,γ(t))+∫0tL(γ,˙γ)−ct. |
So
From Proposition 9 we have
Corollary 2. If
h(γ(t),x)=−∫t0L(γ,˙γ)−cth(x,γ(−t))=−∫0−tL(γ,˙γ)−ct. |
In particular the curve
Theorem 3.7. The function
Corollary 3. Let
w(x)=infz∈Cw0(z)+Φ(z,x) |
1.
2. If
3. If for all
w0(y)−w0(x)≤Φ(x,y), |
then
For
Corollary 4.
A=⋂udominatedI(u) |
Proposition 10. For each
Φ(x,γ(0))−Φ(x,γ(−t))=∫0−tL(γ,˙γ)+ct. |
In particular, for each
Theorem 3.8.
u(x)=minq∈Au(q)+h(q,x) | (9) |
Corollary 5.
h(x,y)=minq∈Ah(x,q)+h(q,y)=minq∈AΦ(x,q)+Φ(q,x) |
Let
The backward Lax semigroup
Ltf(x)=infy∈Gf(y)+ht(y,x). |
It is clear that
It follows at once that
‖Ltf−Ltg‖∞≤‖f−g‖∞ | (10) |
The proof of the following Lemma is the same as in the compact manifold case.
Lemma 4.1. Given
Theorem 4.2. A continuous function
proof. Suppose
u(x)−u(αT(−T))=A(αT)+cT. |
By Lemma 2.3
By Lemma 2.2
∫0−nL(γ,˙γ)+nc≤lim infk→∞∫0−nL(αtk,˙αtk)+nc=lim infk→∞u(x)−u(αtk(−n))=u(x)−u(γ(−n)) |
Suppose now that
u(x)−u(γ(−t))=∫0−tL(γ,˙γ)+ct. |
Thus
u(x)≥u(γ(−t))+ht(γ(t),x)+ct≥Ltu(x)+ct. |
From Proposition 9 and Theorem 4.2 one obtains
Corollary 6. The semigroup
Without loss of generality assume
v(x):=minz∈Gu(z)+h(z,x). | (11) |
Proposition 11. Let
ψ≥v. | (12) |
proof. For
Ltnu(x)=u(γn(0))+A(γn). | (13) |
Passing to a subsequence if necessary we may assume that
ψ(x)=u(y)+lim infn→∞A(γn)≥u(y)+h(y,x). |
Proposition 12. If
proof. For
limt→∞Ltu(x)≤lim inft→∞u(z)+ht(z,x)=v(z) |
which together with Proposition (11) gives
Thus, given
Remark 1. Using Corollary 5 we can write (11) as
v(x)=miny∈AΦ(y,x)+w(y) | (14) |
w(y):=infz∈Gu(z)+Φ(z,y) | (15) |
Item (1) of Corollary 3 states that
Proposition 13. Suppose that
proof. Since
Since the semigroup
Ltu≤Ltv=v for any t>0. |
Thus the uniform limit
We now address the convergence of
For
ωL(u):={ψ∈C(G):∃tn→∞ such that ψ=limn→∞Ltnu}.u_(x):=sup{ψ(x):ψ∈ωL(u)} | (16) |
¯u(x):=inf{ψ(x):ψ∈ωL(u)} | (17) |
From these and Proposition 11
Proposition 14. Let
v≤¯u≤u_ | (18) |
Proposition 15. For
proof. Let
u_(x)−2ε<ψ(x)−ε≤Ltnu(x)=La(Ltn−au)(x)≤Ltn−au(y)+ha(y,x). |
Choose a divergent sequence
u_(x)−3ε<Ltnj−au(y)+ha(y,x)−ε<u_(y)+ha(y,x). |
Denote by
Proposition 16.
proof. Let
Proposition 17. Two dominated functions that coincide on
proof. Let
φi(y)=φi(η(0))−Φ(y,η(0))=φi(η(tn))−Φ(y,η(tn)) |
for every
φ1(y)=limn→∞φ1(η(tn))−Φ(y,η(tn))=φ1(x)−Φ(y,x)=φ2(x)−Φ(y,x)=limn→∞φ2(η(tn))−Φ(y,η(tn))=φ2(y). |
Proposition 18. Let
proof. From Corollary 1, for
(Lsψ)(η(s))−(Ltψ)(η(t))≤∫stL(η(τ),˙η(τ))dτ=φ(η(s))−φ(η(t)) |
Lemma 4.3. There is a
∫t2t1L(ηλ,˙ηλ)≤Φ(ηλ(t1),ηλ(t2))+M(t2−t1)(λ−1)2 | (19) |
for any
proof. Let
∫t2t1L(ηλ(t),˙ηλ(t))dt=∫t2t1[L(η(λt),˙η(λt))+(λ−1)Lv(η(λt),˙η(λt))˙η(λt)+12(λ−1)2Lvv(η(λt),μ˙η(λt))(˙η(λt))2]dt≤λ∫t2t1L(η(λt),˙η(λt))dt+(t2−t1)RK2(λ−1)2=Φ(η(λt1),η(λt2))+(t2−t1)RK2(λ−1)2 |
Proposition 19. Let
(Ltψ)(η(t))−φ(η(t))<ψ(η(0))−φ(η(0)) | (20) |
proof. Fix
(Ltψ)(η(t))−φ(η(t))=(Ltψ)(η(t))≤∫t/λ(1/λ−1)tL(ηλ,˙ηλ)+ψ(η((1−λ)t)), |
thus, by Lemma 4.3
(Ltψ)(η(t))−φ(η(t))≤ψ(η((1−λ)t))−φ(η((1−λ)t))+Mt(λ−1)2. |
If
(Ltψ)(η(t))−φ(η(t))≤m((1−λ)t)+o((1−λ)t))+Mt(λ−1)2, |
where
(Ltψ)(η(t))−φ(η(t))<0. |
Proposition 20. Suppose
proof. Let
‖Ltku−Lskψ1‖∞≤‖Ltk−sku−ψ1‖∞ |
which implies that
l(τ)=limk→∞(Lsk+tψ1)(η(sk+τ+t))−φ(η(sk+τ+t))=(Ltψ)(γ(τ+t))−φ(γ(τ+t)) |
The function
Proposition 21. Let
ψ(η(τ))−v(η(τ))<ε. |
proof. Since the curve
v(η(0))=minz∈Gu(z)+Φ(z,η(0)), |
and hence
v(η(0))+ε2=u(z0)+Φ(z0,η(0))+ε2>u(z0)+∫T0L(γ,˙γ)≥LTu(η(0)). |
Choosing a divergent sequence
‖Ltnu−ψ‖∞<ε2,tn−T>0. |
Take
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
Theorem 4.5. Let
proof. The function
In this section we compare weak KAM and viscosity solutions.
Definition 5.1.
Note that if
Djφ(el)z=(φ∘α)′+(0). |
We consider the Hamiltonian consisting in functions
Hj(x,p)=max{−pz−Lj(x,z):z∈T−xIj,x∈Vz∈TxIj,x∈Ij∖V} |
and the Hamilton Jacobi equations
H(x,Du(x))=c, | (21) |
ut(x,t)+H(x,Dxu(x,t))=0. | (22) |
Note that if
The following definition appeared in [3] and [4].
Definition 5.2. A function
max{Hj(el,Djφ(el)):el∈Ij}≤c. |
max{Hj(el,Djφ(el)):el∈Ij}≥c |
A function
φt(el,t)+max{Hj(el,Djφ(el,t)):el∈Ij}≤c. |
φt(el,t)+max{Hj(el,Djφ(el,t)):el∈Ij}≥c |
Proposition 22. If
proof. Suppose
φ(el)−φ(γ(s))≤u(el)−u(γ(s))≤∫0sLj(γ,˙γ)−csφ(el)−φ(α(t)))t≤1t∫0−tLj(γ,˙γ)+c−Djφ(el)z≤Lj(el,z)+c. |
So
Let
u(el)−u(γ(t))=∫0tLj(γ,˙γ)−ct |
Let
φ(el)−φ(γ(s))≥∫0sLj(γ,˙γ)−cs |
Define
φ(el)−φ(α(t))t≥1t∫0−tLj(γ,˙γ)+c−Djφ(el)z≥Lj(el,z)+c. |
So
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
proof. Since
u(γ(t),t)−u(γ(s),s)≤∫tsL(γ,˙γ) | (23) |
and for any
Let
φ(el,t)−φ(γ(s),s)≤u(el,t)−u(γ(s),s)≤∫tsLj(γ,˙γ)φ(el,t)−φ(α(t−s),s))t−s≤1t−s∫tsLj(γ,˙γ)φt(el,t)−Djxφ(el,t)z≤Lj(el,z). |
So
Let
u(el,t)−u(γ(t−1),t−1)=∫tt−1L(γ,˙γ) |
Let
φ(el,t)−φ(γ(s),s)≥∫tsLj(γ,˙γ) |
Define
φ(el,t)−φ(α(t−s),s)t−s≥1t−s∫tsLj(γ,˙γ)φt(el,t)−Djxφ(el,t)z≥Lj(el,z). |
So
Proposition 24. Suppose the Lagrangian is symmetric at the vertices. Let
proof. Suppose that there are
Φ(x,y,t,s)=u(x,t)−v(y,s)−d(x,y)2+|t−s|22ε−ρ(t+s). |
From the previous definitions we have
δ2≤δ−2ρt∗=Φ(x∗,x∗,t∗,t∗)≤supG2×[0,T]2Φ=Φ(xε,yε,tε,sε). | (24) |
It follows from
d(xε,yε)2+|tε−sε|22ε≤u(xε,tε)−u(yε,sε)+v(xε,tε)−v(yε,sε)≤C(d(xε,yε)2+|tε−sε|2)1/2 |
Thus, there is a sequence
δ2≤Φ(¯x,¯x,¯t,¯t)≤u(¯x,¯t)−v(¯x,¯t), |
and so
Define the test functions
φ(x,t)=v(yε,sε)+d(x,yε)2+|t−sε|22ε+ρ(t+sε)ψ(y,s)=u(xε,tε)−d(xε,y)2+|tε−s|22ε−ρ(tε+s). |
φt(xε,tε)=tε−sεε+ρ,ψs(yε,sε)=tε−sεε−ρ |
Since
2ρ=φt(xε,tε)−ψs(yε,sε)≤max{Hj(yε,−Djy(d(xε,y)22ε)(yε)):yε∈Ij}−max{Hj(xε,Djx(d(x,yε)22ε)(xε)):xε∈Ij} | (25) |
Since
If
Djx(d(x,yε)22ε)(xε)=±d(xε,yε)ε=−Djy(d(xε,y)22ε)(yε). |
If we denote by
2ρ≤Hj(yε,a(xε,yε))−Hj(xε,a(xε,yε)) |
with
Suppose now that
1. Neither
|Dix(d(x,yε)22ε)(xε)|=d(xε,yε)ε|Djy(d(xε,y)22ε)(yε)|=d(xε,yε)ε |
Then (25) becomes
2ρ≤Hj(yε,±d(xε,yε)ε)−Hi(xε,±d(xε,yε)ε). |
2. Suppose
|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
2ρ≤Hj(yε,±d(el,yε)ε)=ha(el,d(el,yε)ε). |
3. If
2ρ≤ha(el,d(xε,el)ε)−Hj(xε,±d(xε,el)ε). |
Since
Corollary 7. Suppose the Lagrangian is symmetric at the vertices. Let
Corollary 8. Suppose the Lagrangian is symmetric at the vertices. Let
proof. We next show that
Let
max{Hj(el,Djφ(el,t)):x∈Ij}≤c=−φt(el,t), |
so
Corollary 9. Suppose the Lagrangian is symmetric at the vertices. Let
proof. By Proposition 22 and Corollary 9,
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 |