Convergence/divergence phenomena in the vanishing discount limit of Hamilton-Jacobi equations
Abstract.
We study the asymptotic behavior of solutions of an equation of the form
(*) |
on a closed Riemannian manifold , where is convex and superlinear in the gradient variable, is globally Lipschitz but not monotone in the last argument, and is the critical constant associated with the Hamiltonian . By assuming that satisfies a positivity condition of integral type on the Mather set of , we prove that any equi-bounded family of solutions of (* ‣ Convergence/divergence phenomena in the vanishing discount limit of Hamilton-Jacobi equations) uniformly converges to a distinguished critical solution as . We furthermore show that any other possible family of solutions uniformly diverges to or . We then look into the linear case and prove that the family of maximal solutions to (* ‣ Convergence/divergence phenomena in the vanishing discount limit of Hamilton-Jacobi equations) is well defined and equi-bounded for small enough. When changes sign and enjoys a stronger localized positivity assumption, we show that equation (* ‣ Convergence/divergence phenomena in the vanishing discount limit of Hamilton-Jacobi equations) does admit other solutions too, and that they all uniformly diverge to as . This is the first time that converging and diverging families of solutions are shown to coexist in such a generality.
Key words and phrases:
weak KAM Theory, vanishing discount problems, Mather measures, viscosity solution theory1991 Mathematics Subject Classification:
Introduction
In this paper we are concerned with the asymptotic behavior of solutions of an equation of the form
(Eλ) |
posed on a closed Riemannian manifold , where is convex and superlinear in the gradient variable, is globally Lipschitz in the last argument, and is the critical constant associated with the Hamiltonian . We refer the reader to Section 1.3 for the definition of and of the other related objects coming from weak KAM Theory that will be mentioned in this introduction. The monotonicity condition on in the last argument, that is standard for these kind of equations, is dropped here in favor of the following much weaker integral condition
-
(L5)
where is the convex conjugate function of , denotes the set of Mather measures for , and (and hence ) satisfies a -type regularity condition near , see conditions (G4) and (L4) in Section 2. Under these assumptions, we prove that any equi-bounded family of solutions of (Eλ) uniformly converges to a distinguished critical solution as . We furthermore show that any family consisting of other possible solutions uniformly diverges to or .
We underline that the conditions presented above on are not sufficient to guarantee the existence and uniqueness of viscosity solutions to (Eλ). This is due to the fact that, without global strict monotonicity of with respect to , there is no comparison principle. This also makes unenforceable Perron’s method, which is the technique customarily employed to prove existence of solutions.
The general issue of existence and uniqueness of such solutions is subsequently addressed in the paper in the linear case under the minimal hypotheses on and that guarantee that the conditions on and mentioned above are in force. We prove that the family of maximal solutions to (Eλ) is well defined and equi-bounded for small enough. When changes sign and in a neighborhood of the projected Aubry set , we show that equation (Eλ) does admit other solutions too that uniformly diverge to as . If we additionally assume on , we furthermore show that any family made up of solutions to (Eλ) that differ from the maximal ones uniformly diverge to . Incidentally, this completely solves the vanishing discount problem for this model case under the sole assumption that on in view of the results established in [41], where was additionally assumed nonnegative on .
Condition (L5) was introduced in [7] and therein employed to solve the vanishing discount problem for an equation of the form (Eλ) under the same set of assumptions considered herein, plus the additional requirement that is globally non-decreasing in . Condition (L5) can be read as a strict monotonicity condition on with respect to , and this is transparent in the linear case . What we find striking about the output of our study is the fact that (L5) is a very weak requirement: it implies that has to be strictly positive only on some portions of the projected Mather set , where the latter is the minimal closed set that contains the projection of the supports of all Mather measures. This set can be very small, such as a finite set of points, see Remark 1.10. Furthermore, it has been conjectured by Mañé [33] that for generic Hamiltonians both and coincide with the support of a closed curve. Many results have been obtained in this direction, see for example [8] and the references therein, showing that condition (L5) generically leaves a lot of space for to take negative values.
The main results proved in this paper keep holding when the superlinearity condition on is relaxed in favor of a simple coercivity. We have decided not to pursue this generalization here since that would add additional technicalities with the drawback of hiding the ideas at the base of our work, see Remark 1.12 for further details.
History of the problem
The so-called ergodic approximation is a technique introduced in [32] to study the existence of solutions of the Hamilton-Jacobi equation111All solutions in the paper are meant in the viscosity sense. The definition will be provided later.
(1) |
on the flat -dimensional torus , where the Hamiltonian is a continuous function on , coercive in the gradient variable, uniformly with respect to , and is a real number. Let and be the unique solution of
According to [32], the functions uniformly converge on , as , to a constant . Furthermore, the solutions are equi-Lipschitz, yielding, by the Arzelá-Ascoli Theorem, that the functions uniformly converge, along subsequences as , to a solution of (1) with equaling . The constant is called critical value of and is characterized by the property of being the unique constant such that (1) admits solutions. At that time, it was not clear if different converging sequences yield the same limit. Some constraints on the possible limit solutions were subsequently found in [22, 29], but the breakthrough came with the work [12], where the authors proved that the unique solution of
() |
converges to a distinguished solution of
(HJ0) |
as under the sole additional assumption that is convex in the gradient variable. The proof relies on techniques and tools issued from weak KAM Theory, in particular on the concept of Mather measure, and it works whenever is a closed Riemannian manifold. This kind of problem is also known as the vanishing discount problem. When the convexity condition on is dropped, the functions may not converge, as it was pointed out in [44] through a counterexample posed on the 1-dimensional torus.
As a nonlinear generalization (see [6, 5, 23] and [39]), one can study the uniform convergence of the unique solution of
as , where is strictly increasing in , and uniformly converges to on compact sets as . This kind of problem is called the vanishing contact structure problem. The vanishing discount problem falls in this framework as a particular case by choosing .
The asymptotic convergence result has been subsequently established in many different situations. For the second order case, one can refer to [26, 27, 34, 43]. For the discrete case, one can refer to [11, 38, 42] and also [2] in the context of twist maps. For the similar problem in the mean field game theory, one can refer to [4]. For the weakly coupled Hamilton-Jacobi systems, one can refer to [17, 14, 24, 25]. For the non-compact setting, one can refer to [18, 28].
A natural and challenging question is to weaken the hypothesis on the monotonicity of the Hamiltonian. A first degenerate case was studied in [41], where the author considered the convergence of the solution of
as , where on , and on the projected Aubry set of . Inspired by the works [5, 41], the authors studied in [7] the vanishing discount problem for contact Hamilton-Jacobi equations of the form (Eλ), where the positivity hypothesis on assumed in [41] is weakened and generalized by introducing the non-degeneracy integral condition (L5). This work highlights once more that the concept of Mather measure plays a central role in the convergence result.
It is worth pointing out that all works mentioned above required a global non-decreasing hypothesis of the Hamiltonians in . If the discounted equation is not increasing in the unknown function , solutions may even not exist, and, if they exist, they may be not unique. In [15, 40], the authors discussed the uniform convergence of the minimal solution of () as . For the non-monotone vanishing discount problem, the second author provided the first example in [35] of nonconvergence. In this example, there exist a convergent family of solutions and a divergent family of solutions at the same time. This phenomenon is new comparing with all the previous works in this direction. In the present paper, we show that the example in [35] is in fact a very general phenomenon when the Hamiltonian is continuous, convex and superlinear in the fibres.
Let us conclude by mentioning that the type of problems we study are also closely linked to optimization problems in economics. The discount factor then models the effect of time through interest rates or inflation. Negative interest rates or deflation have been studied by economists (see for instance [31]). Our results then give possible asymptotics in the presence of coexisting inflation and deflation.
Presentation of our results
We present here our main results. Section 2 contains our analysis on the asymptotic behavior of possible solutions of a general contact Hamilton-Jacobi equation of the form (Eλ) when the discount factor goes to . The Hamiltonian is assumed convex and superlinear in , and globally Lipschitz in , see conditions (G1-3) in Section 2. It is also assumed to satisfy a -type regularity condition in near , see condition (G4) in Section 2. The latter is for instance satisfied when the map is in a neighborhood of in the following sense:
-
(G4′)
there exists such that exists for all and is continuous in .
Let us consider
(Eλ) |
and the limit equation
() |
where is the critical value associated with . We will furthermore assume the non-degeneracy integral condition (L5), where denotes the set of Mather measures for .
Theorem 1.
Under the previous assumptions, there exist a viscosity solution of () and functions , and with
such that, if is a solution of (Eλ) for some , then either one of the following alternatives occurs:
-
(i)
;
-
(ii)
;
-
(iii)
.
We stress that all these three behaviors can happen at the same time for a properly chosen Hamiltonian . Indeed, consider . It is easlily seen that verifies all of the above conditions. The limit Hamiltonian is , its critical constant is and the constant functions are the only solutions to in . Moreover, for all , the three constant functions , and are solutions to the discounted equation.
Theorem 1 shows in great generality that the only possible asymptotic behavior of families of solutions is, up to subsequences, either to uniformly diverge to , or to uniformly converge to a specific solution of (). We also provide two characterizations of , see Theorems 2.2 and 2.3.
The general problem of existence and uniqueness of such solutions is addressed in Section 3. Here we consider the linear case under the minimal hypotheses on and that guarantee that the conditions on and presented above are in force. The discounted equation is then
() |
with limit equation
(HJ0) |
We prove the following existence and convergence result:
Theorem 2.
The previous theorem excludes the possibility of families of solutions that uniformly diverge to , but leaves open the possibility of families uniformly diverging to . By strengthening the non-degeneracy integral condition (L5) with a pointwise positivity condition on on the projected Aubry set , we are able to improve the statement of Theorem 1 as follows.
Theorem 3.
When the previous hypothesis is reinforced, we obtain a stronger conclusion:
Theorem 4.
Namely, in this last case, is the only converging family of solutions.
Let us conclude this presentation by stressing that, combining this analysis with the results of [41], we fully understand the asymptotic behavior of solutions to () when on . The output is the following:
-
(a)
if on the whole , (this situation was considered in [41]) then, for all , there is a unique solution to (), and the family converges as .
- (b)
1. Preliminaries
1.1. Notation
-
Throughout this paper, we assume that is a closed, connected and smooth Riemannian manifold.
-
We fix an auxiliary Riemannian metric on . Let be the distance between and in induced by . By compactness of our results are independent on the choice of .
-
Let diam be the diameter of .
-
We denote by and the tangent and cotangent bundle over respectively. We denote by and points of and respectively.
-
Let denote both canonical projections, the context will make it clear which one is considered.
-
We denote by the norm on both and induced by .
-
If is a smooth manifold, we will denote by the Polish space of continuous functions from to endowed with the metric of local uniform convergence on . We will denote by (resp. ) the set of compactly supported (resp., ) functions from to .
-
We denote by the space of Borel probability measures on endowed with the weak- topology coming from the dual .
-
We denote the set of continuous functions with at most linear growth meaning that
This last quantity defines a norm on the vector space .
-
denotes the set of positive integers.
1.2. Viscosity solutions.
We start by recalling the notion of viscosity solution.
Definition 1.1.
Let be a continuous function, , and consider the equation
(1.1) |
-
(a)
We say that is a viscosity subsolution of (1.1), denoted by
if, for all and such that has a local maximum at , we have . Such a function is termed supertangent to at .
-
(b)
We say that is a viscosity supersolution of (1.1), denoted by
if, for all and such that has a local minimum at , then . Such a function is termed subtangent to at .
-
(c)
We say that is a viscosity solution of (1.1) if it is both a viscosity sub and supersolution.
In this paper, solutions, subsolutions, supersolutions will be always meant in the viscosity sense and implicitly assumed continuous. We recall that, if is on an open set , then it is a viscosity solution (resp. subsolution, supersolution) in if and only if it is a pointwise solution (resp. subsolution, supersolution) in .
The following stability result is well known, see for instance [3].
1.3. Weak KAM solutions and Aubry-Mather theory.
We assume is a continuous Hamiltonian satisfying
-
(H1)
(Convexity) is convex in for all .
-
(H2)
(Superlinearity) .
Let be the convex conjugate function of , i.e.,
It is well known that the Lagrangian is a continuous function on and it is convex and superlinear in . The Fenchel inequality is a direct consequence of this definition:
(1.2) |
Moreover, it can be proven that is itself the convex conjugate of , i.e.,
(1.3) |
Let denote the critical constant defined as follows:
(1.4) |
We present here some facts that we will need about the critical equation, i.e.,
(HJ0) |
Solutions, subsolutions and supersolutions of (HJ0) will be termed critical in the sequel.
Proposition 1.3.
Let . The following properties hold:
- (i)
- (ii)
- (iii)
More precisely, items (ii) and (iii) above require the convexity of in the momentum, while item (i) is a general fact.
Since we are assuming to be superlinear (hence coercive, which is enough), we also have the following characterization of critical subsolutions, see for instance [3, 19].
Proposition 1.4.
For every , we define the minimal action function as
where is taken among all absolutely continuous curves222In the paper, even if not explicitly stated, all curves considered are at least absolutely continuous. satisfying and . The Peierls barrier is the function defined by
(1.5) |
It satisfies the following properties, see for instance [16]:
Proposition 1.5.
-
(i)
The Peierls barrier is finite valued and Lipschitz continuous.
-
(ii)
If is a critical subsolution, then
-
(iii)
For every fixed , the function is a critical solution.
-
(iv)
For every fixed , the function is a critical subsolution.
Theorem 1.6.
There exists a critical subsolution which is both strict and of class C∞ in , meaning that
In particular, the projected Aubry set is nonempty.
The last assertion directly follows from the definition of , see (1.4).
Proposition 1.7.
If and are respectively a sub and supersolution such that on , then on the whole of . In particular, is a uniqueness set for (HJ0), meaning that if two solutions coincide on , then they are equal.
We will say that a Borel probability measure on is closed if it satisfies the following conditions:
-
(a)
;
-
(b)
for all function , we have .
We will denote by the set of such measures.
We will furthermore denote by the family of probability measures that satisfy condition (a) above. The inclusions hold. We will endow with the weak- topology coming from the dual . We refer the reader to [9] for more details on these families of measures.
Theorem 1.8.
The following holds
Measures realizing the above minimum are called Mather measures for . We denote by the set of all Mather measures. This set is compact. The Mather set and the projected Mather set are defined as follows:
These sets are also compact, see [19] for a proof in the regular case.333For the present nonregular case, a proof of this can be found in Appendix A in the ArXiv version of [12]. Furthermore, the following holds, see [41, Proposition 3.13] for a proof in the nonregular case.
Theorem 1.9.
The following inclusion holds: .
Remark 1.10.
In the example of a mechanical Hamiltonian, i.e., , it is well known that , and the Mather measures are convex combinations of delta Diracs concentrated at points with , so that is also equal to .
We conclude this paragraph by a technical lemma that will be of crucial use (see [9, Theorem 2-4.1.3.]).
Lemma 1.11.
Let . The set is compact in .
Note that, as is bounded below, the quantity is well defined for any Borel probability measure. We also remark that in [9, Theorem 2-4.1.3.] the result is proved for a particular subclass of measures, however the proof makes no use of this fact and proves the above result.
1.4. Hamiltonians depending on the unknown function
We recall here known results that can be found in [36, 37] and the references therein. In this section, we consider a continuous Hamiltonian which satisfies the following conditions
-
(G1)
(Lipschitz in ) is -Lipschitz continuous for some , uniformly in ;
-
(G2)
(Convexity in ) is convex for each ;
-
(G3)
(Superlinearity in ) is superlinear for each .
Remark 1.12.
The results of this paper keep holding even when the superlinearity condition (G3) is weakened in favor of a simple coercivity. For instance, Theorems 2.2 and 2.3 can be easily generalized to this setting. Indeed, since we are dealing there with a family of equi-bounded, and hence equi-Lipschitz, solutions, see Lemma 2.6, we could employ the usual trick of modifying outside a compact subset of to make it superlinear. This cannot be done in other parts of the paper since we are dealing with families of solutions that are neither equi-bounded nor equi-Lipschitz in general. And even when they are, as in Section 3.1, this needs to be proved. In fact, this is the core of the analysis performed in Section 3.1, which takes advantage of the fact that the Lagrangian associated with the Hamiltonian via the Fenchel duality is finite-valued. This is no longer true in the purely coercive case, even though the difficulties arising could be handled by showing that all the minimizing curves that come into play in our analysis are indeed supported on the set where the Lagrangian is finite. Yet, we believe that treating this more general case would bring additional technicalities that would have the effect of hiding the ideas at the base of this work. We prefer to leave the coercive case to a possible future work.
Let be the convex conjugate function of , i.e.,
Then it can be proven that verifies similar properties (see [5, Lemma 4.1] with easy adaptations):
-
(L1)
is -Lipschitz continuous uniformly in ;
-
(L2)
is convex for each ;
-
(L3)
is superlinear for each ;
Definition 1.13.
Let be a Hamiltonian satisfying (G1-3) and let be the associated Lagrangian. Let . A function satisfying the following two properties is called a backward (resp. forward) weak KAM solution of
(1.6) |
-
(1)
For each absolutely continuous curve , we have
The above condition reads as is dominated by and will be denoted by .
-
(2)
For each , there exists an absolutely continuous curve (resp. ) with (resp. ) such that
The curves satisfying the above equality are called -calibrated curves.
Lemma 1.14.
We then give an approximation result in this setting. It is an easy consequence of [20, Theorem 8.5].
Theorem 1.15.
Assume is a continuous Hamiltonian verifying (G2). Let be a Lipschitz function verifying for almost every . Then, for every , there is a such that and
Let . Assume there are two points and such that and . Let and consider the equation
(1.7) |
This is a particular case of the previous setting for . In this case, the associated Lagrangian is . The following implicit Lax-Oleinik semigroup is a well defined semigroup of operators that verify, for all ,
(1.8) |
where the infimum is taken among absolutely continuous curves with . A similar property defines and characterizes the forward semigroup , where, for all ,
(1.9) |
Lemma 1.16.
2. General convergence/divergence results
In this section, we consider a continuous Hamiltonian which satisfies the following conditions:
-
(G1)
(Lipschitz in ) is -Lipschitz continuous uniformly in for some ;
-
(G2)
(Convexity in ) is convex for each ;
-
(G3)
(Superlinearity in ) is superlinear for each ;
-
(G4)
(Modulus continuity near ) The partial derivative exists. For every compact subset , we can find a modulus of continuity555A modulus of continuity is a nondecreasing function such that as . such that
The dependence of in the -variable is nonlinear, in general, as in [7], but we do not make any global monotonicity assumption. As established in [5, Lemma 4.1] the associated Lagrangian function defined by
has similar properties:
-
(L1)
(Lipschitz in ) is -Lipschitz continuous uniformly in ;
-
(L2)
(Convexity in ) is convex for each ;
-
(L3)
(Superlinearity in ) is superlinear for each ;
-
(L4)
(Modulus continuity near ) The partial derivative exists. For every compact subset , we can find a modulus of continuity such that
The last condition is easier to state with the Lagrangian function as it involves Mather measures defined on :
-
(L5)
(Non-degeneracy condition) For all Mather measures of , we have
Remark 2.1.
(1) It is classical in convex analysis that for all there is verifying . It is proved in [5, Lemma 4.1] that if and verify the previous formula then . Therefore, an equivalent formulation of (L5) is
-
(G5)
(Non-degeneracy condition) For all Mather measures of ,
where is chosen so to satisfy .
(2) A result of Mañé ([33]) asserts that a generic Hamiltonian has a unique Mather measure.666It has been conjectured by Mañé that for generic Hamiltonians this unique Mather measure is supported on a closed curve. Hence our condition of integral type is quite loose for such a generic . We also refer the reader to Remark 1.10 for an explicit example.
Consider
(Eλ) |
and the limit equation
() |
Here we denote by the critical value of . We denote by (resp. ) the set of all Mather measures (resp. the Mather set) corresponding to . In the sequel, we will always assume that belongs to the interval .
The main theorems of this section are the following.
Theorem 2.2.
Let conditions (L1–5) be in force. Let us assume that there exist an equi-bounded family of solutions to (Eλ), for some . Then the functions uniformly converge in , as , to a solution of the critical equation (). Moreover, is the largest subsolution of () satisfying
(S) |
Under an additional pointwise condition on on the Mather set, the limit critical solution can be characterized as follows.
Theorem 2.3.
Remark 2.4.
Let us stress that the previous theorems hold with same proofs even if is replaced by any family of solutions to (Eλ), where is a subset of having as accumulation point. In particular, this holds for with as .
Concerning the asymptotic behavior of other possible families of solutions to (Eλ), we have the following trichotomy result.
Theorem 2.5.
We start our analysis by remarking that the solutions are equi-Lipschitz continuous. In the remainder of the section, we will denote by the following constant
Lemma 2.6.
The bounded family is equi-Lipschitz continuous.
Proof.
For each , we denote by the distance between them. Take a geodesic satisfying and with constant speed . Denote . Since , we have
The assertion follows by exchanging the role of and . ∎
From the fact that is Lipschitz for every fixed we deduce the following fact.
Lemma 2.7.
Let . Then
Proof.
Pick and choose . By applying Theorem 1.15 to the Hamiltonian and by choosing , we infer that there exists such that for all . By definition of we have that
By integrating this inequality with respect to we get
The result follows letting . ∎
By the Arzelá-Ascoli Theorem and Lemma 2.6, any sequence with admits a subsequence which uniformly converges to a continuous function . By the stability of viscosity solution, see Proposition 1.2, is a solution of (). In the following, we are going to show the uniqueness of the possible limit , thus establishing the convergence result. Define the set of all subsolutions of () satisfying condition (S). We define
(2.1) |
A priori, may be empty, and, even if is not empty, might be . Both these circumstances will be excluded under the hypotheses of Theorem 2.2.
Lemma 2.8.
Any accumulation point of the family as satisfies
In particular, and .
Proof.
Recall that is the uniform bound of . Let . For we have
which implies
The conclusion follows by sending . ∎
In what follows, we will use the notation .
Lemma 2.9.
For and , let be a -calibrated curve with . Then there exists , independent of and of , such that is -Lipschitz continuous for every and .
Proof.
By Lemma 2.6, there is independent of such that is -Lipschitz continuous. By superlinearity of , for each , there is such that
(2.2) |
Thus, we have for
which implies
The proof is now complete. ∎
From now on, we denote by the calibrated curve considered in Lemma 2.9. By the compactness of , there are two constants and with
(2.3) |
We derive from this the following asymptotic informations on the calibrated curves , cf. [7, Corollary 7.4].
Lemma 2.10.
There exist and such that, for all and for all , we have
(2.4) |
for all with .
Proof.
Let us prove the left-hand inequality in (2.4). We argue by contradiction. Assume there is a sequence and a sequence such that
(2.5) |
Define for all a probability measure on by
Here is the set of all continuous functions defined on with compact supports. By Lemma 2.9, all the measures have support in the compact set
Up to extracting a subsequence if necessary, we can assume that converges to in the weak- topology on . Note that is also compactly supported. For , we have
as , which implies that is closed. Since is a calibrated curve, we have
Since is bounded, letting , we find that
Therefore, the limit is a Mather measure of . By (2.5), we get
which contradicts (2.3). The right-hand side inequality in (2.4) can be proved similarly. ∎
In the following, we will denote by and the constants given by Lemma 2.10.
Lemma 2.11.
Let and . The following holds:
-
(i)
for any , we have
As a consequence,
-
(ii)
For any , we have
In particular,
(2.6)
Proof.
We proceed by associating to each calibrated curve a probability measure on . These probability measures will play a key role in the proof of the convergence result stated in Theorem 2.2. They can be regarded as a generalization to the case at issue of the analogous measures first introduced in [11, formula (3.5)]. They already appeared in this exact form in [5, 41, 7].
Definition 2.12.
The following holds, cf. [11, Proposition 3.6], [5, Proposition 4.5], [41, Proposition 5.8], [7, Proposition 7.5].
Lemma 2.13.
The family has support contained in a common compact subset of , in particular it is relatively compact in . Furthermore, if in for , then is a Mather measure.
Proof.
The first part is a direct consequence of Lemma 2.9. It remains to prove that the limit is a Mather measure. We first prove that is closed. For , we have, by integrating by parts,
By (L1) and , we have
where, for the last inequality, we have used (2.6). By (2.6) again, we have
as .
We then prove that is minimizing. Since is Lipschitz continuous, and is a -calibrated curve, for a.e. we have
Then
where
Similarly to the first part of the proof, we have
as , where is a uniform bound on the .
Finally, to bound the error term we use (L1) to find
Therefore, as along the sequence , we conclude that is a Mather measure.∎
The following lemma will be crucial for the proof of the convergence result, cf. [11, Lemma 3.7], [5, Lemma 4.7], [7, Lemma 7.7].
Proof.
For , using Theorem 1.15 we take such that and
Using the Fenchel inequality we have for all ,
Since is Lipschitz continuous, and is a -calibrated curve, for a.e. we have
(2.7) | ||||
where
Let us estimate the error term . Let us set , where is the Lipschitz constant of the curves , according to Lemma 2.9. By (L4) we have
(2.8) |
where is a uniform bound on the . By multiplying both sides of (2.7) by and by rearranging terms, we obtain, for a.e. ,
Integrating the above inequality over the interval where as stated in Lemma 2.11, and using an integration by parts, we have
Letting it follows that,
From Lemma 2.11 we infer that the maps
are in and converge to as . By taking also into account (2.7), we can send in the above inequality, to get, by the Dominated Convergence Theorem,
By Definition 2.12, we have
According to Lemma 2.11-(i), for , we derive that
By (2.7) we also have
The assertion follows by sending . ∎
We are now in position to prove the first two main theorems of this section.
Proof of Theorem 2.2.
Let be an accumulation point of the as . By Lemma 2.8, we know that , so in . Let us prove that . Fix and pick . From Lemmas 2.13 and 2.14 we infer that
for some Mather measure . Since satisfies the constraint (S), we get from this that , hence . By the arbitrariness of the choice of , we get that is finite-valued and that in . We conclude that is the unique accumulation point of the family of functions as . The proof is complete. ∎
Proof of Theorem 2.3.
We denote
Note that is finite-valued, as . We start by remarking that is a subsolution of (). Indeed, for every fixed , the function is a convex combination of the family of critical solutions , where . By the convexity of in the momentum and the equi-Lipschitz character of the critical subsolutions, see Propositions 1.3 and 1.4, it follows that each is a critical subsolution. By Proposition 1.3 again, we infer that a finite valued infimum of critical subsolutions is itself a critical subsolution. Therefore is a critical subsolution.
Let us now prove that . By Proposition 1.4 we know that for all . Let us integrate this inequality with respect to a Mather measure . By assumption and by definition of Mather set, we have for all . We infer
where, for the last inequality, we used that satisfies (S). From this we get
By taking the inf with respect to of the right-hand side term in the above inequality, we get .
We proceed by proving the third main theorem of this section, namely the trichotomy result stated in Theorem 2.5. We start by establishing a sort of Harnack-type inequality for subsolutions of (Eλ).
Lemma 2.15.
There exists a constant such that, if and is a subsolution to (Eλ), then
(2.9) |
Proof.
The left hand side inequality is obvious. Let us prove the right hand side inequality. Let such that . Let such that . Denote . Let be a geodesic satisfying and with constant speed 1. By and (L1), we get
where . By the Gronwall inequality we infer
Taking , and recalling that ,we have
where diam. The result follows taking . ∎
As a consequence, we derive the following key proposition. It will be also used in Section 3 to show the existence of diverging families of solutions.
Proposition 2.16.
Let be a subset of having as accumulation point. Let be a family of subsolutions of (Eλ).
-
(i)
If, for each , there is a point such that as , , then uniformly converges to as , .
-
(ii)
If, for each , there is a point such that as , , then uniformly converges to as , .
Proof.
As in the previous proof, we denote and .
Let us prove assertion (i). The hypothesis is equivalent to as , . Then restricting to , (2.9) implies that hence as , .
Let us prove item (ii). The hypothesis is equivalent to as , . Then restricting to , (2.9) implies that hence as , . ∎
We are now in position to prove Theorem 2.5. We recall that, in what follows, still denotes the critical solution provided by Theorem 2.2
Proof of Theorem 2.5.
We claim that . Otherwise, there is , sequences , and such that
(2.10) |
We claim that is uniformly bounded from below. If not, there is such that
By Proposition 2.16, uniformly converges to , which contradicts (2.10). Then, according to Theorem 2.2 and Remark 2.4, the sequence uniformly converges. to the only possible limit , which contradicts the fact that for all .
In a similar manner, define
and
If , we set . The same proof yields that .
Define
We claim that . Otherwise, there is , a sequence of discount factors and of solutions and such that
(2.11) |
Since , there is also a point such that
Similarly to the first step, we infer from Proposition 2.16 that is uniformly bounded. Then, according to Theorem 2.2 and Remark 2.4, the sequence uniformly converges to the only possible limit . But again this yields a contradiction as for all . This concludes the proof. ∎
3. The linear case
In this section we continue our analysis on the vanishing discount problem in the case when depends linearly on . Hence we will consider a Hamilton-Jacobi equation with discount factor of the form
() |
where we assume that is a continuous function satisfying (H1)-(H2) (convexity and superlinearity) and the coefficient is a continuous function on satisfying the following condition:
() |
where denotes the compact and convex subset of made up by Mather measures associated with . Without any loss of generality, we shall restrict to the case . Equations of the form () can be regarded as a model example for the more general Hamilton-Jacobi equations of contact type considered in Section 2, cf. equation (Eλ). The full Hamiltonian is then given by . The Hamiltonian then fulfills all the hypotheses of the previous section. In particular, it is Lipschitz with respect to with Lipschitz constant .
We will be concerned with the issue of existence of solutions to equation (), at least for small values of , and their asymptotic behavior as . In Section 3.1 we show the existence of the maximal solution to () for values of small enough. These solutions are shown to be equi-bounded and equi-Lispchitz, therefore, in view of the results established in Section 2, they converge to a solution of the limit critical equation
(HJ0) |
In Section 3.2, under further natural conditions on the coefficient , we investigate the existence of possible other families of solutions of () and we describe their asymptotic behavior as .
3.1. A converging family of solutions
We recall that is the Lagrangian associated with via the Legendre transform, namely
The constant is the critical value of . We define the following value function, for ,
where the infimum is taken among absolutely continuous curves with . This definition is inspired by the formula given in [41, Theorem 4.8] to represent the unique solution of () in the case in and on .
The following holds.
Theorem 3.1.
There exists such that for every the following holds:
-
(i)
the value function is finite-valued;
-
(ii)
the functions are equi-bounded and equi-Lipschitz;
- (iii)
-
(iv)
for every , there exists a curve such that
Furthermore, the curve is -Lipschitz, for some independent of and .
Remark 3.2.
According to Theorem 2.2, the functions uniformly converge to a critical solution as , where is the maximal subsolution of (HJ0) satisfying
We can furthermore strengthen the conclusion of Theorem 2.5 on the asymptotic behavior, as , of all possible families of solutions to equation (). Theorem 3.1 in fact rules out the possibility that there exist families of solutions that diverge to . The precise statement is the following.
The remainder of this subsection is devoted to prove the statement of Theorem 3.1. This will be obtained via a series of intermediate results.
A key step in order to establish that the value function is a (Lispchitz) viscosity subsolution to (HJ0) is to prove that it satisfies the Dynamic Programming Principle. Please note that the next proposition is valid even when is not finite-valued.
Proposition 3.4.
(Dynamic Programming Principle). Let . For each absolutely continuous curve , we have
Proof.
We start by claiming that we can reduce to the case , without any loss of generality. Indeed, by multiplying the above inequality for , we get
The claim follows by making the change of variables , and by replacing with .
Let us then prove the assertion with . For each with , we define
Then, by definition of the value function, we have
Taking the infimum among all , we get the assertion. ∎
We now proceed to show is a bounded function on , at least for small enough. We start by proving the following upper bound.
Lemma 3.5.
There is a constant such that for all and .
Proof.
Now we know that the infimum in is taken among curves in
For and , we define
and
Since , is increasing. Then exists for each and , and may equal . We will also need the following auxiliary remark.
Lemma 3.6.
For every and for all curves , we have
Proof.
A direct calculation shows that
By the Cauchy mean value theorem, there is such that
which implies the conclusion. ∎
We distill in the next lemma an argument that we will repeatedly use in the sequel.
Lemma 3.7.
Assume there are sequences , , and such that
(3.1) | ||||
(3.2) |
Define a probability measure by
() |
Then the set is relatively compact in (for the weak- topology coming from ) and any of its accumulation points is a Mather measure associated with .
Proof.
Note that, due to (3.1), we have as . Since is uniformly superlinear in the fibers, there exists a constant such that
which readily implies that the sequence is a well defined sequence in . The asserted precompactness of in follows from Lemma 1.11.
Let be a limit point of a subsequence of , that we will not relabel to ease notations. We are going to show that is a Mather measure.
The measure is closed: we pick . An integration by parts shows that
(3.3) | |||
We infer
By sending and by using Lemma 3.6, hypothesis (3.1) and again Lemma 1.11, we derive
The measure is minimizing: by (3.2) we get
where the last inequality follows from the fact that is continuous and bounded from below and the measures are weakly- converging to in (hence, narrowly converging), see for instance [1, Section 5.1.1].
In view of Theorem 1.8, we conclude that is a Mather measure. ∎
Let be a fixed constant and define, for any fixed integer ,
The information provided by the next lemma will be crucial for our upcoming analysis.
Lemma 3.8.
Let be a fixed integer and assume that for every small enough. Then there exists such that
Remark 3.9.
Note that for every integer and . In particular, when and , we have for every in view of Lemma 3.5.
Proof.
Let be a fixed integer. We argue by contradiction. Assume there exist sequences , and such that
with
In particular, conditions (3.1) and (3.2) in Lemma 3.7 are satisfied with
Let be the probability measure defined in ( ‣ 3.7). According to Lemma 3.7, up to extraction of a subsequence (not relabeled), the measures weakly converge to a Mather measure .
Now we choose a subsolution of (HJ0). For every , there exists, in view of Theorem 1.15, a function satisfying
(3.4) |
For small, we have . We get
The equality appearing above is obtained via an integration by parts (cf. proof of Lemma 3.7 equation (3.3), when we prove that is closed), while for the last inequality we have used that fact that . Now we send and divide the above inequality by . We get
Since as , we get
Since for all the function is also a negative subsolution of (HJ0), by replacing with in the inequality above we obtain
We now proceed to show that the value function is bounded from below, for every fixed , where is the value obtained according to Lemma 3.7 with and , where is the constant provided by Lemma 3.5.
Lemma 3.10.
There exists a constant independent of such that
Proof.
Let us fix . For every we have
where . We are now going to apply Lemma 3.8 with and by choosing in the definition of , so that : from the above inequality we infer
The assertion readily follows from the definition of . ∎
From the information gathered so far, we know that the value function is finite-valued on for every . We now proceed to get bounds for from above and from below on independent of .
Lemma 3.11.
There is a constant such that
Proof.
According to [13, Theorem 4.14 and Proposition 4.4], see also [16, Theorem 3.3], for every there exists a curve with such that, for every subsolution of (HJ0),
(3.5) |
Let us denote by the family of curves satisfying (3.5). We remark for further use that these curves are equi-Lipschitz, see for instance [13, Lemma 4.9].
Pick a subsolution of (HJ0) and fix . From the definition of we get
where is defined by
The second equality in (3.1) is derived via an integration by parts as for (3.3). By assumption () and compactness of the family of Mather measures , there exists such that
(3.7) |
We show that there is such that, for all curves , we have
We argue by contradiction. Assume there exist sequences and such that
(3.8) |
Define by
Due to the fact that the curves are equi-Lipschitz, the measures have equi-compact support. In particular, up to extracting a subsequence (not relabeled), they weakly converge to a probability measure . Furthermore
We claim that is closed. Indeed, for every we have
We proceed to show that is minimizing, namely, a Mather measure. Pick a subsolution to (HJ0). By exploiting (3.5), we get
By (3.8), we also have which leads to a contradiction with (3.7). Then for we have
and
By using these inequalities in (3.1) and recalling that , we finally get
which gives the upper bound of on independent of . ∎
By exploiting the fact that satisfies the Dynamic Programming Principle, we show that the partial upper bound obtained in Lemma 3.11 actually entails a uniform upper bound on the whole .
Proposition 3.12.
There is independent of such that
Proof.
Fix and pick a point . Set , diam and . Take a geodesic with , and . By Proposition 3.4 we have
which gives the sought uniform upper bound of . ∎
Now that we know that is uniformly bounded from above, we can prove that is also uniformly bounded from below.
Proposition 3.13.
There exist and a constant independent of such that
In particular, for all and .
Proof.
By Proposition 3.12, we know that, for , the infimum in is taken among the set
By Lemma 3.8, there exists (depending on ) such that
Let us set .777We recall that is the value obtained according to Lemma 3.7 with and , where is the constant provided by Lemma 3.5. Let be a subsolution of (HJ0). For each , we can take satisfying and for all , given by Theorem 1.15. For small, we have . For each , we have
where
We recall that . Now let to get
Sending we get
The bound from below readily follows from this by definition of . The last assertion is a consequence of Proposition 3.12 and of what we have just shown above. ∎
We now proceed to show that the value function is Lipschitz continuous. This is indeed a consequence of this more general result.
Proposition 3.14.
Let be a bounded function. Assume that satisfies the Dynamic Programming Principle, i.e., for each absolutely continuous curve , we have
(DPP) |
Then is Lipschitz continuous, with a Lipschitz constant that only depends on and .
Proof.
Pick . Let be a geodesic with and , where . By (DPP) we have
There is such that . If , we have
If , we have
Then
Since and , there is such that . Exchanging the role of and , we get the conclusion. ∎
As a consequence of the previous proposition and Proposition 3.13, we derive the following information.
Corollary 3.15.
There is independent of such that the functions are -Lipschitz continuous.
Next, we show that the value function is a viscosity subsolution of the equation (). Indeed, the following result holds.
Proposition 3.16.
Let . Then is a subsolution of () if and only if (DPP) holds for every absolutely curve and every .
Proof.
Let us first assume that is a viscosity subsolution of (HJ0). By Proposition 1.4, we derive that is Lipschitz continuous. Using Theorem 1.15, we take a sequence such that and
For , we have, performing the now usual integration by parts and the Fenchel inequality (1.2),
The assertion follows by sending .
Conversely, let us assume that satisfies (DPP). According to Proposition 3.14, is Lipschitz continuous. We only need to check if is a subsolution where is differentiable, thanks to Proposition 1.4. Let be differentiable at . We take a curve with and . Then
Letting , we get
Taking the supremum with respect to , we get, thanks to (1.3),
which implies that is a subsolution. ∎
We proceed to show that the value function is the maximal viscosity subsolution of (), and hence a solution by maximality. We need an auxiliary lemma first.
Lemma 3.17.
Let and let be an absolutely continuous curve. Let us assume that
(3.9) |
Then as .
Proof.
We argue by contradiction. Assume there exist an increasing sequence and a small enough such that
Then, for all , we have
hence
Let us pick with . Up to extracting a subsequence, we can assume that for all . Let us set . We have
as , which contradicts (3.9). ∎
Let us prove the result announced above. We recall that the appearing in the next two statements is the real number in provided by Proposition 3.13.
Proposition 3.18.
Proof.
Let be a subsolution of (). Let us fix . By definition, for all , there is such that
where for . By Proposition 3.16, we have
By the definition of , we have
Combining all the above inequalities, we conclude
We know that , cf. proof of Proposition 3.13. By Lemma 3.17, we derive that as . By finally sending we conclude that . This, together with Propositions 3.4 and 3.16, proves the asserted maximality of .
Now we show that is a viscosity solution of (). Since is the maximal subsolution, we only need to show that it is a supersolution. The argument is standard and depends on the bump construction. We give it below for the reader’s convenience. Assume, by contradiction, that the supersolution test fails at some point . This means that there is a strict subtangent of at 888Meaning that has a strict local minimum at . such that
Up to adding a constant to , we can assume that . Let be the open ball centered at with the radius . Let us choose and small enough so that
(3.10) |
Set
Due to (3.10), the function is a subsolution of () in as the maximum of two subsolutions in that open ball, and it agrees with in an open neighborhood of . This readily implies that is a subsolution of () on the whole . Yet, we have , contradicting the fact that is the maximal subsolution of (). This shows that is indeed a solution to () in . ∎
From now on, we denote by the value function , since it is the maximal solution.
Proposition 3.19.
There exists such that, for every fixed and , we can find a curve with such that
Furthermore, the curve is -Lipschitz continuous, for some constant independent of and .
Proof.
Let us fix . According to Proposition 3.18, is a viscosity, hence a weak KAM, solution of
By Lemma 2.9, we know that, for every fixed , there is a Lipschitz curve with such that
(3.11) |
for all . The above equality with and can be restated as
(3.12) |
where
Note that the function and are in . Since is uniformly bounded from below, this implies, in view of (3.12), that is in . The operator defined by
is a contraction when , in particular there is a unique fixed point of . This gives the existence of the local unique solution of (3.12), which is defined, for example, on with . Since the time of local existence is independent of the initial data, the maximal solution is defined for all , and is given by
as an explicit computation shows. This gives directly
(3.13) |
for all . Due to the fact that the functions are equi-bounded and equi-Lipschitz, this implies that the curve is -Lipschitz, with a Lipschitz constant that is independent of and , see Lemma 2.9.
We want to show that there exists such that as whenever . According to Lemma 3.17, it suffices to show that there exists a such that
(3.14) |
where we have denoted by the family of absolutely continuous curves with that satisfy (3.11). Notice in fact that (3.14) implies in particular that, for every fixed , condition (3.9) in Lemma 3.17 is met. We argue by contradiction. Let us assume the claim false. Then there exist sequences , , and such that as . Notice that the latter implies that and as . Let be the probability measure defined in ( ‣ 3.7). Then (3.13) can be restated as
(3.15) |
for all . With the aid of Lemma 3.6, it is easy to check that the right-hand side of (3.15) goes to as . We can therefore apply Lemma 3.7 to infer that, up to extracting a subsequence, weakly converges to a Mather measure .
Now we choose a subsolution of (HJ0). For , we take satisfying and for all , given by Theorem 1.15. For small enough, we have . By (3.15) we get
where for the first inequality we have used Fenchel’s inequality (1.2) and for the subsequent equality an integration by parts. Let to get
Now we divide by , we use the fact that and Lemma 3.6 to get
Since for, all , is also a subsolution of (HJ0), we get, by applying the previous inequality to , that
which implies that . This leads to a contradiction with () being a Mather measure, as we recalled above.
Let us fix and . By sending in (3.13) and by exploiting the information just gathered, we derive that
We conclude that
(3.16) | ||||
Indeed (3.14) means that the positive function is in . Furthermore, the fact that the curve is Lipschitz implies that the function is in . Equality (3.16) follows from this by making use of the Dominated Convergence Theorem. ∎
3.2. Diverging families of solutions
We have seen in Theorem 3.1 that, under the assumption (), the maximal solution of () uniformly converges as to a critical solution . In this section, we will show that if we furthermore assume
-
(2)
there is a point such that ;
-
(3)
in an open neighborhood of the projected Aubry set ,
then there exists a family of solutions to () that uniformly diverges to . Furthermore, when condition (a) is reinforced in favor of the following one
-
(3′)
on ,
then all solutions to () different from the maximal ones uniformly diverge to as . We summarize all this in the following statement. We refer the reader to Section 1.3 for the definition of the projected Aubry set. We underline that, in view of Theorem 1.9, condition (a) is stronger than the integral condition (). Throughout this section, we will denote by the maximal solution of ().
Theorem 3.20.
Proof.
By Theorem 1.6, there is a subsolution of (HJ0) which is and strict in . Therefore, for every open neighborhood of there is a depending on such that
Let us prove (i). Let be the open neighborhood of given by condition (3). Define
For small enough, we have for all . Up to choosing a smaller if necessary, we have
and
for all . Therefore, is a subsolution of (). Now we denote by (resp. ) the Lax-Oleinik semigroup defined in (1.8) (resp. the forward semigroup defined in (1.9)) associated to . Define
By Lemma 1.16, , in particular uniformly diverges to as . By Lemma 1.17, is a solution of (), and there is a point such that for each . We derive that as . By Proposition 2.16, we conclude that uniformly diverges to as .
Let us prove (ii). From assumption (3′) we infer that there is an open neighborhood of and such that for all . One can easily check that there exist small enough and such that
and
for all . Therefore, is a strict subsolution of ().
Claim:
Let . There is no solution of () satisfying
in .
We argue by contradiction. Assume there is such a solution . We have
Since is a strict subsolution of (), by Lemma 1.17,
We get
which contradicts the fact that is a fixed point of .
Therefore, for each solution of (), there is a point such that
By Proposition 2.16, uniformly converges to as . ∎
Remark 3.21.
We note that the proof above also shows that the solutions diverge to with speed (at least) of order .
Acknowledgements
Andrea Davini is a member of the INdAM Research Group GNAMPA. He is supported for this research by Sapienza Università di Roma - Research Funds 2021. Panrui Ni and Maxime Zavidovique are supported by ANR CoSyDy (ANR-CE40-0014). Jun Yan is supported by National Natural Science Foundation of China (Grant Nos. 12171096, 12231010).
Declarations
Conflict of interest statement: The author states that there is no conflict of interest.
Data availability statement: Data sharing not applicable to this article as no datasets were generated or analysed during the current study.
References
- [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, 2005.
- [2] M.-C. Arnaud and M. Zavidovique, Weak K.A.M. solutions and minimizing orbits of twist maps, Trans. Amer. Math. Soc., 376 (2023), pp. 8129–8171.
- [3] G. Barles, Solutions de viscosité des équations de Hamilton-Jacobi, vol. 17 of Mathématiques & Applications (Berlin) [Mathematics & Applications], Springer-Verlag, Paris, 1994.
- [4] P. Cardaliaguet and A. Porretta, Long time behavior of the master equation in mean field game theory, Analysis & PDE, 12 (2019), pp. 1397–1453.
- [5] Q. Chen, Convergence of solutions of Hamilton–Jacobi equations depending nonlinearly on the unknown function, Advances in Calculus of Variations, (2021).
- [6] Q. Chen, W. Cheng, H. Ishii, and K. Zhao, Vanishing contact structure problem and convergence of the viscosity solutions, Comm. Partial Differential Equations, 44 (2019), pp. 801–836.
- [7] Q. Chen, A. Fathi, M. Zavidovique, and J. Zhang, Convergence of the solutions of the nonlinear discounted Hamilton–Jacobi equation: The central role of Mather measures, Journal de Mathématiques Pures et Appliquées, 181 (2024), pp. 22–57.
- [8] G. Contreras, A. Figalli, and L. Rifford, Generic hyperbolicity of Aubry sets on surfaces, Invent. Math., 200 (2015), pp. 201–261.
- [9] G. Contreras and R. Iturriaga, Global minimizers of autonomous Lagrangians, 22o Colóquio Brasileiro de Matemática. [22nd Brazilian Mathematics Colloquium], Instituto de Matemática Pura e Aplicada (IMPA), Rio de Janeiro, 1999.
- [10] G. Contreras, R. Iturriaga, G. P. Paternain, and M. Paternain, Lagrangian graphs, minimizing measures and Mañé’s critical values, Geom. Funct. Anal., 8 (1998), pp. 788–809.
- [11] A. Davini, A. Fathi, R. Iturriaga, and M. Zavidovique, Convergence of the solutions of the discounted equation: the discrete case, Math. Z., 284 (2016), pp. 1021–1034.
- [12] , Convergence of the solutions of the discounted Hamilton-Jacobi equation, Invent. Math., 206 (2016), pp. 29–55.
- [13] A. Davini and A. Siconolfi, A generalized dynamical approach to the large time behavior of solutions of Hamilton-Jacobi equations, SIAM J. Math. Anal., 38 (2006), pp. 478–502.
- [14] A. Davini, A. Siconolfi, and M. Zavidovique, Random Lax-Oleinik semigroups for Hamilton-Jacobi systems, J. Math. Pures Appl. (9), 120 (2018), pp. 294–333.
- [15] A. Davini and L. Wang, On the vanishing discount problem from the negative direction, Discrete Contin. Dyn. Syst., 41 (2021), pp. 2377–2389.
- [16] A. Davini and M. Zavidovique, Weak KAM theory for nonregular commuting Hamiltonians, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), pp. 57–94.
- [17] , Convergence of the solutions of discounted Hamilton-Jacobi systems, Advances in Calculus of Variations, 14 (2019), pp. 1–15.
- [18] I. C. Dolcetta and A. Davini, On the vanishing discount approximation for compactly supported perturbations of periodic Hamiltonians: the 1d case, Comm. Partial Differential Equations, 48 (2023), p. 576–622.
- [19] A. Fathi, Weak KAM from a PDE point of view: viscosity solutions of the Hamilton-Jacobi equation and Aubry set, Proc. Roy. Soc. Edinburgh Sect. A, 142 (2012), pp. 1193–1236.
- [20] A. Fathi and E. Maderna, Weak KAM theorem on non compact manifolds, NoDEA Nonlinear Differential Equations Appl., 14 (2007), pp. 1–27.
- [21] A. Fathi and A. Siconolfi, PDE aspects of Aubry-Mather theory for quasiconvex Hamiltonians, Calc. Var. Partial Differential Equations, 22 (2005), pp. 185–228.
- [22] D. A. Gomes, Generalized Mather problem and selection principles for viscosity solutions and Mather measures, Adv. Calc. Var., 1 (2008), pp. 291–307.
- [23] D. A. Gomes, H. Mitake, and H. V. Tran, The selection problem for discounted Hamilton-Jacobi equations: some non-convex cases, J. Math. Soc. Japan, 70 (2018), pp. 345–364.
- [24] H. Ishii, The vanishing discount problem for monotone systems of Hamilton-Jacobi equations. Part 1: linear coupling, Mathematics in Engineering, 3 (2021), pp. 1–21.
- [25] H. Ishii and L. Jin, The vanishing discount problem for monotone systems of Hamilton-Jacobi equations. Part 2: nonlinear coupling, Calc. Var., 59 (2020), pp. 1–28.
- [26] H. Ishii, H. Mitake, and H. Tran, The vanishing discount problem and viscosity Mather measures. Part 1: The problem on a torus, J. Math. Pures Appl., 108 (2017), p. 125–149.
- [27] , The vanishing discount problem and viscosity Mather measures. part 2: Boundary value problems, J. Math. Pures Appl., 108 (2017), pp. 261–305.
- [28] H. Ishii and A. Siconolfi, The vanishing discount problem for Hamilton-Jacobi equations in the Euclidean space, Comm. Partial Differential Equations, 45 (2020), pp. 525–560.
- [29] R. Iturriaga and H. Sánchez-Morgado, Limit of the infinite horizon discounted Hamilton-Jacobi equation, Discrete Contin. Dyn. Syst. Ser. B, 15 (2011), pp. 623–635.
- [30] L. Jin, J. Yan, and K. Zhao, Nonlinear semigroup approach to the Hamilton-Jacobi equation—a toy model, Minimax Theory Appl., 8 (2023), pp. 61–84.
- [31] N. R. Kocherlakota, On the ‘discount’ factor in growth economies, Journal of Monetary Economics, 25 (1990), pp. 43–47.
- [32] P.-L. Lions, G. Papanicolaou, and S. Varadhan, Homogenization of Hamilton-Jacobi equation. unpublished preprint, 1987.
- [33] R. Mañé, Generic properties and problems of minimizing measures of Lagrangian systems, Nonlinearity, 9 (1996), pp. 273–310.
- [34] H. Mitake and H. Tran, Selection problems for a discount degenerate viscous Hamilton-Jacobi equation, Adv. Math., 306 (2017), pp. 684–703.
- [35] P. Ni, Multiple asymptotic behaviors of solutions in the generalized vanishing discount problem, Proc. Amer. Math. Soc., 151 (2023), pp. 5239–5250.
- [36] P. Ni and L. Wang, A nonlinear semigroup approach to Hamilton-Jacobi equations—revisited, J. Differential Equations, 403 (2024), pp. 272–307.
- [37] P. Ni, L. Wang, and J. Yan, A representation formula of the viscosity solution of the contact Hamilton-Jacobi equation and its applications, Chinese Ann. Math. Ser. B, (to appear).
- [38] P. Ni and M. Zavidovique, Nonlinear and degenerate discounted approximation in discrete weak KAM theory, 2024.
- [39] Y. Wang, J. Yan, and J. Zhang, Convergence of viscosity solutions of generalized contact Hamilton-Jacobi equations, Arch. Rational Mech. Anal., 241 (2021), pp. 1–18.
- [40] , On the negative limit of viscosity solutions for discounted Hamilton-Jacobi equations, J. Dyn. Differ. Equ., (2022).
- [41] M. Zavidovique, Convergence of solutions for some degenerate discounted Hamilton–Jacobi equations, Analysis & PDE, 15 (2022), pp. 1287 – 1311.
- [42] , Discrete and Continuous Weak KAM Theory: an introduction through examples and its applications to twist maps, 2023.
- [43] J. Zhang, Limit of solutions for semilinear Hamilton-Jacobi equations with degenerate viscosity, Advances in Calculus of Variations, (2023).
- [44] B. Ziliotto, Convergence of the solutions of the discounted Hamilton-Jacobi equation: A counterexample, J. Math. Pures Appl., 128 (2019), pp. 330–338.