Blow-up Behaviors of Ground States in Ergodic Mean-field Games Systems with Hartree-type Coupling
Abstract
In this paper, we investigate the concentration behaviors of ground states to stationary Mean-field Games systems (MFGs) with the nonlocal coupling in , With the mass critical exponent imposed on Riesz potentials, we first discuss the existence of ground states to potential-free MFGs, which corresponds to the establishment of Gagliardo-Nirenberg type’s inequality. Next, with the aid of the optimal inequality, we classify the existence of ground states to stationary MFGs with Hartree-type coupling in terms of the -norm of population density defined by . In addition, under certain types of coercive potentials, the asymptotics of ground states to ergodic MFGs with the nonlocal coupling are captured. Moreover, if the local polynomial expansions are imposed on potentials, we study the refined asymptotic behaviors of ground states and show that they concentrate on the flattest minima of potentials.
MSC: 35J47, 35J50, 46N10
Keywords: Mean-field Games, Variational Method, Nonlocal Coupling, Ground States, Blow-up Profiles
1 Introduction
In this paper, we are concerned with the following ergodic stationary Mean-field Games systems
(1.4) |
where denotes a solution, is a so-called Lagrange multiplier, is the potential function and is defined as the Riesz potential satisfying
(1.5) |
Here represents the population density and is the value function of a typical player. In particular, Hamiltonian is in general assumed to be convex uniformly and the typical form is
(1.6) |
Correspondingly, the Lagrangian is defined by and if is given by (1.6), can be written as
(1.7) |
where is the conjugate number of
Assume in system (1.4) is given by (1.6) and has polynomial lower and upper bounds when is large enough, then Cesaroni and Bernardini [5, 4] studied the existence and concentration of ground states to (1.4) under the subcritical mass exponent case by using the variational method. Motivated by their results and our analysis focused on Mean-field Games systems with the local coupling [14], we shall utilize the variational approach to discuss the existence and asymptotic behaviors of ground states to (1.4) under the critical mass exponent case, i.e. in (1.5).
1.1 Mean-field Games Theory and Systems
Motivated by the theories of statistical physics, Huang et al. [20] and Lasry et al. [21] in 2007 developed Mean-field Games theories and proposed a class of coupled PDE systems to describe the differential games among a huge number of players, which have rich applications in the fields of economics, finance and management.
The general form of time-dependent Mean-field Games systems reads as
(1.8) |
where and denote the density and the value function, respectively. Here represents the initial data of density and is the terminal data of the value function. Now, we give a brief summary of the derivation of (1.8). Suppose the dynamics of the -th player satisfies
(1.9) |
where is the initial condition, is the velocity and represents the Brownian motion. Assume for are independent and all players are homogeneous, then we have for follow the same process and drop “" in (1.9). On the other hand, each player aims to minimize the following expected cost:
(1.10) |
where is the Lagrangian, measures the spatial preference and is the coupling. Invoking the dynamic programming principle [2, 3], one can formulate the time-dependent system (1.8) by analyzing the minimization of (1.10). We point out that many results are concentrated on the study of global well-posedness to (1.8), see [9, 8, 7, 13, 18, 17, 16].
As stated in [14], the corresponding stationary problem of (1.8) is
(1.11) |
where the triple denotes the solution, is the potential function and is the cost function. There are also some results concerning the existence and qualitative properties of non-trivial solutions to the stationary problem (1.11), see [10, 15, 19, 23, 12, 14, 5, 4]. We mention that when the cost is monotone increasing, as shown in [21], the uniqueness of the solution to (1.11) can be in general guaranteed. Whereas, when the cost is monotone decreasing and unbounded, the case is delicate and (1.11) may admit many distinct solutions. In particular, the pioneering work in the study of ground states to stationary Mean-field Games systems with decreasing cost was finished by Cesaroni and Cirant [10].
We also would like to point out the stationary Mean-field Games systems can be trivialized to nonlinear -Laplacian Schrödinger equations when is chosen as (1.6). Indeed, Fokker-Planck equation in (1.11) can be reduced into the following form:
(1.12) |
Similarly as shown [11], we define and obtain from (1.12) and the -equation in (1.11) that
(1.15) |
where is the -Laplacian and given by . It is well-known that nonlinear -Laplacian Schrödinger equation (1.15) admits the following variational structures:
(1.16) |
where denotes the anti-derivative of In particular, when and in (1.15), the equation is the standard nonlinear Schrödinger equation with the Hartree-type aggregation term.
Inspired by the relation between Schrödinger equations and Mean-field Games systems discussed above, furthermore, the results of Cirant et al. [14] and Bernardini et al. [4, 5], we focus on the existence and asymptotic behaviors of ground states to (1.4) when In particular, Bernardini and Cesaroni studied the subcritical mass exponent case with extensively via the variational method. It is well-known that system (1.4) admits the following variational structure:
(1.17) |
where for and for Here Lagrangian is defined by
(1.21) |
To explain the range of exponent , we are concerned with the following constrained minimization problem:
(1.22) |
where the admissible set is given by
(1.23) |
with
(1.24) |
It is straightforward to show that Indeed, by choosing with determined by and , then one has and which implies Now, we mention that the lower bound is a necessary condition to guarantee that for all . To clarify this, we find if for any
where is defined as and is chosen such that Based on the discussion stated above, Bernardini and Cesaroni employed the direct method and the concentration-compactness approach to investigate the ground states to (1.4) with given by (1.6) when satisfies in (1.5). In this paper, similarly as the work finished in [14], we shall study the existence and blow-up behaviors of ground states to (1.4) under the critical mass exponent case. We also would like to mention that there exists the other critical exponent from the restriction of Sobolev embedding Theorem. Next, we state our main results in Subsection 1.2.
1.2 Main results
We consider Hamiltonian satisfies (1.6) and in (1.11), where is the Riesz potential of order defined by . In particular, we assume potential is locally Hölder continuous and satisfies
-
(V1).
.
-
(V2).
there exist positive constants and such that
(1.25a) (1.25b) (1.25c) -
(V3).
with .
With the assumptions shown above, we shall classify the existence of ground states to (1.4) with in terms of the total mass of density via the variational method. Compared to the arguments for the existence of ground states to the Mean-field Games system with a local coupling, one has to control in some space with the aid of the nonlinearity in (1.17). Motivated by this, we exploit Hardy-Littlewood-Sobolev inequality stated in Appendix A and establish desired estimates.
One of our main goals is to study the attainability of the constrained minimization problem (1.22)
with the critical mass exponent , namely,
(1.26) |
where is given in (1.1) and the energy (1.17) is precisely written as
(1.27) |
and
To this end, similarly as discussed in [14], we have to first investigate the Gagliardo-Nirenberg type’s inequality up to the critical mass exponent, which is
(1.28) |
where
(1.29) |
with defined as (1.24) and . It is worthy mentioning that problem (1.28) is scaling invariant under the scaling for any and
With the help of the conclusions shown in [4], we can prove the existence of minimizers to (1.28) for any . Then, we perform the approximation argument to study the case of . In fact, we have
Theorem 1.1.
Suppose in (1.28) and , then we have is finite and attained by some minimizer . Moreover, we have there exists a classical solution , to the following Mean-field Games systems:
(1.33) |
where
(1.34) |
In particular, there exists constants and such that .
Theorem 1.1 implies the best constant in (1.28) exists even if Next, with the aid of Theorem 1.1, we are able to study the attainability of and classify the existence of minimizers to (1.26), which is
Theorem 1.2.
Suppose satisfies assumption (V1)-(V2) and , where is shown in (1.34), the we have the following alternatives:
Remark 1.2.
Theorem 1.2 indicates that the minimizers to (1.26) do not exist when is large enough. A natural question is the behaviors of ground states as , where is the existence threshold defined by (1.34). To explore this, we perform the scaling argument and investigate the convergence to get
Theorem 1.3.
Suppose that satisfies and let be the minimizer of given in Theorem 1.2 with . Then, we have
-
(i).
(1.39) - (ii).
Theorem 1.3 implies as , the ground states to (1.4) concentrate and their basic blow-up behaviors are captured by the least energy solution to potential-free Mean-field Games systems with some mild assumptions imposed on . Moreover, by imposing some typical local expansions on potential , one can obtain the refined asymptotics of ground states, which are summarized as
Theorem 1.4.
Assume that all conditions in Theorem 1.3 hold and suppose that has distinct zeros denoted by and there exist , and such that
Define
where and with
Let be the sequence given by (1.40) and be the limiting solution. Then we have . Moreover, as
and
(1.44) |
where and are given by (1.22) and (1.39), respectively. In particular, up to a subsequence,
(1.45) |
Theorem 1.4 demonstrates that under certain types of potentials with the local polynomial expansions, ground states to (1.4) are localized as , in which the locations converge to the flattest minima of .
The rest of this paper is organized as follows. In Section 2, we give some preliminaries for the investigation of ground states to (1.4) with . Section 3 is dedicated to the formulation of the optimal Gagliardo-Nirenberg type’s inequality and the proof of Theorem 1.1. In Section 4, we prove Theorem 1.2 by using the blow-up analysis and the Gagliardo-Nirenberg inequality shown in Theorem 1.1. Finally, in Section 5, we focus on Theorem 1.3 and 1.4, i.e. discuss the existence and concentration behaviors of ground states in some singular limit of given in (1.4). Without confusing readers, is chosen as a generic constant, which may vary line to line.
2 Preliminaries
This section is devoted to some preliminary results including existence and regularities of the solutions to Hamilton-Jacobi and Fokker-Planck equations.
2.1 Hamilton-Jacobi Equations
Consider the following Hamilton-Jacobi equation:
(2.1) |
where is a bounded domain with the smooth boundary, and . For the local estimates of the solutions to (2.1), we have
Lemma 2.1 (C.f. Theorem 1.1 in [14]).
Let , , and . Suppose solves (2.1) in the strong sense. Then for each and , we have
where and the constant .
Since our arguments in Section 3, 4 and 5 involve some limits of solution sequences, we also focus on the following sequence of Hamilton-Jacobi equations:
(2.2) |
where and are fixed. Here denote the solution pair to (2.2). Concerning the regularities of , we obtain
Lemma 2.2 (C.f. Lemma 3.1 in [14]).
Assume that satisfies and . Suppose the potential functions are uniformly local Hölder continuous satisfying as and there exists sufficiently large such that
(2.3) |
where the positive constants , , , and are independent of . Define as the solutions to (2.2). Then, we have for all ,
(2.4) |
where constant depends on , , , , , and
In particular, if each satisfies
(2.5) |
where and independent of then we have
(2.6) |
where constant depends on , , , , , and
For the lower bounds of , we have the following results:
Lemma 2.3 (C.f. Lemma 3.2 in [14]).
Suppose all conditions in Lemma 2.2 hold. Let be a family of solutions and assume that are bounded from below uniformly. Then there exist positive constants and independent of such that
(2.7) |
In particular, if the following conditions hold on
(2.8) |
where constants and are independent of then we have
(2.9) |
If in (2.8) and there exist and independent of such that
(2.10) |
then (2.9) also holds.
The following results are concerned with the existence of the classical solution to (2.2), which are
Lemma 2.4 (C.f. Lemma 3.3 in [14]).
Suppose are locally Hölder continuous and bounded from below uniformly in . Define
(2.11) |
Then we have
-
(i).
are finite for every and (2.2) admits a solution with and being bounded from below (may not uniform in ). Moreover,
- (ii).
(iii). If satisfies (1.25b) with replaced by and positive constants , and independent of then there exist uniformly bounded from below classical solutions to problem (2.2) satisfying estimate (2.7).
2.2 Fokker-Planck Equations
Now, we focus on the following Fokker-Planck equations:
(2.14) |
where is given and denotes the solution. Firstly, we state the regularity results of solutions to equation (2.14), which are
Lemma 2.5.
Proof.
See the proof of Lemma 3.5 in [14]. ∎
Lemma 2.6 (C.f. Corollary 1.1 in [14]).
Assume that is the solution to (2.14) with
Then for , there exists a positive constant depending only on and such that
(2.15) |
Moreover, there exists a positive constant only depending on and such that
(2.16) |
Next, we discuss the exponential decay property of the solutions to system (1.11) and obtain
Lemma 2.7.
Assume that with and is the solution of the following potential-free problem
(2.19) |
Suppose is bounded from below. Then, we have there exist such that
(2.20) |
Proof.
Noting that with , we use Sobolev embedding to get for some , and thus . Moreover, by using the fact that and the interpolation inequality, one finds for every . Therefore, invoking Lemma A.1 and Lemma A.2, one can obtain that for some and , which implies
The rest of proof follows from [4, Proposition 4.2] and [14, Lemma 3.6]. ∎
Thanks to Lemma 2.7, we establish Pohozaev identities satisfied by the solution to system (2.19), which are
Lemma 2.8 (C.f. Lemma 3.1 in [5]).
Assume all conditions satisfied by hold in Lemma 2.7 and denote . Then the following identities hold:
(2.23) |
Proof.
Proceeding the similar argument shown in Lemma 3.7 of [14], we can prove this lemma. For the sake of completeness, we exhibit the proof briefly. First all, we multiply the -equation and -equation in (2.19) by and , respectively, then integrate them by parts and subtract the two identities to get
(2.24) |
where we have used the exponential decay property of shown in Lemma 2.7 and the uniformly boundedness of stated in Lemma 2.2.
Next, we focus on the proof of the following identity:
(2.25) |
In fact, by testing the first equation and the second equation in (2.19) against and , we apply the integration by parts to obtain
(2.26) |
and
(2.27) |
where the boundary integrals vanish due to the decay property of and the upper bound of Also, we find
(2.28) |
Collecting (2.26), (2.27) and (2.2), we have the following equality holds:
(2.29) |
With the help of the integration by parts, one further gets
which indicates (2.25) by using the -equation in (2.19). In addition, since and , we obtain
(2.30) |
Finally, by using (2.24), (2.25) and (2.30), we conclude that (2.23) holds.
We mention that the argument shown above hold only when and in this case, the Fokker-Planck equation can be solved in the strong sense. When , one can only solve the Fokker-Planck equation in the weak sense. Whereas, we can replace with in (1.6) and proceed the same argument shown above with , then take the limit to get the desired conclusion.
∎
3 Optimal Gagliardo-Nirenberg Type’s Inequality
In this section, we are going to discuss the existence of minimizers to problem (1.28) and prove Theorem 1.1. As mentioned above, problem (1.28) is scaling invariant under the scaling for any and Therefore, one can verify that (1.28) is equivalent to
(3.1) |
where
(3.2) |
Now, we start by studying the subcritical mass exponent case of problem (3.1), namely, . For this case, Bernardini [4] proved that there exists for every solving the following system
(3.6) |
which is the classical solution to system (3.6). Furthermore, the author showed there exist such that
(3.7) |
In particular, Bernardini obtained the following minimization problem
(3.8) |
with
(3.9) |
is attained by the pair with . In addition, invoking Lemma 2.8, one finds
(3.12) |
Collecting the results shown above, we are able to investigate a relationship between , the minimizer of (3.8) and the minimizer of problem (3.1), which is
Lemma 3.1.
For any fixed and , problem (1.28) is attained by with . More precisely, we have
(3.13) |
Proof.
We follow the procedures shown in Lemma 4.1 of [14] to prove this lemma. First of all, we define
(3.14) |
With the definition (3.14), the minimization problem (3.1) can be rewritten as
(3.15) |
Now, we aim to verify that (3.15) is attained by , which is the minimizer of (3.8). First of all, we estimate the energy defined by (3.8) from below. We remark that provided with . Thus, we only need to consider the case that satisfying . Define for , then we have
(3.16) |
where the equality holds if and only if
It then follows from the definition of and (3) that
which yields
(3.17) |
Denote
then we invoke (3.17) to obtain
(3.18) |
where we have used the definition (3.14) and the fact
We can see from Lemma 3.1 that for all , Gagliardo-Nirenberg type inequalities given by (3.1) can be attained under the subcritical mass exponent case . In addition, invoking (3.12) and (3.13), we obtain that
(3.20) |
and
(3.21) |
The next lemma will indicate that defined in (3.1) is uniformly bounded as , which is essential for us to investigate the mass critical exponent case and prove Theorem 1.1.
Lemma 3.2.
There are constants and independent of such that for all with small,
(3.22) |
Proof.
We first estimate from above uniformly in . By setting with , we have for any and
(3.23) |
where we have used the following inequality
We next focus on the positive lower bound satisfied by uniformly in . To show this, we argue by contradiction and assume
(3.24) |
Lemma 3.1 implies there is a minimizer of problem (1.28). Since (1.28) is invariant under the scaling for any and , we normalize to get
(3.25) |
By using this equality and the Hardy-Littlewood-Sobolev inequality given in (A.2), we obtain
(3.26) |
where is the best constant. On the other hand, since , we follow the argument shown in [22, Theorem 4.3] to get
(3.27) |
Hence,
(3.28) |
Then it follows from (3.24), (3.28) and (1.28) that, as
(3.29) |
Proceeding the same argument shown in Lemma 3.5 of [14], one finds By using the Sobolev embedding theorem, we obtain
(3.30) |
On the other hand, the following interpolation inequality holds:
where With the help of (3.30), we further get as and
which reaches a contradiction to (3.25). Thus, we have independent of such that
(3.31) |
Finally, combining (3.31) with (3.23), one finds (3.22) holds. This completes the proof of the lemma. ∎
With the aid of the uniform boundedness of , we next establish the uniform bound of as , which is
Lemma 3.3.
Let , be the solution of
(3.35) |
Define . Assume that each is bounded from below and there exists a constant independent of such that
(3.36) |
then there is independent of such that
(3.37) |
Proof.
The proof is similar as shown in [14, Lemma 4.3]. We proceed by contradiction and suppose that up to a subsequence,
(3.38) |
Now, we fix without loss of generality, as this is due to the fact that is bounded from below. Define
(3.39) |
then, by (3.36) and (3.38), we obtain that up to a subsequence
(3.40) |
and
(3.41) |
Recall the definition of , then we deduce from (1.7) and (3.39) that
(3.42) |
In light of (3.38) and (3.39), we infer that
(3.43) |
This together with (3.40) implies that for any ,
(3.44) |
On the other hand, invoking (3.39) and (3.35), one can obtain that
(3.48) |
It follows from (3.36) and (3.38) that In addition, we claim that
(3.49) |
where independent of . Indeed, by the definition of , we get
(3.50) |
For , we apply (3.43) to get
(3.51) |
For , taking into account the condition (3.36), we get from (3.40) that
(3.52) |
Then, we conclude that (3.49) holds by collecting (3.38), (3.50)-(3.52) and the fact . Hence, applying Lemma 2.2 to the first equation in (3.48), one finds
(3.53) |
Noting that , we deduce from (3.53) that
(3.54) |
Now, we turn our attention to Hölder estimates of and the proof of Hölder continuity of is the same as shown in Lemma 4.3 of [14]. In fact, we obtain for some
(3.55) |
Assume that is a maximum point of , i.e., Then we deduce from (3.55) that there exists independent of such that . Thus,
which contradicts (3.44), and the proof of the lemma is finished. ∎
With the aid of Lemma 3.1, Lemma 3.2 and Lemma 3.3, we are able to show conclusions stated in Theorem 1.1, which are
Proof of Theorem 1.1:
Proof.
We first recall that, for any and , denotes the solution to system (3.6), and the pair with is a minimizer of the minimization problem (3.15), in which satisfies the estimate (3.7). Now, we take
in (3.6), then one can deduce from (3.21) that
Moreover, we obtain that, up to a subsequence,
(3.56) |
Since depends on , to emphasize the dependence of a solution on , we will rewrite as . Hence, we know from (3.6) that satisfies
(3.60) |
We can infer from Lemma 3.2 that, up to a subsequence,
In addition, invoking (3.56) we have that as , where
(3.61) |
Moreover, due to the relation (3.21), we obtain that, up to a subsequence,
(3.62) |
and it follows from (3.12) that
(3.63) |
Applying Lemma 3.3, we derive from (3.62) and (3.63) that
(3.64) |
Then, by using the estimate (2.6) with from Lemma 2.3, we have
(3.65) |
which, together with the definition of , yields
(3.66) |
Proceeding the arguments similar as those used in the proof of (3.55), we collect (3.63)-(3.66) to obtain that
(3.67) |
and thus
(3.68) |
We may assume that due to the fact that is bounded from below. Hence, by the first equation of (3.60), one has
which together with (3.64) and a similar argument as used in [5, Lemma 4.1] to obtain that there are and a large independent of such that,
(3.69) |
Now, we rewrite the first equation of (3.60) as
(3.70) |
where . By performing the same procedure shown in (3.50), one can see that
for some independent of . Then, we apply the standard elliptic regularity to (3.70) and obtain
(3.71) |
where Performing the standard diagonal procedure, we take the limit and apply Arzelà-Ascoli theorem, (3.67) and (3.71) to obtain that there exists such that
(3.72) |
Combining (3.60), (3.62) and (3.72), we conclude that satisfies
(3.76) |
In light of (3.69) and Fatou’s lemma, we have
(3.77) |
Moreover, by Lemma 2.7, we obtain that there exists some such that . In addition, by using (3.65), we get . It then follows from Lemma 2.8 that
(3.78) |
Next, we discuss the relationship between and with We claim that
(3.79) |
Indeed, we first utilize Lemma 3.1 and obtain
(3.80) | ||||
(3.81) |
Then, we derive from (3.63) that, as ,
Moreover, one takes the limit in (3.80) to get
(3.82) |
To complete the proof of our claim, it suffices to prove that the "=" holds in (3.82). Suppose the contrary that , then by the definition of we get that there exists given in (1.28) such that
(3.83) |
where is sufficiently small. On the other hand, by the definition of one finds
(3.84) |
Since
then we can pass a limit in (3.83) and (3) to get
which reaches a contradiction. Hence, the claim holds, i.e. .
Next, we prove Since solves (3.76) and with , we conclude from (3.72) and Lemma 2.2 that Then by standard elliptic estimates, the boundedness of and the exponentially decaying property of , one can prove that .
Finally, it follows from (3.63) and (3.82) that
(3.85) |
where is given in (3.61). Then, by the fact , we deduce from (3.77), (3.78) and (3.85) that
(3.86) |
which shows is a minimizer of and
These facts together with (3.76) indicate (1.33) holds. Now, we finish The proof of Theorem 1.1.
∎
4 Existence of Ground States: Coercive Potential MFGs
In this section, we shall discuss the existence of minimizers to problem (1.26). To this end, we have to perform the regularization procedure on (1.27) since when , the -component enjoys the worse regularity. In detail, we first consider the following auxiliary minimization problem
(4.1) |
where is given by (3.2) and
(4.2) |
and is the standard mollifier with
for is sufficiently small. With the regularized energy (4.2), we are able to study the existence of minimizers to (1.26) by taking the limit. The crucial step in this procedure, as discussed in [10], is the uniformly boundedness of in , in which is assumed to be a minimizer of (4.2).
Before proving Theorem 1.2, we collect some vital result shown in Section 3, which is
(4.3) |
where is given by (Blow-up Behaviors of Ground States in Ergodic Mean-field Games Systems with Hartree-type Coupling) and is defined by (1.34).
Then, we shall first prove energy given by (1.27) has a minimizer if and only if , where is defined by (1.1). Next, we show that there exists such that is a solution to (1.38) when is assumed to satisfy (1.25) when Following the procedures discussed above, we are able to prove conclusions stated in Theorem 1.2. We would like to remark that with (1.25b) in assumption (V2) imposed on potential , the condition in (3.2) must be satisfied for any minimizer. With this assumption, Gagliardo-Nirenberg type’s inequality (4.3) is valid. Next, we state some crucial propositions and lemmas, which will be used in the proof of Theorem 1.2, as follows:
Lemma 4.1.
Let
Assume that with . Then, the embedding is compact for any , where if and if .
In light of we establish the following lemma for the uniformly boundedness of
Lemma 4.2.
Suppose that is locally Hölder continuous and satisfies (1.25). Let be solutions to the following systems
(4.7) |
where with , with satisfies for all , and ,
(4.8) |
and
(4.9) |
Assume that
(4.10) |
and for all , is bounded from below uniformly. Then we have
(4.11) |
Proof.
By slightly modifying the argument shown in [14, Lemma 5.2], we finish the proof of this lemma. ∎
With the preliminary results shown above, we now begin the proof of Theorem 1.2.
Proof of Theorem 1.2:
Proof.
We first prove the Conclusion (i) in Theorem 1.2. To this end, we focus on the auxiliary problem (4.1). Invoking the Young’s inequality for convolution and the property of mollifier, one finds
(4.12) |
for any . Here, we have used the following Hardy-Littlewood-Sobolev inequality
(4.13) |
As a consequence, in light of (4.3) and (4.12), we get
(4.14) |
Next, we show that the minimization problem (4.1) is attainable. We first show that there exists independent of such that
(4.15) |
where is given by (4.1). Indeed, choosing
one can find
which indicates that (4.15) holds. Let be a minimizing sequence of problem (4.1), then we have from (4.15) that there exists independent of such that
(4.16) |
Moreover, it follows from (4.3), (4.14), (4.16) and the fact that
(4.17) |
and
(4.18) |
where is independent of . The subsequent argument for proving Conclusion (i) is similar as shown in the proof of Theorem 1.3 in [14]. In fact, with the aid of the key Lemma 2.5, we obtain from (4.17) that
(4.19) |
where is defined by (1.24) and is some constant independent of . As a consequence, there exists such that
(4.20) |
In light of the assumption (V1), , given in Subsection 1.2, one can deduce from Lemma 4.1 that
(4.21) |
Therefore, up to a subsequence,
(4.22) |
In addition, thanks to the convexity of , by letting in (4.17), we have there exists independent of such that
(4.23) |
Moreover,
(4.24) |
and
(4.25) |
Combining (4.20) and (4.21) with (4.25), we deduce that . Then, one invokes (4.22) and (4.23) to get
which indicates is a minimizer of problem (4.1). Finally, similarly as the proof of Proposition 3.4 in [10] and the arguments shown in Proposition 5.1 and Proposition 5.2 in [14], we apply Lemma 2.4 to obtain that there exists bounded from below (depending on ) and such that
(4.29) |
For each fixed , we utilize Lemma 2.2 to obtain that there exists depends on such that . Noting that and , we have from the classical regularity of the -equation in (4.29) that . We next prove
(4.30) |
To show this, we apply the integration by parts to the -equation and the -equation in (4.29), then get
and
(4.31) |
where we have used the fact that . Collecting (4.23), (4.24) and (4.31), one finds (4.30) holds.
Next, we let and show the existence of the minimizer to problem (1.26). Noting satisfies (4.10) with replaced by We utilize Young’s inequality for convolution and Hardy-Littlewood-Sobolev inequality (A.1) to get
and
Then, collecting (4.23) and (4.30), we invoke Lemma 4.2 to conclude that
(4.32) |
Then, by using Lemma 2.2, we obtain
(4.33) |
Since is bounded from below, without loss of generality, we assume that . In light of (2.7), one finds that as , which indicates each admits its minimum at some finite point . By using (4.30), (4.32) and the coercivity of , we obtain from the -equation of (4.29) that is uniformly bounded with respect to . The fact together with (4.33) implies that there exists independent of such that
where we have used (1.25c) in the second inequality. Since are bounded from below uniformly, one can employ Lemma 2.3 to get that Thus, with the assumptions (1.25) imposed on , we get
(4.34) |
where are independent of .
In light of (4.32) and (4.33), one finds for any and ,
(4.35) |
where the constant depends only on , and is independent of . Then, with the help of Lemma 2.5, we obtain from (4.35) that . Taking large enough, we utilize Sobolev embedding theorem to get
(4.36) |
To estimate we rewrite the -equation of (4.29) as
(4.37) |
Since , then with . Thus, we deduce from Lemma A.2 that for some . Now, by using (4.32), (4.33) and the fact that is locally Hölder continuous, we obtain that for any ,
Then we utilize the standard elliptic regularity in (4.37) to get
(4.38) |
where Letting and proceeding the standard diagonalization procedure, we invoke Arzelà-Ascoli theorem to find there exists such that
(4.39) |
In addition, by using Lemma 2.5 and (4.23), we find there exists such that
(4.40) |
Moreover, invoking Lemma 4.1, one finds
(4.41) |
Passing to the limit as in (4.29), we then obtain from (4.30) and (4.39)-(4.41) that there exists such that satisfies (1.38). In addition, we infer from (4.33) and (4.34) that
(4.42) |
Recall that a.e. as in , then we use (4.32) to get that . Then, proceeding the same argument as shown in the proof of Proposition 5.2 in [14], one can further find from (1.38) and (4.42) that
and , . | (4.43) |
Finally, we prove that is a minimizer of . To this end, we claim that for ,
(4.44) |
where is given in (1.26). On one hand, in view of (4.12), it is straightforward to get . On the other hand, we aim to show . Due to the definition of , for any , we choose such that . In light of (4.12), we conclude that for small enough, . Thus,
Letting at first and then , one has . Combining the two facts, we finish the proof of (4.44).
We collect (4.40), (4.41), (4.44) and the convexity of to get
which implies is a minimizer of . This completes the proof of Conclusion (i).
Now, we focus on Conclusion (ii) of Theorem 1.2. We have the fact that given in Theorem 1.1 is a minimizer of problem (1.28) with . To simplify notation, we rewrite as , then define
(4.45) |
where the constraint set and are defined by (1.1) and (1.34), respectively. Since and decays exponentially as stated in Theorem 1.1, we utilize Lemma 2.8 to find
(4.46) |
Thanks to (4.46), we substitute (4.45) into (1.27), then obtain that if
(4.47) |
Therefore, we have for which indicates that problem (1.22) does not admit any minimizer.
Now, we are concentrated at the critical case and plan to show Conclusion (iii). To begin with, we prove that up to a subsequence,
(4.48) |
Indeed, since as shown in and is defined by (1.22), we have for any , such that
(4.49) |
Noting that , we further obtain
(4.50) | ||||
By a straightforward computation, one has as
(4.51) |
We collect (4.49), (4.50) and (4.51) to get
(4.52) |
Letting in (4.52), one has from (4.52) that
(4.53) |
In addition, define as a minimizer of for any fixed , then we find and
It follows that
(4.54) |
Combining (4.53) with (4.54), one has
(4.55) |
In light of assumptions (V1) and (V2) stated in Subsection 1.2 for potential , we set in (4) to get
Now, we focus on the proof Conclusion (iii). If conclusion (iii) is not true, then we assume that has a minimizer . By using (4.48), we further obtain
Combining this with (4.3), one gets
(4.56) |
which implies Whereas, with the assumption (1.25c) and the fact we have It follows that a.e., which is a contradiction. Consequently, we complete the proof of Conclusion (iii). ∎
Theorem 1.2 implies that when the potential satisfies some mild assumptions given by (V1), (V2) and (V3) stated in Section 1, system (1.4) admits the ground states only when , where is explicitly shown in Theorem 1.1 and has a strong connection with the existence of ground states to the potential-free nonlocal Mean-field Games system. In the next section, we shall discuss the asymptotic behaviors of ground states to problem (1.4) as
5 Asymptotics of Ground States as
This section is devoted to the proof of Theorem 1.3 and Theorem 1.4. More precisely, we shall describe the asymptotic profile of least energy solutions to (1.11) as
5.1 Basic Blow-up Behaviors
In this subsection, we analyze the basic asymptotic behaviors of ground states to (1.11) as and prove Theorem 1.3.
Proof of Theorem 1.3:
Proof.
To prove Conclusion (i), we perform the blow-up argument and assume
Then we utilize Lemma 2.5 to get
(5.1) |
Consequently, we have there exists such that
(5.2) |
Now, we prove given by (1.1). Indeed, noting (5.1), we have
(5.3) |
By using the assumptions (V1), (V2) and (V3) satisfied by , we conclude from (5.2), (5.3) and Lemma 4.1 that
(5.4) |
which implies Moreover, thanks to (5.2), one gets weakly. It follows that
which implies . Hence, we obtain and further due to (5.2) and (5.4). Moreover, one has from (4.48) that
Therefore, is a minimizer of which yields a contradiction to Conclusion (iii) in Theorem 1.2. This finishes the proof of Conclusion (i).
(ii). Note that
As stated in Conclusion (i) of Theorem 1.2, we have each is bounded from below and satisfies . Hence, there exists such that , which indicates thanks to the definition given in (1.40).
In light of (1.38) and (1.40), we find that satisfies the following system
(5.8) |
Collecting (1.39), (4.3) and (4.48), one gets
(5.9) |
(5.10) |
and
(5.11) |
Following the similar argument employed in the derivation of (4.31), we utilize (5.8) and (5.9) to obtain
which implies
(5.12) |
We apply the maximum principle to the -equation in (5.8), then deduce that
(5.13) |
which indicates
(5.14) |
Now, we claim that there exists some constant such that
(5.15) |
If this is not the case, one can find some subsequence such that . Then, with the aid of (5.13), one has
(5.16) |
where is some constant independent of Define
(5.17) |
then one has
By substituting (5.17) into (5.8), we find
(5.20) |
By using the assumption (1.25b), one gets
where is some constant independent of Noting that
we utilize the maximal regularity shown in Lemma 2.1 to obtain
where and are some constants. Focusing on the -equation of (5.20), we similarly apply the standard elliptic regularity estimates (See Theorem 1.6.5 in [6]) to obtain with independent of . By a direct calculation, we conclude from (5.16) that
(5.21) |
This together with the Hölder’s continuity of implies that
(5.22) |
where sufficiently large and independent of . In light of , we have the fact that As a consequence, there exists such that Then It follows from (5.22) that
as Whereas,
which reaches a contradiction. This completes the proof of our claim (5.15).
Moreover, since satisfies (1.25b), one further obtains for large enough,
(5.23) |
where constant depends on and is independent of
Similarly as discussed in the proof of Theorem 1.2, we estimate and rewrite the -equation of (5.8) as
(5.24) |
Noting that , we utilize Lemma 2.1 to get , i.e. . By using Lemma 2.6, we further obtain that for some since satisfies the second equation in (5.8).
(5.25) |
In light of (5.14), we have from (5.25) that there exists a constant such that
(5.26) |
Now, we claim that up to a subsequence,
(5.27) |
If not, one has either or with In the two cases, we both have a.e. in for some . It then follows from (5.26) that
By using (5.9), we find there exists such that
(5.28) |
where thanks to (5.26) and is given by (1.24). Furthermore, invoking (5.25), one has in . Moreover, combining (5.8) with (5.12), we obtain satisfies
(5.32) |
where we have followed the procedure performed in the proof of (3.86) shown in Section 3. In particular, we have used Lemma 2.8 to obtain that is a minimizer of (1.28) and . Thus, we have from (5.32) that satisfies (1.33). On the other hand, we obtain in , and then with the aid of (5.28), one finds
which indicates that (1.41) holds.
Finally, we prove that (1.43) holds when (1.42) is imposed on . To this end, we argue by contradiction and assume that, then, up to a subsequence,
(5.33) |
Define
(5.37) |
Now, we claim that and independent of such that
(5.38) |
Invoking (5.37), we have (5.38) is equivalent to
(5.39) |
In light of (5.14), we find
(5.40) |
To show (5.39), we have from the first equation in (5.8) that satisfies
(5.41) |
Following the argument shown in [10, Theorem 4.1], we consider the following two cases:
Case 1: Assume that there exists some constant independent of such that satisfies
Then thanks to (5.40), we follow the same argument performed in the derivation of (5.24), (5.25) and (5.26) to obtain the claim (5.38).
Case 2: Suppose that satisfies
(5.42) |
Define
(5.43) |
then obtain from (5.8) that satisfies
(5.47) |
Since satisfies (1.42), we utilize Lemma 2.2 to get
(5.48) |
Denote and , which are the minimum and maximum points of and respectively. With the aid of (5.27), we obtain Then, we obtain from (5.48) that
(5.49) |
As a consequence,
(5.50) |
Collecting (5.42), (5.49) and (5.50), we proceed the same argument shown in [10, Theorem 4.1] to get with and independent of Since is maximum point of , we combine (5.40) with (5.43) to get Hence, we have there exists some independent of such that
Noting , we find from the above estimate and (5.43) that (5.39) holds.
In Theorem 1.3, we see that as the ground states to problem (1.22) concentrate and become localized patterns, in which the profiles are determined by the minimizer to problem (3.1). We mention that with some typical expansions imposed on potential locally, the detailed asymptotics of ground states can be captured and we shall discuss them in Subsection 5.2.
5.2 Refined Blow-up Behaviors
In this subsection, we shall analyze the refined asymptotic profiles of the rescaled minimizer and prove Theorem 1.4. As shown in Theorem 1.4, we assume has distinct zeros defined by ; moreover, , , such that
(5.51) |
Define , and denote
(5.52) |
Set consisted of all weighted flattest zeros of Collecting the above notations, we first establish the precise upper bound of as stated as follows:
Lemma 5.1.
The , defined by (1.26), satisfies
(5.53) |
Proof.
The proof is similar as the argument shown in [14, Lemma 6.1] with slight modifications. We omit the details. ∎
In Section 5, we find converges to in the following sense:
where is the minimizer of and correspondingly, satisfies (1.33). Moreover, Lemma 2.7 and Lemma 2.8 imply and such that
(5.54) |
and
(5.55) |
Proof of Theorem 1.4:
Proof.
Thanks to Theorem 1.3, we have for some In addition, noting that is the minimizer of problem (1.26), one gets
(5.56) |
By the direct calculation, we obtain
(5.57) |
In light of , then one has
(5.58) |
Now, we claim that
(5.59) |
Indeed, if is not true, then we have either or up to a subsequence, . Then by using Fatou’s lemma, we conclude from (1.41), (5.57) and (5.58) that
for any constant large enough. Combining (5.9) with (5.56), one gets
which contradicts Lemma 5.1. This completes the proof of claim (5.59).
With the help of (5.59), we obtain that such that, up to a subsequence,
Then we aim to prove that satisfies (1.45), i.e. with . To begin with, noting , we apply Fatou’s lemma then conclude from (5.51), (5.52) and (1.41) that
(5.60) |
where the last two equalities hold if and only if (1.45) holds. Thus, we have
(5.61) |
where the equality holds in the second step if and only if
(5.62) |
Thus, combining (5.61) with (5.53), one has all equalities in (5.61) hold. It immediately follows that all "=" in (5.60) also hold. Now, we obtain (1.44) and (1.45), which completes the proof of Theorem. 1.4. ∎
6 Discussion
In this paper, we mainly investigated the existence of ground states to (1.4) with critical mass exponent in the nonlocal coupling. First of all, we analyzed the attainability of the best constant in the Gagliardo-Nirenberg type’s ratio defined by (1.28), which corresponds the existence of ground states to the potential-free Mean-field Games system. Next, with the aid of Gagliardo-Nirenberg type’s inequality, we employ the variational approach to classify the existence of minimizers to the constrained minimization problem (1.22). In particular, while discussing the existence of classical solutions to (1.4) under the subcritical mass, we introduced the mollifier and showed the of to the mollified minimization problems, in which the Hardy-Littlewood-Sobolev inequality is crucial. Then taking the limit and applying standard elliptic regularities, we obtained the existence of classical solutions to (1.4) under the subcritical mass. Finally, with some assumptions imposed in the potential we performed the scaling argument and blow-up analysis to derive the asymptotic behaviors of ground states to (1.4) in the singular limit of , where the Pohozaev identities have been intensively used for the convergence of
There are some interesting problems that deserve the explorations in the future. In Section 3, some technical restriction on was imposed, which is the boundedness of for sufficiently small It is an open problem to remove this condition while establishing the Gagliardo-Nirenberg type’s inequality. It is also intriguing to investigate the properties of ground states including uniqueness, symmetries, etc. to potential-free Mean-field Games systems (1.11) with the Hartree coupling and polynomial Hamiltonian. The extension of our results into a general class of potential is a challenging problem due to the lower bounds of the value function
Appendix A Basic proerties of Riesz potential
This Appendix is devoted to some well-known results for the estimates involving Riesz potential, which can be found in [22, Theorem 4.3], [25, Theorem 14.37] and [5, Theorem 2.8].
Lemma A.1 (Hardy Littlewood-Sobolev inequality).
Assume that and . Then for any , it holds
(A.1) |
where constant depending on , and
Moreover, suppose that with , and . Then, we have there exists a sharp constant independent of and such that
(A.2) |
In particular, we find from Lemma A.1 that if in (A.2) and , then there exists a sharp constant independent of and such that
(A.3) |
Lemma A.2 (C.f. Theorem 2.8 in [5]).
Let and be positive constants such that and . Then for any , we have
(A.4) |
where and . Moreover, if , we have
(A.5) |
In particular, there exists constant such that
Lemma A.2 exhibits the and Hölder estimates of under certain conditions of and .
Acknowledgments
Xiaoyu Zeng is supported by NSFC (Grant Nos. 12322106, 12171379, 12271417) . Huan-Song Zhou is supported by NSFC (Grant Nos. 11931012, 12371118) .
References
- [1] T. Bartsch and Z. Wang. Existence and multiplicity results for some superlinear elliptic problems on . Commun. Part. Diff. Eq., 20(9-10):1725–1741, 1995.
- [2] A. Bensoussan and J. Frehse. Nonlinear elliptic systems in stochastic game theory. J. Reine Angew. Math., 350:23–67, 1984.
- [3] A. Bensoussan and J. Frehse. Ergodic Bellman systems for stochastic games in arbitrary dimension. Proc. Roy. Soc. London Ser. A, 449(1935):65–77, 1995.
- [4] C. Bernardini. Mass concentration for Ergodic Choquard Mean-Field Games. ESAIM: COCV, 2023.
- [5] C. Bernardini and A. Cesaroni. Ergodic mean-field games with aggregation of choquard-type. J. Differential Equations, 364:296–335, 2023.
- [6] V. Bogachev, N. Krylov, M. Röckner, and S. Shaposhnikov. Fokker–Planck–Kolmogorov Equations, volume 207. American Mathematical Society, 2022.
- [7] F. Cagnetti, D. Gomes, H. Mitake, and H. Tran. A new method for large time behavior of degenerate viscous Hamilton-Jacobi equations with convex Hamiltonians. Ann. Inst. H. Poincaré, 32(1):183–200, 2015.
- [8] P. Cardaliaguet. Long time average of first order mean field games and weak KAM theory. Dyn. Games Appl., 3(4):473–488, 2013.
- [9] P. Cardaliaguet, J. Lasry, P. Lions, and A. Porretta. Long time average of mean field games. Netw. Heterog. Media, 7(2):279–301, 2012.
- [10] A. Cesaroni and M. Cirant. Concentration of ground states in stationary mean-field games systems. Anal. PDE, 12(3):737–787, 2018.
- [11] M. Cirant. A generalization of the Hopf–Cole transformation for stationary Mean-Field Games systems. C. R. Math. Acad. Sci. Paris, 353(9):807–811, 2015.
- [12] M. Cirant. Stationary focusing mean-field games. Commun. Part. Diff. Eq., 41(8):1324–1346, 2016.
- [13] M. Cirant and A. Goffi. Maximal -regularity for parabolic Hamilton-Jacobi equations and applications to mean field games. Ann. PDE, 7(2):Paper No. 19, 40, 2021.
- [14] M. Cirant, F. Kong, J. Wei, and X. Zeng. Critical mass phenomena and blow-up behavior of ground states in stationary second order mean-field games systems with decreasing cost. preprint, 2024.
- [15] D. Gomes and H. Mitake. Existence for stationary mean-field games with congestion and quadratic Hamiltonians. NoDEA Nonlinear Differential Equations Appl., 22(6):1897–1910, 2015.
- [16] D. Gomes, S. Patrizi, and V. Voskanyan. On the existence of classical solutions for stationary extended mean field games. Nonlinear Anal., 99:49–79, 2014.
- [17] D. Gomes, E. Pimentel, and H. Sánchez-Morgado. Time-dependent mean-field games in the superquadratic case. ESAIM Control Optim. Calc. Var., 22(2):562–580, 2016.
- [18] D. Gomes, G. Pires, and H. Sánchez-Morgado. A-priori estimates for stationary mean-field games. Netw. Heterog. Media, 7(2):303–314, 2012.
- [19] D. A. Gomes, E. Pimentel, and V. Voskanyan. Regularity theory for mean-field game systems. Springer, 2016.
- [20] M. Huang, R. Malhamé, and P. Caines. Large population stochastic dynamic games: closed-loop McKean-Vlasov systems and the Nash certainty equivalence principle. Commun. Inf. Syst., 6(3):221–251, 2006.
- [21] J. Lasry and P. Lions. Mean field games. Jpn. J. Math., 2(1):229–260, 2007.
- [22] E. H. Lieb and M. Loss. Analysis, volume 14 of Grad. Stud. Math. Providence, RI: AMS, American Mathematical Society, 1996.
- [23] A. Mészáros and F. Silva. A variational approach to second order mean field games with density constraints: the stationary case. J. Math. Pures Appl., 104(6):1135–1159, 2015.
- [24] M. Reed. Methods of modern mathematical physics: Functional analysis. Elsevier, 2012.
- [25] R. L. Wheeden and A. Zygmund. Measure and integral. An introduction to real analysis, volume 43 of Pure Appl. Math., Marcel Dekker. Marcel Dekker, Inc., New York, NY, 1977.