Loading [MathJax]/jax/element/mml/optable/GeneralPunctuation.js
Research article Special Issues

Deep reinforcement learning with emergent communication for coalitional negotiation games

  • Received: 19 December 2021 Revised: 17 January 2022 Accepted: 07 February 2022 Published: 07 March 2022
  • For tasks intractable for a single agent, agents must cooperate to accomplish complex goals. A good example is coalitional games, where a group of individuals forms coalitions to produce jointly and share surpluses. In such coalitional negotiation games, how to strategically negotiate to reach agreements on gain allocation is however a key challenge, when the agents are independent and selfish. This work therefore employs deep reinforcement learning (DRL) to build autonomous agent called DALSL that can deal with arbitrary coalitional games without human input. Furthermore, DALSL agent is equipped with the ability to exchange information between them through emergent communication. We have proved that the agent can successfully form a team, distribute the team's benefits fairly, and can effectively use the language channel to exchange specific information, thereby promoting the establishment of small coalition and shortening the negotiation process. The experimental results shows that the DALSL agent obtains higher payoff when negotiating with handcrafted agents and other RL-based agents; moreover, it outperforms other competitors with a larger margin when the language channel is allowed.

    Citation: Siqi Chen, Yang Yang, Ran Su. Deep reinforcement learning with emergent communication for coalitional negotiation games[J]. Mathematical Biosciences and Engineering, 2022, 19(5): 4592-4609. doi: 10.3934/mbe.2022212

    Related Papers:

    [1] Sergey A. Rashkovskiy . Quantization of Hamiltonian and non-Hamiltonian systems. Communications in Analysis and Mechanics, 2023, 15(2): 267-288. doi: 10.3934/cam.2023014
    [2] Antoine Bendimerad-Hohl, Ghislain Haine, Laurent Lefèvre, Denis Matignon . Stokes-Lagrange and Stokes-Dirac representations of N-dimensional port-Hamiltonian systems for modeling and control. Communications in Analysis and Mechanics, 2025, 17(2): 474-519. doi: 10.3934/cam.2025020
    [3] Richard Cushman . Normalization and reduction of the Stark Hamiltonian. Communications in Analysis and Mechanics, 2023, 15(3): 457-469. doi: 10.3934/cam.2023022
    [4] Henryk Żoła̧dek . An example in Hamiltonian dynamics. Communications in Analysis and Mechanics, 2024, 16(2): 431-447. doi: 10.3934/cam.2024020
    [5] Andrea Brugnoli, Ghislain Haine, Denis Matignon . Stokes-Dirac structures for distributed parameter port-Hamiltonian systems: An analytical viewpoint. Communications in Analysis and Mechanics, 2023, 15(3): 362-387. doi: 10.3934/cam.2023018
    [6] Xiao Qing Huang, Jia Feng Liao . Existence and asymptotic behavior for ground state sign-changing solutions of fractional Schrödinger-Poisson system with steep potential well. Communications in Analysis and Mechanics, 2024, 16(2): 307-333. doi: 10.3934/cam.2024015
    [7] Cheng Yang . On the Hamiltonian and geometric structure of Langmuir circulation. Communications in Analysis and Mechanics, 2023, 15(2): 58-69. doi: 10.3934/cam.2023004
    [8] Chunming Ju, Giovanni Molica Bisci, Binlin Zhang . On sequences of homoclinic solutions for fractional discrete p-Laplacian equations. Communications in Analysis and Mechanics, 2023, 15(4): 586-597. doi: 10.3934/cam.2023029
    [9] Long Ju, Jian Zhou, Yufeng Zhang . Conservation laws analysis of nonlinear partial differential equations and their linear soliton solutions and Hamiltonian structures. Communications in Analysis and Mechanics, 2023, 15(2): 24-49. doi: 10.3934/cam.2023002
    [10] Wenqian Lv . Ground states of a Kirchhoff equation with the potential on the lattice graphs. Communications in Analysis and Mechanics, 2023, 15(4): 792-810. doi: 10.3934/cam.2023038
  • For tasks intractable for a single agent, agents must cooperate to accomplish complex goals. A good example is coalitional games, where a group of individuals forms coalitions to produce jointly and share surpluses. In such coalitional negotiation games, how to strategically negotiate to reach agreements on gain allocation is however a key challenge, when the agents are independent and selfish. This work therefore employs deep reinforcement learning (DRL) to build autonomous agent called DALSL that can deal with arbitrary coalitional games without human input. Furthermore, DALSL agent is equipped with the ability to exchange information between them through emergent communication. We have proved that the agent can successfully form a team, distribute the team's benefits fairly, and can effectively use the language channel to exchange specific information, thereby promoting the establishment of small coalition and shortening the negotiation process. The experimental results shows that the DALSL agent obtains higher payoff when negotiating with handcrafted agents and other RL-based agents; moreover, it outperforms other competitors with a larger margin when the language channel is allowed.



    In this paper, we are interested in the existence and some asymptotic properties of ground state homoclinic orbits for the following first-order Hamiltonian system

    ˙z=JHz(t,z), (1.1)

    with the Hamiltonian function

    H(t,z)=12Lzz+A(ϵt)G(|z|). (1.2)

    Here, z=(u,v)RN×RN=R2N, ϵ>0 is a parameter, L is a symmetric 2N×2N matrix-valued function, and

    J=(0II0).

    is the usual symplectic matrix with I being the identity matrix in RN. As usual, we refer to a solution z of system (1.1) as a homoclinic orbit if z(t)0 and z(t)0 as |t|.

    It is widely known that Hamiltonian systems are very important dynamical systems, which have many applications in several natural science areas such as relativistic mechanics, celestial mechanics, gas dynamics, chemical kinetics, optimization and control theory and so on. The complicated dynamical behavior of Hamiltonian systems has the attracted attention of many mathematicians and physicists ever since Newton wrote down the differential equations describing planetary motions and derived Kepler's ellipses as solutions. For more details on applications of Hamiltonian systems, one can refer to [1] and the monograph [2] of Mawhin and Willem. We also refer the readers to see [3] for the Stokes-Dirac structures of the port-Hamiltonian systems.

    In the past few decades, Hamiltonian systems have attracted a considerable amount of interest due to many powerful applications in different fields; the literature related to these systems is extensive and encompasses several interesting lines of research on the topic of nonlinear analysis, including the existence, nonexistence, multiplicity and finer qualitative properties of homoclinic orbits. Here we cannot provide a complete and fully detailed list of references, but rather we limit ourselves to mentioning the works which are closely related to the content of the present paper.

    A major breakthrough was the pioneering paper of Rabinowitz [4] from 1978 who, for the first time, obtained periodic solutions of system (1.1) by using variational methods. After the celebrated work of Rabinowitz [4], based on the dual action and mountain pass argument, Coti-Zelati et al. [5] obtained the existence and multiplicity of homoclinic orbits under the condition of strictly convexity. Later on, this result was further detailed by [6,7] in which the authors established the existence result for infinitely many homoclinic orbits. Without the convexity condition, Hofer and Wysocki [8] independently investigated the existence of homoclinic orbits by combining the Fredholm operator theory and the linking argument. Tanaka [9] employed a suitable subharmonic approach to obtain one homoclinic orbit by relaxing the convexity condition. In [10], Rashkovskiy studied the quantization process of Hamiltonian and non-Hamiltonian systems.

    The main unusual feature of the first-order Hamiltonian system is that the associated energy functional is strongly indefinite. Generally speaking, for the strongly indefinite functionals refined variational methods like the Nehari manifold method and mountain pass theorem still do not apply. Some general critical point theories like the generalized linking theorem and other weaker versions for strongly indefinite functionals were subsequently developed by Kryszewski and Szulkin in [11] and Bartsch and Ding in [12]. Since then, based on the critical point theorems from [11,12] for strongly indefinite functionals, many scholars have gradually begun to investigate the existence and multiplicity of homoclinic orbits for non-autonomous Hamiltonian systems under some different conditions. More precisely, under the conditions that H depends periodically on t and has super-quadratic growth in z, Arioli and Szulkin [13], Chen and Ma [14], and Ding and Willem [15] obtained the existence result. Ding [16], Ding and Girardi [17], and Zhang et al. [18] studied the multiplicity result for homoclinic orbits. Concerning the asymptotic quadratic growth case, we refer to the work done by Szulkin and Zou [19] and Sun et al. [20]. Here we would like to emphasize that the periodicity condition is used to resolve the issue stemming from the lack of compactness since system (1.1) is set on the whole space R.

    On the other hand, without the condition of periodicity, the non-periodic problem is quite different due to the lack of compactness of Sobolev embeddings. In an early paper [21], Ding and Li utilized the coercive property of L to establish a variational framework with compactness, and they proved the existence of homoclinic orbits for the super-quadratic growth case. Also under the framework of compactness, Zhang and Liu [22] studied the sub-quadratic growth case. Regarding the asymptotically quadratic case, based on the infinite-dimensional linking argument, Ding and Jeanjean [23] established a multiplicity result for homoclinic orbits. Moreover, they imposed a control on the size of G with respect to the behavior of L to recover sufficient compactness. For the existence and exponential decay of homoclinic orbits for system (1.1) with nonperiodic super-quadratic and lack of compactness, we refer the reader to [24]. We also mention the recent paper by Zhang et al. [25] in which the existence and decay of ground state homoclinic orbits for system (1.1) with asymptotic periodicity are explored. For other results related to the Hamiltonian systems with strongly variational structure, we refer the reader to [26,27,28,29,30,31,32] and the references therein.

    It is worth pointing out that, in all of the works mentioned above, the authors were concerned mainly with the study of the existence and multiplicity of homoclinic orbits, and there are no papers considering the asymptotic properties of homoclinic orbits. Inspired by this fact and the work done by Alves and Germano [33] in which the authors investigated the existence and concentration of ground state solutions for the Schrödinger equation; in the present paper we aim to further study the existence and some asymptotic properties of ground-state homoclinic orbits for system (1.1) with Hamiltonian function (1.2). This is a very interesting issue that has motivated the present work.

    To continue the discussion, we introduce the following notation

    S=(Jddt+L),

    then, system (1.1) takes the following form

    Sz=A(ϵt)g(|z|)z,tR. (1.3)

    Before stating our results, we suppose the following conditions hold for L, A and G.

    (L) L is a constant symmetric 2N×2N matrix such that σ(JL)iR=, where σ denotes the spectrum of operator JL.

    (A) AC(R,R) and 0<inftRA(t)A:=lim|t|A(t)<A(0)=maxtRA(t);

    (g1) Gz(|z|)=g(|z|)z, gC(R+,R+), and there exist p>2 and c0>0 such that

    |g(s)|c0(1+|s|p2)for allsR+;

    (g2) g(s)=o(1) as s0, and G(s)/s2+ as s+;

    (g3) g(s) is strictly increasing in s on (0,+).

    Next we state the main result of this paper as follows.

    Theorem 1.1. Assume that conditions (L), (A), and (g1)-(g3) hold. Then we have the following results:

    (a) there exists ϵ0>0 such that system (1.1) has a ground state homoclinic orbit zϵ for each ϵ(0,ϵ0);

    (b) |zϵ| attains its maximum at tϵ, then,

    limϵ0A(ϵtϵ)=A(0),

    moreover, zϵ(t+tϵ)z as ϵ0, where z is a ground state homoclinic orbit of the limit system

    Sz=A(0)g(|z|)z,tR;

    (c) additionally, if A,gC1, and g(s)s=o(1) as s0, then there exist constants c,C>0 such that

    |z(t)|Cexp(c|ttϵ|)for alltR.

    We would like to emphasize that since our problem is carried out in the whole space, then the strongly indefiniteness of energy functionals and the lack of compactness are two major difficulties that we encounter in order to guarantee the existence of homoclinic orbits. More precisely, one reason is that strongly indefinite functionals are unbounded from below and from above so that the classical methods from the calculus of variations do not apply. The other reason is that the lack of compactness leads to the energy functionals not satisfying the necessary compactness property.

    Let us now outline the methods involved to prove Theorem 1.1. Indeed, based on the above reasons, first, we will take advantage of the method of the generalized Nehari manifold developed by Szulkin-Weth [34] to handle system (1.1), this is because such a strategy helps to overcome the difficulty caused by strongly indefinite features. Second, we must verify that the energy functional possesses the necessary compactness property at some energy level. This target will be accomplished by applying the energy comparison argument to establish some precise comparison relationships for the ground-state energy value between the original problem and certain auxiliary problems. Finally, combining the compactness analysis technique, Kato's inequality, and the sub-solution estimate, we can obtain the concentration property and decay of homoclinic orbits. Then Theorem 1.1 follows naturally.

    This paper is organized as follows. In Section 2, we establish the functional analytic setting associated with system (1.1). In Section 3, we present some technical results, and obtain the existence result for ground-state homoclinic orbits for the autonomous system. Section 4 is devoted to proofs of Theorem 1.1.

    Throughout the present paper, we will use the following notations:

    denotes the norm of the Lebesgue space L^{s}(\mathbb{R}) , 1\leq s\leq +\infty ;

    (\cdot, \cdot)_{2} denotes the usual inner product of the space L^{2}(\mathbb{R}) ;

    c , c_i , C_{i} represent various different positive constants.

    In what follows, we will establish the variational framework to work for system (1.1).

    Recall that S = -(\mathscr{J}\frac{d}{dt}+L) is a self-adjoint operator on the space L^{2}: = L^{2}(\mathbb{R}, \mathbb{R}^{2N}) with the domain \mathcal{D}(S) = H^{1}(\mathbb{R}, \mathbb{R}^{2N}) ; according to the discussion in [13], we can know that, under the condition (L) , there exists a > 0 such that (-a, a)\cap \sigma(S) = \emptyset (see also [16,19]). Therefore, the space L^{2} has the following orthogonal decomposition

    \begin{equation} \nonumber L^{2} = L^{-}\oplus L^{+}, \; \; \; z = z^{-}+z^{+} \end{equation}

    corresponding to the spectrum decomposition of S such that S is positive definite (resp. negative definite) in L^{+} (resp. L^{-} ).

    We use |S| to denote the absolute value of S , and |S|^{1/2} denotes the square root of |S| . Let E = \mathscr{D}(|S|^{1/2}) be the domain of the self-adjoint operator |S|^{1/2} , which is a Hilbert space equipped with the following inner product

    \begin{equation} \nonumber (z, w) = (|S|^{1/2}z, |S|^{1/2}w)_{2}, \ \hbox{for} \ z, w\in E, \end{equation}

    and the norm \|z\|^{2} = (z, z) . Evidently, E possesses the following decomposition

    \begin{equation} \nonumber E = E^{-}\oplus E^{+}, \; \; \hbox{where}\; \; E^{\pm} = E\cap L^{\pm}, \end{equation}

    which is orthogonal with respect to the two inner products (\cdot, \cdot)_{2} and (\cdot, \cdot) . Moreover, by using the polar decomposition of S we can obtain that

    \begin{equation} \nonumber Sz^{-} = -|S|z^{-}, \ Sz^{+} = |S|z^{+}\; \; \hbox{for all}\; z = z^{+}+z^{-}\in E. \end{equation}

    Furthermore, from [16] we have the embedding theorem, that is, E embeds continuously into L^{q} for each q\geq 2 and compactly into L^{q}_{loc} for all q\geq1 . Hence, there exists a constant \gamma_{q} > 0 such that for all z\in E

    \begin{eqnarray} \|z\|_{q}\leq \gamma_{q}\|z\|\; \hbox{for all}\; q\geq 2. \end{eqnarray} (2.1)

    From the assumptions (g_{1}) and (g_{2}) , we can deduce that for any \epsilon > 0 , there exists a positive constant C_{\epsilon} such that

    \begin{equation} |g(s)|\leq\epsilon+C_{\epsilon}|s|^{p-2}\; \; \hbox{and}\; \; |G(s)|\leq\epsilon|s|^{2}+C_{\epsilon}|s|^{p}\; \; \hbox{for each}\; \; s\in \mathbb{R}^{+}. \end{equation} (2.2)

    Next, we define the corresponding energy functional of system (1.3) on E by

    \begin{equation} \nonumber I_{\epsilon}(z) = \frac{1}{2}\int_{\mathbb{R}}Sz\cdot z\mathrm{d}t-\int_{\mathbb{R}}A(\epsilon t)G(|z|)\mathrm{d}t \end{equation}

    Applying the polar decomposition of S , then the energy functional I_{\epsilon} has another representation as follows

    \begin{equation} \nonumber I_{\epsilon}(z) = \frac{1}{2}(\|z^{+}\|^{2}-\|z^{-}\|^{2}) -\int_{\mathbb{R}}A(\epsilon t)G(|z|)\mathrm{d}t. \end{equation}

    Evidently, according to condition (L) , we can see that I_{\epsilon} is strongly indefinite. Furthermore, from conditions (g_{1}) and (g_{2}) we can infer that I_{\epsilon}\in C^{1}(E, \mathbb{R}) , and we have

    \begin{equation*} \langle I_{\epsilon}^\prime(z), \psi\rangle = (z^+, \psi^+)-(z^-, \psi^-)-\int_{\mathbb{R}}A(\epsilon t)g(|z|)z \psi\mathrm{d}t, \; \forall \psi\in E. \end{equation*}

    Making use of a standard argument we can check that critical points of I_{\epsilon} are homoclinic orbits of system (1.1).

    We shall make use of the techniques of the limit problem to prove the main results; in this section we introduce some related results for the autonomous system.

    For any constant \mu > 0 , in what follows we consider the autonomous system given by

    \begin{equation} Sz = \mu g(|z|)z, \; \; t\in \mathbb{R}. \end{equation} (3.1)

    Similarly, following the above comments, we define the energy functional I_{\mu} corresponding to system (3.1) as follows

    \begin{equation} \nonumber I_{\mu}(z) = \frac{1}{2}\left(\|z^{+}\|^{2}-\|z^{-}\|^{2}\right)-\mu\int_{\mathbb{R}}G(|z|)\mathrm{d}t. \end{equation}

    Evidently, we have

    \begin{equation*} \langle I_{\mu}^\prime(z), \psi\rangle = (z^+, \psi^+)-(z^-, \psi^-)-\mu\int_{\mathbb{R}}g(|z|)z \psi\mathrm{d}t, \; \forall \psi\in E. \end{equation*}

    In order to obtain the ground state homoclinic orbits of system (3.1), we will use the method of the generalized Nehari manifold developed by Szulkin and Weth [34]. To do this, we first introduce the following generalized Nehari manifold

    \begin{equation} \nonumber \mathscr{N}_{\mu} = \{z\in E\backslash E^{-}: \langle I'_{\mu}(z), z\rangle = 0\; \hbox{and}\; \langle I'_{\mu}(z), v\rangle = 0, \; \; \forall v\in E^{-}\}, \end{equation}

    and we define the ground state energy value c_{\mu} of I_{\mu} on \mathscr{N}_{\mu}

    \begin{equation} \nonumber c_{\mu} = \inf\limits_{z\in \mathscr{N}_{\mu}}I_{\mu}(z). \end{equation}

    Furthermore, for every z\in E\backslash E^{-} , we also need to define the subspace

    \begin{equation} \nonumber E(z) = E^{-}\oplus \mathbb{R}z = E^{-}\oplus \mathbb{R}z^{+}, \end{equation}

    and the convex subset

    \begin{equation} \nonumber \widehat{E}(z) = E^{-}\oplus [0, +\infty)z = E^{-}\oplus [0, +\infty)z^{+}. \end{equation}

    We note that E(z) and \widehat{E}(z) do not depend on \mu , but depend on the operator S .

    The main result in this section is the following theorem:

    Theorem 3.1. Assume that condition (L) holds and let (g_{1}) - (g_{3}) be satisfied, then, problem (3.1) has at least a ground state homoclinic orbit z\in\mathscr{N}_{\mu} such that I_{\mu}(z) = c_{\mu} > 0 .

    In this subsection, we are going to prove some technical results which will be used in the proof of Theorem 3.1. The following result involves the translation that will be used frequently in this paper, the proof can be found in [33, Lemma 2.1].

    Lemma 3.1. For all u = u^{+}+u^{-}\in E and y\in\mathbb{R} , if v(t): = u(t+y) , then v\in E with v^{+}(t) = u^{+}(t+y) and v^{-}(t) = u^{-}(t+y) .

    We give an important estimate, which plays a crucial role in the later proof.

    Lemma 3.2. Let z\in \mathscr{N}_{\mu} , then, for each v\in \mathcal{H}: = \{sz+w: s\geq -1, w\in E^{-}\} and v\neq 0 , we have the following energy estimate

    \begin{equation} \nonumber I_{\mu}(z+v) < I_{\mu}(z). \end{equation}

    Hence z is a unique global maximum of I_{\mu}|_{\widehat{E}(z)} .

    Proof. We follow the similar ideas explored in [34, Proposition 2.3.] to prove this lemma. Observe that, for any z\in \mathscr{N}_{\mu} , we directly obtain

    \begin{equation} \nonumber 0 = \langle I'_{\mu}(z), \varphi\rangle = (Az, \varphi)_{2}-\mu\int_{\mathbb{R}}g(|z|)z \varphi \mathrm{d}t \; \hbox{for all}\; \varphi\in E(z). \end{equation}

    Let v = sz+w\in\mathcal{H} , then, z+v = (1+s)z+w\in \widehat{E}(z) . By an elemental computation, we can get

    \begin{equation} \nonumber \begin{split} &I_{\mu}(z+v)-I_{\mu}(z)\\ & = \frac{1}{2}\bigg[\big(A(z+v), (z+v)\big)_{2}-\big(Az, z\big)_{2}\bigg]+\mu\int_{\mathbb{R}}\bigg[G(|z|)-G(|z+v|)\bigg]\mathrm{d}t\\ & = \frac{1}{2}\bigg[\big(A((1+s)z+w), (1+s)z+w\big)_{2}-\big(Az, z\big)_{2}\bigg]+\mu\int_{\mathbb{R}}\bigg[G(|z|)-G(|z+v|)\bigg]\mathrm{d}t\\ & = -\frac{\|w\|^{2}}{2}+ \big(Az, s(\frac{s}{2}+1)z+(1+s)w\big)_{2}+\mu\int_{\mathbb{R}}\bigg[G(|z|)-G(|z+v|)\bigg]\mathrm{d}t\\ & = -\frac{\|w\|^{2}}{2}+\mu\int_{\mathbb{R}}\Big[g(|z|)z\cdot \Big(s(\frac{s}{2}+1)z(t)+(1+s)w(t)\Big)+G(|z|)-G(|z+v|)\bigg]\mathrm{d}t\\ & = -\frac{\|w\|^{2}}{2}+\mu\int_{\mathbb{R}}\widetilde{g}(s, z, v)\mathrm{d}t, \end{split} \end{equation}

    where

    \begin{equation} \nonumber \widetilde{g}(s, z, v) = g(|z|)z\cdot \Big(s(\frac{s}{2}+1)z(t)+(1+s)w(t)\Big)+G(|z|)-G(|z+v|). \end{equation}

    Using (g_2) and (g_3) and combining the arguments used in [35] (see also [36]), we can conclude that \widetilde{g}(s, z, v) < 0 . Therefore, we have

    \begin{equation} \nonumber I_{\mu}(z+v) < I_{\mu}(z). \end{equation}

    Evidently, we know that z is a unique global maximum of I_{\mu}|_{\widehat{E}(z)} .

    Lemma 3.3. Assume that (g_{1}) and (g_{2}) are satisfied, then, we have the following two conclusions:

    \rm{(i)} there exists \rho > 0 such that c_{\mu} = \inf_{\mathscr{N}_{\mu}}I_{\mu}\geq \inf_{S_{\rho}}I_{\mu} > 0 , where

    S_{\rho}: = \{ z\in E^{+}:\|z\| = \rho\};

    \rm{(ii)} for any z\in \mathscr{N}_{\mu} , then \|z^{+}\|^{2}\geq \max\{\|z^{-}\|^{2}, 2c_{\mu}\} > 0 .

    Proof. (ⅰ) Let z\in E^{+} , we can deduce from (2.1) and (2.2) that

    \begin{equation} \nonumber I_{\mu}(z) = \frac{1}{2}\|z\|^{2}-\mu\int_{\mathbb{R}}G(|z|)\mathrm{d}t \geq \left(\frac{1}{2}-\epsilon \mu\gamma_{2}^{2}\right)\|z\|^{2}-\mu \gamma_{p}^{p}C_{\epsilon}\|z\|^{p}. \end{equation}

    Evidently, we can see that there is \rho > 0 , for \|z\| = \rho small enough such that \inf_{S_{\rho}}I_{\mu} > 0 .

    On the other hand, for each z\in \mathscr{N}_{\mu} , there exists a positive constant s such that s\|z\| = \rho , then sz\in \widehat{E}(z)\cap S_{\rho} . From Lemma 3.2, one can derive that

    \begin{equation} \nonumber I_{\mu}(z) = \max\limits_{v\in \widehat{E}(z)}I_{\mu}(v)\geq I_{\mu}(sz). \end{equation}

    Therefore, we have

    \begin{equation} \nonumber \inf\limits_{\mathscr{N}_{\mu}}I_{\mu}\geq \inf\limits_{S_{\rho}}I_{\mu} > 0, \end{equation}

    which shows that the conclusion (ⅰ) holds.

    (ⅱ) First we note that, from (g_{3}) , it follows that

    \begin{equation} \nonumber \frac{1}{2}g(s)s^{2}\geq G(s) > 0 \ \hbox{for all} \ s\in\mathbb{R}^{+}. \end{equation}

    For each z\in \mathscr{N}_{\mu} , combining this with the definition of c_{\mu} , we have

    \begin{equation} \nonumber 0 < c_{\mu} \leq \frac{1}{2}\|z^{+}\|^{2}-\frac{1}{2}\|z^{-}\|^{2}-\mu\int_{\mathbb{R}}G(|z|)\mathrm{d}t \leq \frac{1}{2}\|z^{+}\|^{2}-\frac{1}{2}\|z^{-}\|^{2}. \end{equation}

    Hence, we can derive that \|z^{+}\|^{2}\geq \max\{\|z^{-}\|^{2}, 2c_{\mu}\} > 0 . The proof is finished.

    Lemma 3.4. Assume that \Omega\subset E^{+}\setminus \{0\} is a compact subset, thus, there exists R > 0 such that I_{\mu} < 0 on E(z)\setminus B_{R}(0) for all z\in \Omega .

    Proof. The proof follows as in [34, Lemma 2.5], here, we omit the details.

    Following the result in [34] (see [34, Lemma 2.6]), we can establish the uniqueness of maximum point of I_{\mu} on the set \widehat{E}(z) .

    Lemma 3.5. For any z\in E\backslash E^{-} , then the set \mathscr{N}_{\mu}\cap \widehat{E}(z) consists of precisely one point \widetilde{m}_{\mu}(z)\neq0 , which is the unique global maximum of I_{\mu}|_{\widehat{E}(z)} .

    Proof. On account of Lemma 3.2, it is sufficient to prove that \mathscr{N}_{\mu}\cap \widehat{E}(z)\neq\emptyset . Since \widehat{E}(z) = \widehat{E}(z^{+}) , without loss of generality, we can suppose that z\in E^{+} and \|z\| = 1 . By Lemma 3.3-(ⅰ), we obtain that I_{\mu}(sz) > 0 for s\in (0, +\infty) small enough. Lemma 3.4 yields that I_{\mu}(sz) < 0 for sz\in \widehat{E}(z)\backslash B_{R}(0) . Consequently, we can deduce that 0 < \sup I_{\mu}(\widehat{E}(z)) < \infty . Because \widehat{E}(z) is convex and the functional I_{\mu} is weakly supper semi-continuous on \widehat{E}(z) , we conclude that there exists \widehat{z}\in \widehat{E}(z) such that I_{\mu}(\widehat{z}) = \sup I_{\mu}(\widehat{E}(z)) . This shows that \widehat{z} is a critical point of I_{\mu}|_{\widehat{E}(z)} ; therefore,

    \begin{equation} \nonumber \langle I'_{\mu}(\widehat{z}), \widehat{z}\rangle = \langle I'_{\mu}(\widehat{z}), \varphi\rangle = 0 \; \; \hbox{for all}\; \; \varphi\in \widehat{E}(z), \end{equation}

    hence, \widehat{z}\in \mathscr{N}_{\mu} . So, \widehat{z}\in \mathscr{N}_{\mu}\cap \widehat{E}(z) . The proof is finished.

    Combining Lemma 3.2 with Lemma 3.5, we obtain the following conclusion.

    Lemma 3.6. For each z\in E\backslash E^{-} , then, there is a unique pair (s^{\ast}, \varphi^{\ast}) with s^{\ast}\in (0, +\infty) and \varphi^{\ast}\in E^{-} such that s^{\ast}z+\varphi^{\ast}\in \mathscr{N}_{\mu}\cap \widehat{E}(z) and

    \begin{equation} \nonumber I_{\mu}(s^{\ast}z+\varphi^{\ast}) = \max\limits_{w\in \widehat{E}(z)}I_{\mu}(w). \end{equation}

    Moreover, if z\in\mathscr{N}_{\mu} , then we have that s^{*} = 1 and \varphi^{\ast} = z^{-} .

    Lemma 3.7. I_{\mu} is coercive on \mathscr{N}_{\mu} , that is, I_{\mu}(z)\to+\infty as \|z\|\to+\infty , z\in\mathscr{N}_{\mu} .

    Proof. Seeking for a contradiction, assume that there exists \{z_{n}\}\subset \mathscr{N}_{\mu} such that

    \begin{equation} \nonumber I_{\mu}(z_{n})\leq \widehat{c}\; \; \hbox{for some}\; \; \widehat{c}\in[c_{\mu}, +\infty)\; \; \hbox{as}\; \; \|z_{n}\|\rightarrow +\infty. \end{equation}

    Setting w_{n}: = z_{n}/\|z_{n}\| , we obtain that \|z^{+}_{n}\|\geq \|z^{-}_{n}\| from Lemma 3.3(ⅱ), then, for every n\in \mathbb{N} , we get that \|w^{+}_{n}\|^{2}\geq \|w^{-}_{n}\|^{2} and \|w^{+}_{n}\|^{2}\geq \frac{1}{2} . There exist \{y_{n}\}\subset \mathbb{Z} , r > 0 and \delta > 0 such that

    \begin{equation} \int_{B_{r}(y_{n})}|w^{+}_{n}|^{2}\mathrm{d}t\geq \delta, \; \; \forall n\in \mathbb{N}. \end{equation} (3.2)

    If this is not true, then according to Lions' concentration-compactness principle, we can conclude that w^{+}_{n}\rightarrow 0 in L^{q}(\mathbb{R}) for q > 2 . Combining (2.1) and (2.2), we know that, for every s > 0 ,

    \begin{equation} \nonumber \mu\int_{\mathbb{R}}G(|sw^{+}_{n}|)\mathrm{d}t\leq \epsilon \mu \gamma_{2}^{2}s^{2}\|w^{+}_{n}\|^{2} +\mu C_{\epsilon}\gamma_{p}^{p}s^{p}\|w^{+}_{n}\|^{p}\rightarrow 0, \end{equation}

    then we get

    \begin{equation} \nonumber \begin{aligned} \widehat{c}&\geq I_{\mu}(sw^{+}_{n}) = \frac{1}{2}s^{2}\|w^{+}_{n}\|^{2}-\mu\int_{\mathbb{R}}G(|sw^{+}_{n}|)\mathrm{d}t\\ &\geq \frac{s^{2}}{4}-\mu\int_{\mathbb{R}}G(|sw^{+}_{n}|)\mathrm{d}t\rightarrow \frac{s^{2}}{4}. \end{aligned} \end{equation}

    This yields a contradiction if s > \sqrt{4\widehat{c}} ; hence we prove that (3.2) holds.

    Let us define \tilde{z}_{n}(t): = z_{n}(t+y_{n}) , and \tilde{w}_{n}(t): = w_{n}(t+y_{n}) , then, \tilde{w}_{n}^{+}\rightharpoonup \tilde{w}^{+} , and (3.2) yields that \tilde{w}^{+}\neq 0 . Since \tilde{z}_{n}(t) = \tilde{w}_{n}(t)\|\tilde{z}_{n}\| , it follows that \tilde{z}_{n}(t)\rightarrow +\infty almost everywhere in \mathbb{R} as \|\tilde{z}_{n}\| = \|z_{n}\|\rightarrow +\infty . Applying the Fatou's lemma, we can derive that

    \begin{equation} \nonumber \int_{\mathbb{R}}\frac{G(|z_{n}|)}{\|z_{n}\|^{2}}\mathrm{d}t = \int_{\mathbb{R}}\frac{G(|\tilde{z}_{n}|)}{\|\tilde{z}_{n}\|^{2}}\mathrm{d}t = \int_{\mathbb{R}}\frac{G(|\tilde{z}_{n}|)}{|\tilde{z}_{n}|^{2}}|\tilde{w}_{n}|^{2}\mathrm{d}t \geq \int_{[\tilde{z}_{n}\neq 0]}\frac{G(|\tilde{z}_{n}|)}{|\tilde{z}_{n}|^{2}}|\tilde{w}_{n}|^{2}\mathrm{d}t\rightarrow +\infty, \end{equation}

    where [\tilde{z}_{n}\neq 0] is the Lebesgue measure of the set \{t\in \mathbb{R}: \tilde{z}_{n}(t)\neq 0\} . Therefore

    \begin{equation} \nonumber \begin{aligned} 0&\leq \frac{I_{\mu}(z_{n})}{\|z_{n}\|^{2}} = \frac{1}{2}\|w_{n}^{+}\|^{2}-\frac{1}{2}\|w_{n}^{-}\|^{2} -\mu\int_{\mathbb{R}}\frac{G(|z_{n}|)}{\|z_{n}\|^{2}}\mathrm{d}t\\ &\leq \frac{1}{2}-\mu\int_{\mathbb{R}}\frac{G(|\tilde{z}_{n}|)}{|\tilde{z}_{n}|^{2}}|\tilde{w}_{n}|^{2}\mathrm{d}t \rightarrow -\infty, \end{aligned} \end{equation}

    we get a contradiction. The proof is finished.

    We want to utilize the method of the generalized Nehari manifold to prove the main result. To do this, we set S^{+}: = \{z\in E^{+}: \|z\| = 1\} in E^{+} , and we define the following mapping

    \begin{equation} \nonumber \widetilde{m}_{\mu}:E^{+}\backslash\{0\}\to\mathscr{N}_{\mu}\; \hbox{and}\; m_{\mu} = \widetilde{m}_{\mu}|_{S^{+}}, \end{equation}

    and the inverse of m_{\mu} is

    \begin{equation} \nonumber m^{-1}_{\mu}:\mathscr{N}_{\mu}\to S^{+}, \; m^{-1}_{\mu}(z) = z^{+}/\|z^{+}\|. \end{equation}

    Following from the proof of [34, Lemma 2.8], we can see that \widetilde{m}_{\mu} is continuous and m_{\mu} is a homeomorphism.

    We now consider the reduced functionals

    \begin{equation} \nonumber \widetilde{\Phi}_{\mu}(z) = I_{\mu}(\widetilde{m}_{\mu}(z))\; \hbox{and}\; \Phi_{\mu} = \widetilde{\Phi}_{\mu}|_{S^{+}}. \end{equation}

    which is continuous since \widetilde{m}_{\mu} is continuous.

    The following results establish some crucial properties involving the reduced functionals \widetilde{\Phi}_{\mu} and \Phi_{\mu} , which play important roles in our arguments. And their proofs follow the proofs of [34, Proposition 2.9, Corollary 2.10].

    Lemma 3.8. The following conclusions are true:

    \rm{(a)} \widetilde{\Phi}_{\mu}\in C^{1}(E^{+}\backslash\{0\}, \mathbb{R}) and for z, v\in E^{+} and z\neq0 ,

    \begin{equation} \nonumber \langle\widetilde{\Phi}'_{\mu}(z), v\rangle = \frac{\|\widetilde{m}_{\mu}(z)^{+}\|}{\|z\|}\langle I'_{\mu}(\widetilde{m}_{\mu}(z)), v\rangle. \end{equation}

    \rm{(b)} \Phi_{\mu}\in C^{1}(S^{+}, \mathbb{R}) and for each z\in S^{+} and v\in T_{z}(S^{+}) = \{u\in E^{+}:(z, u) = 0\} ,

    \begin{equation} \nonumber \langle \Phi'_{\mu}(z), v\rangle = \|\widetilde{m}_{\mu}(z)^{+}\|\langle I'_{\mu}(\widetilde{m}_{\mu}(z)), v\rangle. \end{equation}

    \rm{(c)} \{z_n\} is a (PS)-sequence for \Phi_{\mu} if and only if \{\widetilde{m}_{\mu}(z_n)\} is a (PS)-sequence for I_{\mu} .

    \rm{(d)} We have

    \begin{equation} \nonumber \inf\limits_{S^{+}}\Phi_{\mu} = \inf\limits_{\mathscr{N}_{\mu}}I_{\mu} = c_{\mu}. \end{equation}

    Moreover, z\in S^{+} is a critical point of \Phi_{\mu} if and only if \widetilde{m}_{\mu}(z)\in \mathscr{N}_{\mu} is a critical point of I_{\mu} and the corresponding critical values coincide.

    Based on the above preliminaries, in this subsection we give the complete proof of Theorem 3.1, and further study the monotonicity and continuity of the ground-state energy c_{\mu} .

    Proof of Theorem 3.1: According to Lemma 3.3, it is easy to see that c_{\mu} > 0 . We note that, if z\in \mathscr{N}_{\mu} with I_{\mu}(z) = c_{\mu} , then m^{-1}_{\mu}(z)\in S^{+} is a minimizer of \Phi_{\mu} ; hence, it is a critical point of \Phi_{\mu} . Consequently, Lemma 3.8 yields that z is a critical point of I_{\mu} . We have to prove that there exists a minimizer \tilde{z}\in \mathscr{N}_{\mu} such that I_{\mu}(\tilde{z}) = c_{\mu} . Actually, Ekeland's variational principle yields that there exists a sequence \{v_{n}\}\subset S^{+} such that \Phi_{\mu}(v_n)\to c_{\mu} and \Phi'_{\mu}(v_n)\to 0 as n\to\infty . For all n\in\mathbb{N} , setting z_n = \widetilde{m}_{\mu}(v_n)\in \mathscr{N}_{\mu} , then I_{\mu}(z_n)\to c_{\mu} and I'_{\mu}(z_n)\to 0 by Lemma 3.8. By virtue of Lemma 3.7, we can see that \{z_n\} is bounded in E . Moreover, it satisfies

    \begin{equation} \nonumber \varliminf_{n\rightarrow \infty}\sup \limits_{y\in {\mathbb R}} \int_{B_{1}(y)}|z_{n}|^{2}\mathrm{d}t > 0. \end{equation}

    If this is not true, then Lions' concentration-compactness principle implies that z_n\to 0 in L^{q}(\mathbb{R}) for any q > 2 . Therefore, from (2.1) and (2.2), we can derive that

    \begin{equation} \nonumber \int_{\mathbb{R}}\left[\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\right]\mathrm{d}t = o_{n}(1). \end{equation}

    Then, we get

    \begin{equation} \nonumber \begin{aligned} c_{\mu}+o_{n}(1)& = I_{\mu}(z_{n})-\frac{1}{2}\langle I_{\mu}'(z_{n}), z_{n}\rangle\\ & = \mu\int_{\mathbb{R}}\bigg[\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\bigg]\mathrm{d}t = o_{n}(1). \end{aligned} \end{equation}

    Since c_{\mu} > 0 , obviously we get a contradiction. Thus, there exist \{y_n\}\subset\mathbb{Z} and \delta > 0 such that

    \begin{equation} \nonumber \int_{B_{2}(y_{n})}|z_n|^{2}\mathrm{d}t\geq\delta. \end{equation}

    Let us define \tilde{z}_n(t) = z_{n}(t+y_n) , then, we have

    \begin{equation} \int_{B_{2}(0)}|\tilde{z}_n|^{2}\mathrm{d}t\geq\delta. \end{equation} (3.3)

    Observe that I_{\mu} is the invariant under translation since (3.1) is autonomous, then, we have \|\tilde{z}_n\| = \|z_n\| and

    \begin{equation} I_{\mu}(\tilde{z}_n)\to c_{\mu}\; \; \hbox{and}\; \; I_{\mu}'(\tilde{z}_n)\to 0. \end{equation} (3.4)

    Passing to a subsequence, we may suppose that \tilde{z}_n \rightharpoonup \tilde{z} in E , \tilde{z}_n \to \tilde{z} in L_{loc}^{q}(\mathbb{R}) for q > 2 , and \tilde{z}_n(t) \to \tilde{z}(t) almost everywhere on \mathbb{R} . According to (3.3) and (3.4), then we can derive that \tilde{z}\neq 0 and I_{\mu}'(\tilde{z}) = 0 . This implies that \tilde{z}\in \mathscr{N}_{\mu} and I_{\mu}(\tilde{z})\geq c_{\mu} . On the other hand, applying Fatou's lemma we can obtain

    \begin{equation} \nonumber \begin{aligned} c_{\mu}& = \lim\limits_{n\to\infty}\left(I_{\mu}(\tilde{z}_n)-\frac{1}{2}\langle I_{\mu}'(\tilde{z}_n), \tilde{z}_n\rangle\right)\\ & = \lim\limits_{n\to \infty}\mu\int_{\mathbb{R}}\bigg(\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\bigg)\mathrm{d}t\\ &\geq \mu\int_{\mathbb{R}}\lim\limits_{n\to \infty}\bigg(\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\bigg)\mathrm{d}t\\ & = I_{\mu}(\tilde{z})-\frac{1}{2}\langle I_{\mu}'(\tilde{z}), \tilde{z}\rangle = I_{\mu}(\tilde{z}), \end{aligned} \end{equation}

    that is, I_{\mu}(\tilde{z})\leq c_{\mu} . Consequently, I_{\mu}(\tilde{z}) = c_{\mu} and \tilde{z} is a critical point of I_{\mu} , which implies that \tilde{z} is a ground-state homoclinic orbit of problem (3.1). So, we have completed the proof of Theorem 3.1.

    As a byproduct of the Theorem 3.1, we show the monotonicity and continuity of c_{\mu} .

    Lemma 3.9. The function \mu \mapsto c_{\mu} is decreasing and continuous on (0, +\infty) .

    Proof. In what follows, let z_{\mu_{1}} and z_{\mu_{2}} be as ground state homoclinic orbits of I_{\mu_{1}} and I_{\mu_{2}} , respectively. Assume that \mu_{1} > \mu_{2} . First of all, we want to verify that the function \mu \mapsto c_{\mu} is decreasing. On account of Lemma 3.6 we can find that there exist s_{1} > 0 and \varphi_{1}\in E^{-} such that

    \begin{equation} \nonumber I_{\mu_{1}}(s_{1}z_{\mu_{2}}+\varphi_{1}) = \max\limits_{z\in \widehat{E}(z_{\mu_{2}})}I_{\mu_{1}}(z), \end{equation}

    then we have

    \begin{equation} \nonumber \begin{aligned} c_{\mu_{1}}&\leq I_{\mu_{1}}(s_{1}z_{\mu_{2}}+\varphi_{1})\\ & = I_{\mu_{2}}(s_{1}z_{\mu_{2}}+\varphi_{1}) +(\mu_{2}-\mu_{1})\int_{\mathbb{R}}G(|s_{1}z_{\mu_{2}}+\varphi_{1}|)\mathrm{d}t\\ &\leq I_{\mu_{2}}(z_{\mu_{2}}) +(\mu_{2}-\mu_{1})\int_{\mathbb{R}}G(|s_{1}z_{\mu_{2}}+\varphi_{1}|)\mathrm{d}t\\ & = c_{\mu_{2}}+(\mu_{2}-\mu_{1})\int_{\mathbb{R}}G(|s_{1}z_{\mu_{2}}+\varphi_{1}|)\mathrm{d}t. \end{aligned} \end{equation}

    Combining the fact that

    \begin{equation} \nonumber \int_{\mathbb{R}}G(|s_{1}z_{\mu_{2}}+\varphi_{1}|)\mathrm{d}t\geq 0, \end{equation}

    with the inequality \mu_{2}-\mu_{1} < 0 , we can infer that

    \begin{equation} \nonumber c_{\mu_{1}}\leq c_{\mu_{2}}. \end{equation}

    We finish the proof by demonstrating that the function \mu \mapsto c_{\mu} is decreasing on (0, +\infty) .

    In order to claim the continuity of c_{\mu} , we divide the proof into two steps:

    Step 1: Let \{\mu_{n}\} be a sequence such that \mu_{1}\leq \mu_{2}\leq \cdots \leq \mu_{n}\leq \cdots \leq \mu and \mu_{n}\to\mu .

    Claim 1: c_{\mu_{n}}\to c_{\mu} as n\to\infty .

    Indeed, let z_{\mu} be the ground state homoclinic orbit of system (3.1). On account of Lemma 3.6, we can see that there exist s_{n} > 0 and \varphi_{n}\in E^{-} such that

    \begin{equation} \nonumber I_{\mu_{n}}(s_{n}z_{\mu}+\varphi_{n}) = \max\limits_{z\in \widehat{E}(z_{\mu})}I_{\mu_{n}}(z)\; \; \hbox{for all}\; \; n\in \mathbb{N}. \end{equation}

    Note that, using (g_3) and computing directly, we get

    \begin{equation} \nonumber I_{\mu_{1}}(z)-I_{\mu_{n}}(z) = (\mu_{n}-\mu_{1})\int_{\mathbb{R}}G(|z|)\mathrm{d}t\geq 0, \end{equation}

    so, for all n\in \mathbb{N} and z\in E , we have that I_{\mu_{1}}(z)\geq I_{\mu_{n}}(z) . Then by Lemma 3.4, it holds that there exists R > 0 such that

    \begin{equation} I_{\mu_{n}}(z)\leq I_{\mu_{1}}(z)\leq 0, \; \; \forall z\in \widehat{E}(z_{\mu})\backslash B_{R}(0). \end{equation} (3.5)

    According to Lemma 3.3 and the monotonicity of c_{\mu} , we can obtain

    \begin{equation} \nonumber I_{\mu_{n}}(s_{n}z_{\mu}+\varphi_{n}) = \max\limits_{z\in \widehat{E}(z_{\mu})}I_{\mu_{n}}(z)\geq c_{\mu_{n}}\geq c_{\mu} > 0, \end{equation}

    consequently, it follows that

    \begin{equation} I_{\mu_{n}}(s_{n}z_{\mu}+\varphi_{n}) > 0. \end{equation} (3.6)

    From (3.5) and (3.6), one can check that \|s_{n}z_{\mu}+\varphi_{n}\|\leq R ; this shows that the sequence \{s_{n}z_{\mu}+\varphi_{n}\} is bounded in E . Hence, it is easy to see that

    \int_{\mathbb{R}}G(|s_{n}z_{\mu}+\varphi_{n}|)\mathrm{d}t\; \; \hbox{is also bounded},

    then we get

    \begin{equation} \nonumber \begin{aligned} c_{\mu_{n}}&\leq I_{\mu_{n}}(s_{n}z_{\mu}+\varphi_{n})\\ & = I_{\mu}(s_{n}z_{\mu}+\varphi_{n}) +(\mu-\mu_{n})\int_{\mathbb{R}}G(|s_{n}z_{\mu}+\varphi_{n}|)\mathrm{d}t\\ &\leq I_{\mu}(z_{\mu})+(\mu-\mu_{n})\int_{\mathbb{R}}G(|s_{n}z_{\mu}+\varphi_{n}|)\mathrm{d}t\\ & = c_{\mu}+o_{n}(1). \end{aligned} \end{equation}

    On the other hand, since c_{\mu}\leq c_{\mu_{n}} for all n\in \mathbb{N} , we can infer that

    \begin{equation} \nonumber c_{\mu_{n}}\to c_{\mu} \ \hbox{as} \ n\to \infty. \end{equation}

    Step 2: Let \{\mu_{n}\} be a sequence such that \mu_{1}\geq \mu_{2}\geq \cdots \geq \mu_{n}\geq \cdots \geq \mu and \mu_{n}\to\mu .

    Claim 2: c_{\mu_{n}}\to c_{\mu} as n\to\infty .

    In fact, let z_{n} be the ground state homoclinic orbit of the system (3.1) with \mu = \mu_{n} , then, there exist s_{n} > 0 and \varphi_{n}\in E^{-} such that

    \begin{equation} \nonumber I_{\mu}(s_{n}z_{n}+\varphi_{n}) = \max\limits_{z\in \widehat{E}(z_{n})}I_{\mu}(z). \end{equation}

    We can easily obtain that the sequence \{z_{n}\} is bounded by Lemma 3.7. Moreover, we can find that there exist \delta > 0 , r > 0 and \{y_{n}\}\subset \mathbb{Z} such that for each n\in \mathbb{N} , we have

    \begin{equation} \int_{B_{r}(y_{n})}|z_{n}^{+}|^{2}\mathrm{d}t\geq \delta. \end{equation} (3.7)

    Otherwise, using Lions' concentration-compactness principle we can deduce that z_{n}^{+}\rightarrow 0 in L^{q}(\mathbb{R}) for all q > 2 . Combining (2.1) with (2.2), it holds that

    \begin{equation} \nonumber \int_{\mathbb{R}}g(|z_{n}|)z_{n} z_{n}^{+}\mathrm{d}t\rightarrow 0. \end{equation}

    Therefore, we have

    \begin{equation} \nonumber \begin{aligned} 0& = \langle I'_{\mu_{n}}(z_{n}), z_{n}^{+}\rangle = \|z_{n}^{+}\|^{2}-\mu_{n}\int_{\mathbb{R}}g(|z_{n}|)z_{n} z_{n}^{+}\mathrm{d}t\\ & = \|z_{n}^{+}\|^{2}+o_{n}(1), \end{aligned} \end{equation}

    this shows that \|z_{n}^{+}\|^{2}\rightarrow 0 , which is a contradiction to the inequality \|z_{n}^{+}\|^{2}\geq 2c_{\mu_{n}} > 0 from Lemma 3.3. So, (3.7) holds.

    Setting \tilde{z}_{n}(t): = z_{n}(t+y_{n}) , one can check that \{\tilde{z}_{n}\} is bounded in E ; passing to a subsequence, \tilde{z}_{n}^{+}\rightharpoonup \tilde{z}^{+}\neq 0 in E . Set \mathcal{V}: = \{\tilde{z}_{n}^{+}\}\subset E^{+}\backslash \{0\} ; hence, \mathcal{V} is bounded and the sequence does not weakly converge to zero in E . Then by Lemma 3.4, there exists R > 0 such that for every z\in \mathcal{V} , we obtain

    \begin{equation} I_{\mu}(w) < 0, \; \hbox{for}\; w\in E(z)\backslash B_{R}(0). \end{equation} (3.8)

    Define \tilde{\varphi}_{n}(t): = \varphi_{n}(t+y_{n}) , we have

    \begin{equation} I_{\mu}(s_{n}\tilde{z}_{n}+\tilde{\varphi}_{n}) = I_{\mu}(s_{n}z_{n}+\varphi_{n}) = \max\limits_{z\in \widehat{E}(z_{n})}I_{\mu}(z)\geq c_{\mu} > 0, \; \; \; \forall n\in \mathbb{N}. \end{equation} (3.9)

    In view of (3.8) and (3.9), we can conclude that \|s_{n}\tilde{z}_{n}+\tilde{\varphi}_{n}\|\leq R for all n\in \mathbb{N} , then, \|s_{n}z_{n}+\varphi_{n}\|\leq R , which implies that the sequence \{s_{n}z_{n}+\varphi_{n}\} is bounded in E , and \int_{\mathbb{R}}G(|s_{n}z_{n}+\varphi_{n}|)\mathrm{d}t is also bounded. Thus, we obtain

    \begin{equation} \nonumber \begin{aligned} c_{\mu}&\leq I_{\mu}(s_{n}z_{n}+\varphi_{n})\\ & = I_{\mu_{n}}(s_{n}z_{n}+\varphi_{n}) +(\mu_{n}-\mu)\int_{\mathbb{R}}G(|s_{n}z_{n}+\varphi_{n}|)\mathrm{d}t\\ &\leq I_{\mu_{n}}(z_{n}) +(\mu_{n}-\mu)\int_{\mathbb{R}}G(|s_{n}z_{n}+\varphi_{n}|)\mathrm{d}t\\ & = c_{\mu_{n}}+o_{n}(1). \end{aligned} \end{equation}

    Combining this with the fact that c_{\mu}\geq c_{\mu_{n}} for all n\in \mathbb{N} , then we have

    \begin{equation} \nonumber c_{\mu_{n}}\to c_{\mu} \ \hbox{as} \ n\to \infty. \end{equation}

    We have finished the proof of the lemma.

    In this subsection we will give the proof involving the existence result for ground state homoclinic orbits for system (1.1). As before, we define the associated generalized Nehari manifold

    \begin{equation} \nonumber \mathscr{N}_{\epsilon}: = \{z\in E\backslash E^{-}: \langle I'_{\epsilon}(z), z\rangle = 0 \; \; \hbox{and}\; \; \langle I'_{\epsilon}(z), \varphi\rangle = 0, \forall \varphi\in E^{-}\} \end{equation}

    and the ground state energy value

    \begin{equation} \nonumber c_{\epsilon} = \inf\limits_{\mathscr{N}_{\epsilon}} I_{\epsilon}. \end{equation}

    Applying the same arguments explored in the Section 3 , we can show that for every z\in E\backslash E^{-} the set \mathscr{N}_{\epsilon}\cap \widehat{E}(z) is a singleton set, and the element of this set is the unique global maximum of I_{\epsilon}|_{\widehat{E}(z)} , that is, there exists a unique pair t > 0 and \varphi \in E^{-} such that

    \begin{equation} \nonumber I_{\epsilon}(tz+\varphi) = \max\limits_{w\in \widehat{E}(z)}I_{\epsilon}(w). \end{equation}

    Therefore, the following mapping is well-defined:

    \begin{equation} \nonumber \widetilde{m}_{\epsilon}:E^{+}\backslash\{0\}\to\mathscr{N}_{\epsilon}\; \hbox{and}\; m_{\epsilon} = \widetilde{m}_{\epsilon}|_{S^{+}}, \end{equation}

    and the inverse of m_{\epsilon} is

    \begin{equation} \nonumber m^{-1}_{\epsilon}:\mathscr{N}_{\epsilon}\to S^{+}, \; m^{-1}_{\epsilon}(z) = z^{+}/\|z^{+}\|. \end{equation}

    Accordingly, the reduced functional \widetilde{\Phi}_{\epsilon}:E^{+}\backslash\{0\}\to\mathbb{R} and the restriction \Phi_{\epsilon}: S^{+}\to\mathbb{R} can respectively be defined by

    \begin{equation} \nonumber \widetilde{\Phi}_{\epsilon}(z) = I_{\epsilon}(\widetilde{m}_{\epsilon}(z))\; \hbox{and}\; \Phi_{\epsilon} = \widetilde{\Phi}_{\epsilon}|_{S^{+}}. \end{equation}

    Moreover, from the above discussions in Section 3 , we can check that all related conclusions in Section 3 hold for I_{\epsilon} , c_{\epsilon} , \mathscr{N}_{\epsilon} , \widetilde{m}_{\epsilon} , m_{\epsilon} , \widetilde{\Phi}_{\epsilon} and \Phi_{\epsilon} , respectively.

    Meanwhile, concerning the limit problem given by

    \begin{equation} Sz = A(0)g(|z|)z, \; \; t\in \mathbb{R}, \end{equation} (4.1)

    for the sake of simplicity, we will use the notations I_{0} , c_{0} and \mathscr{N}_{0} to denote I_{A(0)} , c_{A(0)} and \mathscr{N}_{A(0)} , respectively.

    Next, we will state the relationship of the ground state energy value between system (1.3) and limit system (4.1), and this is very significant in our following arguments.

    Lemma 4.1. The limit \lim\limits_{\epsilon\rightarrow 0}c_{\epsilon} = c_{0} holds.

    Proof. Let be \epsilon_{n}\to 0 as n\to\infty . Evidently, using Lemma 3.9 we obtain that c_{0}\leq c_{\epsilon_{n}} for all n\in\mathbb{N} ; thus, c_{0}\leq\liminf_{n\to\infty}c_{\epsilon_{n}} .

    On the other hand, Theorem 3.1 shows that the limit system (4.1) has a ground state homoclinic orbit z_{0} . Then, according to Lemma 3.6, we can find that there are s_{n}\in(0, +\infty) and \varphi_{n}\in E^{-} such that s_{n}z_{0}^{+}+\varphi_{n}\in \mathscr{N}_{\epsilon_{n}} , and

    \begin{equation} \nonumber I_{\epsilon_{n}}(s_{n}z_{0}^{+}+\varphi_{n})\geq c_{\epsilon_{n}}\geq c_{0} > 0, \; \; \; \forall n\in \mathbb{N}. \end{equation}

    As in the previous section, we can see that \{s_{n}z_{0}^{+}+\varphi_{n}\} is bounded in E . Thus, without loss of generality, we assume that s_{n}\rightarrow s_{0} and \varphi_{n}\rightharpoonup \varphi in E^{-} . Therefore, we can deduce from the weakly lower semi-continuity of the norm and Fatou's Lemma that

    \begin{equation} \nonumber \begin{aligned} c_{0} = &\liminf\limits_{n\rightarrow \infty}c_{\epsilon_{n}}\leq \limsup\limits_{n\rightarrow \infty}c_{\epsilon_{n}}\leq\limsup\limits_{n\rightarrow \infty}I_{\epsilon_{n}}(s_{n}z_{0}^{+}+\varphi_{n})\\ \leq& \limsup\limits_{n\rightarrow \infty}\bigg[\frac{1}{2}s_{n}^{2}\|z_{0}^{+}\|^{2}-\frac{1}{2}\|\varphi_{n}\|^{2} -\int_{\mathbb{R}}A(\epsilon_{n}t)G(|s_{n}z_{0}^{+}+\varphi_{n}|)\mathrm{d}t\bigg]\\ \leq& \frac{1}{2}s_{0}^{2}\|z_{0}^{+}\|^{2}-\frac{1}{2}\|\varphi\|^{2}-A(0)\int_{\mathbb{R}}G(|s_{0}z_{0}^{+}+\varphi|)\mathrm{d}t\\ = &I_{0}(s_{0}z_{0}^{+}+\varphi)\leq I_{0}(z_{0}) = c_{0}. \end{aligned} \end{equation}

    Obviously, we can get

    \begin{equation} \nonumber \lim\limits_{n\rightarrow \infty}c_{\epsilon_{n}} = c_{0}, \end{equation}

    finishing the proof.

    In view of the above discussion, we obtain that I_{0}(s_{0}z_{0}^{+}+\varphi) = I_{0}(z_{0}) = c_{0} ; then, both s_{0}z_{0}^{+}+\varphi and z_{0} are the elements of \mathscr{N}_{0}\cap \widehat{E}(z_{0}) . But, according to Lemma 3.6, there is only one element in \mathscr{N}_{0}\cap \widehat{E}(z_{0}) , so we can conclude that s_{0}z_{0}^{+}+\varphi = z_{0} and s_{n}\rightarrow s_{0} = 1 , where \varphi_{n}\rightharpoonup \varphi = z_{0}^{-} .

    As a byproduct of the Lemma 4.1, we can directly obtain the following result.

    Lemma 4.2. Assume that condition (A) holds, then, there is \epsilon_{0} > 0 such that c_{\epsilon} < c_{A_{\infty}} for \epsilon\in (0, \epsilon_{0}) .

    Proof. From condition (A) we can see that A(0) > A_{\infty} . Then from Lemma 3.9 we have that c_{0} < c_{A_{\infty}} . Observe that, Lemma 4.1 yields that there exists \epsilon_{0} > 0 small enough such that c_{\epsilon} < c_{A_{\infty}} for all \epsilon\in (0, \epsilon_{0}) . Therefore, we get that c_{\epsilon} < c_{A_{\infty}} for \epsilon\in (0, \epsilon_{0}) .

    Using similar arguments as for the proof of Lemma 3.7, one can easily check the following lemma.

    Lemma 4.3. The energy functional I_{\epsilon} is coercive on \mathscr{N}_{\epsilon} for each \epsilon\geq 0 .

    Next we give the proof involving the existence result for ground state homoclinic orbits for system (1.1).

    Lemma 4.4. Assume that conditions (L) , (A) and (g_1) - (g_3) are satisfied, then, system (1.1) has a ground-state homoclinic orbit for each \epsilon \in(0, \epsilon_{0}) .

    Proof. Following the proof of Theorem 3.1 and using Lemma 3.8, we must prove that there exists z\in\mathscr{N}_{\epsilon} such that I_{\epsilon}(z) = c_{\epsilon} . Indeed, applying Ekeland's variational principle, there exists \{u_{n}\}\subset S^{+} such that \Phi_{\epsilon}(u_n)\to c_{\epsilon} and \Phi'_{\epsilon}(u_n)\to 0 . Put z_n = \widetilde{m}_{\epsilon}(u_n)\in \mathscr{N}_{\epsilon} for all n\in\mathbb{N} . Then from Lemma 3.8 we have that I_{\epsilon}(z_n)\to c_{\epsilon} and I'_{\epsilon}(z_n)\to 0 . Furthermore, in view of Lemma 4.3, we can prove that \{z_n\} is bounded. Then, up to a subsequence, we can suppose that z_{n}\rightharpoonup z in E . Evidently, I'_{\epsilon}(z) = 0 .

    In what follows we need to show that z\neq 0 and I_{\epsilon}(z) = c_{\epsilon} . Combining the fact that z_n\in \mathscr{N}_{\epsilon} with Lemma 3.3, we have

    \begin{equation} \nonumber \begin{aligned} o_{n}(1)& = \langle I'_{\epsilon}(z_{n}), z_{n}^{+}\rangle = \|z_{n}^{+}\|^{2}-\int_{\mathbb{R}}A(\epsilon t)g(|z_{n}|)z_{n}z_{n}^{+}\mathrm{d}t\\ &\geq 2c_{\epsilon}-\int_{\mathbb{R}}A(\epsilon t)g(|z_{n}|)z_{n}z_{n}^{+}\mathrm{d}t, \end{aligned} \end{equation}

    which yields that

    \begin{equation} \nonumber \int_{\mathbb{R}}A(\epsilon t)g(|z_{n}|)z_{n}z_{n}^{+}\mathrm{d}t\geq 2c_{\epsilon} > 0. \end{equation}

    As in the previous section, we can check that there exists a sequence \{y_{n}\}\subset \mathbb{Z} , r > 0 and \delta > 0 such that

    \begin{equation} \int_{B_{r}(y_{n})}|z^{+}_{n}|^{2}\mathrm{d}t\geq \delta, \; \; \forall n\in \mathbb{N}. \end{equation} (4.2)

    Now, we need to prove that the sequence \{y_{n}\} is bounded in \mathbb{R} . Arguing by contradiction we can suppose that \{y_{n}\} is unbounded and |y_{n}|\rightarrow +\infty as n\rightarrow \infty . Setting w_{n}(t): = z_{n}(t+y_{n}) , then, w_{n}\rightharpoonup w in E ; we can obtain w\neq 0 from (4.2). By choosing the test function \psi\in C_{0}^{\infty}(\mathbb{R}) , we get

    \begin{equation} \begin{aligned} o_{n}(1) = &\langle I'_{\epsilon}(z_{n}), \psi(t-y_{n})\rangle\\ = &\left(z^{+}_{n}, \psi^{+}(t-y_{n})\right)-\left(z^{-}_{n}, \psi^{-}(t-y_{n})\right)-\int_{\mathbb{R}}A(\epsilon t)g(|z_{n}|)z_{n} \psi(t-y_{n})\mathrm{d}t\\ = &\left(w^{+}_{n}, \psi^{+}\right)-\left(w^{-}_{n}, \psi^{-}\right)-\int_{\mathbb{R}}A(\epsilon t+\epsilon y_{n})g(|w_{n}|)w_{n} \psi(t)\mathrm{d}t. \end{aligned} \end{equation} (4.3)

    Letting n\rightarrow +\infty , then we obtain

    \begin{equation} \left(w^{+}, \psi^{+}\right)-\left(w^{-}, \psi^{-}\right)-\int_{\mathbb{R}}A_{\infty}g(|w|)w \psi(t)\mathrm{d}t = \langle I'_{\infty}(w), \psi \rangle = 0. \end{equation} (4.4)

    This shows that w is a nontrivial solution of system (3.1) with \mu = A_{\infty} and w\in \mathscr{N}_{A_{\infty}} .

    Employing the Fatou's lemma we can derive that

    \begin{equation} \nonumber \begin{aligned} c_{A_{\infty}}&\leq I_{A_{\infty}}(w) = I_{A_{\infty}}(w)-\frac{1}{2}\langle I'_{A_{\infty}}(w), w\rangle\\ & = \int_{\mathbb{R}}A_{\infty}\left[\frac{1}{2}g(|w|)|w|^{2}-G(|w|)\right]\mathrm{d}t\\ &\leq\liminf\limits_{n\rightarrow \infty}\int_{\mathbb{R}}A(\epsilon t+\epsilon y_{n})\left[\frac{1}{2}g(|w_{n}|)|w_{n}|^{2}-G(|w_{n}|)\right]\mathrm{d}t\\ & = \liminf\limits_{n\rightarrow \infty}\int_{\mathbb{R}}A(\epsilon t)\left[\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\right]\mathrm{d}t\\ & = \liminf\limits_{n\rightarrow \infty}\left[I_{\epsilon}(z_{n})-\frac{1}{2}\langle I'_{\epsilon}(z_{n}), z_{n}\rangle\right] = c_{\epsilon}. \end{aligned} \end{equation}

    Therefore, it follows that

    \begin{equation} \nonumber c_{A_{\infty}}\leq c_{\epsilon}, \; \; \forall \epsilon > 0. \end{equation}

    However, Lemma 4.2 yields that c_{\epsilon} < c_{A_{\infty}} when \epsilon < \epsilon_{0} , which leads to a contradiction. So, we can conclude that \{y_{n}\} is bounded. Then for all n\in \mathbb{N} , there exists r_{0} > 0 such that B_{r}(y_{n})\subset B_{r_{0}}(0) , it holds that

    \begin{equation} \nonumber \int_{B_{r_{0}}(0)}|z_{n}|^{2}\mathrm{d}t\geq \int_{B_{r}(y_{n})}|z_{n}|^{2}\mathrm{d}t\geq \delta. \end{equation}

    Therefore, we obtain that z_{n}\rightharpoonup z in E with z\neq 0 . By repeating the steps in (4.3) and (4.4), we know that z\in\mathscr{N}_{\epsilon} is a nontrivial solution for system (1.1), thus, c_{\epsilon}\leq I_{\epsilon}(z) .

    On the other hand, according to Fatou's lemma, we infer that

    \begin{equation} \nonumber \begin{aligned} c_{\epsilon}& = \liminf\limits_{n\rightarrow \infty}\left[I_{\epsilon}(z_{n})-\frac{1}{2}\langle I'_{\epsilon}(z_{n}), z_{n}\rangle\right]\\ & = \liminf\limits_{n\rightarrow \infty}\int_{\mathbb{R}}A(\epsilon t)\left[\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\right]\mathrm{d}t\\ &\geq\int_{\mathbb{R}}A(\epsilon t)\left[\frac{1}{2}g(|z|)|z|^{2}-G(|z|)\right]\mathrm{d}t\\ & = I_{\epsilon}(z)-\frac{1}{2}\langle I'_{\epsilon}(z), z\rangle = I_{\epsilon}(z). \end{aligned} \end{equation}

    Thus, c_{\epsilon} = I_{\epsilon}(z) . Evidently, it is easy to see that z is a ground state homoclinic orbit of system (1.1). We complete the proof.

    Let

    \mathcal{K}_{\epsilon}: = \left\{z\in E\backslash\{0\}: I'_{\epsilon}(z) = 0\right\}

    be the set of all nontrivial critical points of I_{\epsilon} . In order to describe some important properties of ground state homoclinic orbits, next, we get the following regularity result by taking advantage of the bootstrap argument (see [37] for the iterative steps), this result can also be found in [24, Lemma 2.3].

    Lemma 4.5. If z\in \mathcal{K}_{\epsilon} with |I_{\epsilon}(z)|\leq C_{1} and \|z\|_{2}\leq C_{2} ; then, z\in W^{1, q}(\mathbb{R}, \mathbb{R}^{2N}) for any q > 2 , and \|z\|_{W^{1, q}}\leq C_{q} , where C_{q} depends only on C_{1}, C_{2} and q .

    Below, we use \mathscr{L} to denote the set of all ground state homoclinic orbits of system (1.1). Let z\in\mathscr{L} , then, I_{\epsilon}(z) = c_{\mu} ; applying a standard argument we can show that \mathscr{L} is bounded in E ; therefore, \|z\|_{2}\leq \widehat{c} for all z\in\mathscr{L} and some \widehat{c} > 0 . Hence, making use of Lemma 4.5, we see that, for each q > 2 , there exists C_{q} such that

    \begin{equation} \|z\|_{W^{1, q}}\leq C_{q}, \; \; \forall z\in\mathscr{L}. \end{equation} (4.5)

    Moreover, combining the Sobolev embedding theorem, we can show that there exists C_{\infty} > 0 such that

    \begin{equation} \|z\|_{\infty}\leq C_{\infty}, \; \; \forall z\in\mathscr{L}. \end{equation} (4.6)

    We now shall prove the concentration behavior of the maximum points of the ground state homoclinic orbit. Let z_{\epsilon} be a ground state homoclinic orbit of system (1.1), which can be obtained by Lemma 4.4. Our aim is to show that if t_{\epsilon} is a maximum point of |z_{\epsilon}| , then,

    \begin{equation} \nonumber \lim\limits_{\epsilon\rightarrow 0}A(\epsilon t_{\epsilon}) = A(0). \end{equation}

    In other words, we must show that if \epsilon_{n}\rightarrow 0 , up to a subsequence, \epsilon_{n}t_{\epsilon_{n}}\rightarrow t_{0} for some t_{0}\in \mathscr{A} , where

    \begin{equation} \nonumber \mathscr{A} = \{t\in \mathbb{R}: A(t) = A(0)\} \end{equation}

    denotes the set of the maximum points of A(t) .

    Let \{\epsilon_{n}\}\subset (0, \epsilon_{0}) with \epsilon_{n}\rightarrow 0 as n\to\infty and z_{\epsilon_{n}}\in\mathscr{L} ; we write z_{n}: = z_{\epsilon_{n}} . Then, we have

    \begin{equation} \nonumber I_{\epsilon_{n}}(z_{n}) = c_{\epsilon_{n}}\; \; \hbox{and}\; \; I'_{\epsilon_{n}}(z_{n}) = 0 \end{equation}

    Evidently, in view of Lemma 4.3, we can easily check that \{z_{n}\} is bounded in E .

    Lemma 4.6. There exist a sequence \{y_{n}\}\subset \mathbb{Z} and two constants r > 0 , \delta > 0 such that

    \begin{equation} \nonumber \int_{B_{r}(y_{n})}|z_{n}|^{2}\mathrm{d}t\geq \delta. \end{equation}

    Proof. Arguing by contradiction, we suppose that for any r_{1} > 0 ,

    \begin{equation} \nonumber \lim\limits_{n\rightarrow \infty}\sup \limits_{y\in\mathbb{R}} \int_{B_{r_{1}}(y)}|z_{n}|^{2}\mathrm{d}t = 0. \end{equation}

    Then, according to Lions' concentration-compactness principle, we conclude that z_{n}\rightarrow 0 in L^{q}(\mathbb{R}) for all q > 2 . Furthermore, by (2.1) and (2.2), we obtain

    \begin{equation} \nonumber \int_{\mathbb{R}}A(\epsilon_{n}t)\left[\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\right]\mathrm{d}t\rightarrow 0. \end{equation}

    Therefore, it follows that

    \begin{equation} \nonumber c_{\epsilon_{n}} = I_{\epsilon_{n}}(z_{n})-\frac{1}{2}\langle I'_{\epsilon_{n}}(z_{n}), z_{n}\rangle = \int_{\mathbb{R}}A(\epsilon_{n}t)\left[\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\right]\mathrm{d}t\rightarrow 0. \end{equation}

    Evidently, this is impossible because c_{\epsilon_{n}} > 0 (see Lemma 3.3). We complete the proof.

    Lemma 4.7. The sequence \{\epsilon_{n}y_{n}\} is bounded, and \lim\limits_{n\rightarrow \infty}\epsilon_{n}y_{n} = x_{0}\in \mathscr{A} .

    Proof. Setting v_{n}(t): = z_{n}(t+y_{n}) , up to a subsequence, it is easy to see that v_{n}\rightharpoonup v in E with v\neq 0 from Lemma 4.6. In what follows, we want to prove that the sequence \{\epsilon_{n}y_{n}\} is bounded. If this is not true, we can suppose that there is a subsequence \{\epsilon_{n}y_{n}\} such that |\epsilon_{n}y_{n}|\rightarrow +\infty as n\rightarrow +\infty . Since z_{n} is the ground state homoclinic orbit of system (1.1), v_{n} solves the following system

    \begin{equation} Sv_{n} = A(\epsilon_{n}t+\epsilon_{n}y_{n})g(|v_{n}|)v_{n}, \end{equation} (4.7)

    and the energy

    \begin{equation} \nonumber \begin{aligned} \widehat{I}_{\epsilon_{n}}(v_{n})& = \frac{1}{2}\left(\|v_{n}^{+}\|^{2}-\|v_{n}^{-}\|^{2}\right) -\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})G(|v_{n}|)\mathrm{d}t\\ & = \frac{1}{2}\left(\|z_{n}^{+}\|^{2}-\|z_{n}^{-}\|^{2}\right) -\int_{\mathbb{R}}A(\epsilon_{n}t)G(|z_{n}|)\mathrm{d}t\\ & = \int_{\mathbb{R}}A(\epsilon_{n}t)\left[\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\right]\mathrm{d}t\\ & = I_{\epsilon_{n}}(z_{n}) = c_{\epsilon_{n}}. \end{aligned} \end{equation}

    Furthermore, for every \phi\in E , we have

    \begin{equation} \nonumber (v_{n}^{+}, \phi^{+})-(v_{n}^{-}, \phi^{-})-\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})g(|v_{n}|)v_{n}\phi \mathrm{d}t = 0. \end{equation}

    Since A(\epsilon_{n}t+\epsilon_{n}y_{n})\to A_{\infty} , given that v_n\rightharpoonup v and \phi\in C^{\infty}_{0}(\mathbb{R}) , we get

    \begin{equation} \nonumber (v^{+}, \phi^{+})-(v^{-}, \phi^{-})-\int_{\mathbb{R}}A_{\infty}g(|v|)v\phi \mathrm{d}t = 0. \end{equation}

    Thereby, v is a nontrivial homoclinic orbit of system (3.1) with \mu = A_{\infty} and v\in \mathscr{N}_{A_{\infty}} . In view of Lemma 4.1 and Fatou's lemma, we can conclude that

    \begin{equation} \begin{aligned} c_{A_{\infty}}&\leq I_{A_{\infty}}(v) = I_{A_{\infty}}(v)-\frac{1}{2}\langle I'_{A_{\infty}}(v), v\rangle\\ & = A_{\infty}\int_{\mathbb{R}}\left[\frac{1}{2}g(|v|)|v|^{2}-G(|v|)\right]\mathrm{d}t\\ &\leq \liminf\limits_{n\rightarrow \infty}\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})\left[\frac{1}{2}g(|v_{n}|)|v_{n}|^{2}-G(|v_{n}|)\right]\mathrm{d}t\\ & = \liminf\limits_{n\rightarrow \infty}\int_{\mathbb{R}}A(\epsilon t)\left[\frac{1}{2}g(|z_{n}|)|z_{n}|^{2}-G(|z_{n}|)\right]\mathrm{d}t\\ & = \liminf\limits_{n\rightarrow \infty}\left[I_{\epsilon_{n}}(z_{n})-\frac{1}{2}\langle I'_{\epsilon_{n}}(z_{n}), z_{n}\rangle\right]\\ & = \liminf\limits_{n\rightarrow \infty}I_{\epsilon_{n}}(z_{n}) = \lim\limits_{n\rightarrow \infty}c_{\epsilon_{n}} = c_{0}. \end{aligned} \end{equation} (4.8)

    However, according to Lemma 4.2 we know that c_{0} < c_{A_{\infty}} . Evidently, this is a contradiction. Therefore, \{\epsilon_{n}y_{n}\} is bounded in \mathbb{R} , and passing to a subsequence, we can assume that \epsilon_{n}y_{n}\rightarrow x_{0} . According to the above argument, for \forall \psi\in E , we get

    \begin{equation} \nonumber (v^{+}, \psi^{+})-(v^{-}, \psi^{-})-\int_{\mathbb{R}}A(x_{0})g(|v|)v\psi \mathrm{d}t = 0, \end{equation}

    Obviously, we can see that v is a ground state homoclinic orbit of the following system

    \begin{equation} Sv = A(x_{0})g(|v|)v, \; \; t\in \mathbb{R}, \end{equation} (4.9)

    and v\in \mathscr{N}_{A(x_{0})} . Following to the proof of (4.8), we can get tht c_{A(x_{0})}\leq c_{0} , then, using Lemma 3.9, it follows that A(x_0)\geq A(0) ; together with condition (A) , we can obtain that A(x_0) = A(0) . Hence, we show that \lim\limits_{n\rightarrow \infty}\epsilon_{n}y_{n} = x_{0} and x_{0}\in \mathscr{A} . The proof is completed.

    According to Lemma 4.7, we see that v is a ground state homoclinic orbit of system (4.7), then, I_{0}(v) = c_{0} and I'_{0}(v) = 0 . Using Lemma 4.1 and Fatou's lemma, we directly obtain

    \begin{equation} \nonumber \begin{aligned} c_{0}&\leq \int_{\mathbb{R}}A(0)\left[\frac{1}{2}g(|v|)|v|^{2}-G(|v|)\right]\mathrm{d}t\\ &\leq\liminf\limits_{n\to\infty}\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})\left[\frac{1}{2}g(|v_{n}|)|v_{n}|^{2}-G(|v_{n}|)\right]\mathrm{d}t\\ & = \liminf\limits_{n\to\infty}\widehat{I}_{\epsilon_{n}}(v_{n})\leq\limsup\limits_{n\to\infty}I_{\epsilon_{n}}(z_{n})\leq c_{0}. \end{aligned} \end{equation}

    Hence, we have

    \begin{equation} \lim\limits_{n\to\infty}\widehat{I}_{\epsilon_{n}}(v_{n}) = \lim\limits_{n\to\infty}c_{\epsilon_{n}} = c_{0} = I_{0}(v). \end{equation} (4.10)

    Lemma 4.8. We have the convergence conclusion: v_{n}\to v in E as n\to\infty .

    Proof. Let \eta:[0, +\infty)\rightarrow [0, 1] be a smooth function satisfying that \eta(s) = 1 if s\leq 1 , and \eta(s) = 0 if s\geq 2 . Define \tilde{v}_{n}(t) = \eta(2|t|/n)v(t) , then, for q\in [2, +\infty) , one has

    \begin{equation} \|v-\tilde{v}_{n}\|\rightarrow 0\; \hbox{and}\; \|v-\tilde{v}_{n}\|_{q}\rightarrow 0\; \hbox{as}\; n\rightarrow \infty. \end{equation} (4.11)

    Setting \theta_{n} = v_{n}-\tilde{v}_{n} , it is not difficult to verify that along a subsequence

    \begin{equation} \lim\limits_{n\rightarrow \infty}\left|\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})\left[G(|v_{n}|) -G(|\theta_{n}|)-G(|\tilde{v}_{n}|)\right]\mathrm{d}t\right| = 0 \end{equation} (4.12)

    and

    \begin{equation} \lim\limits_{n\rightarrow \infty}\left|\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n}) \left[g(|v_{n}|)v_{n}-g(|\theta_{n}|)\theta_{n}-g(|\tilde{v}_{n}|)\tilde{v}_{n}\right]\varphi \mathrm{d}t\right| = 0 \end{equation} (4.13)

    uniformly in \varphi\in E with \|\varphi\|\leq 1 . Using the fact that A(\epsilon_{n}t+\epsilon_{n}y_{n})\rightarrow A_{0} as n\rightarrow \infty uniformly on any bounded set of t , and combining the decay of v and (4.11) we can easily check the following result

    \begin{equation} \int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})G(|\tilde{v}_{n}|)\mathrm{d}t\rightarrow \int_{\mathbb{R}}A_{0}G(|v|)\mathrm{d}t. \end{equation} (4.14)

    Consequently, using (4.10), (4.11), (4.12) and (4.14) we infer that

    \begin{equation} \nonumber \begin{aligned} \widehat{I}_{\epsilon_{n}}(\theta_{n})& = \widehat{I}_{\epsilon_{n}}(v_{n})-I_{0}(v)\\ &+\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})\left[G(|v_{n}|)-G(|\theta_{n}|)-G(|\tilde{v}_{n}|)\right]\mathrm{d}t+o_{n}(1)\\ & = o_{n}(1)\; \hbox{as}\; n\rightarrow \infty, \end{aligned} \end{equation}

    which implies that \widehat{I}_{\epsilon_{n}}(\theta_{n})\rightarrow 0 . Similarly, we also obtain

    \begin{equation} \nonumber \begin{aligned} \langle\widehat{I}'_{\epsilon_{n}}(\theta_{n}), \varphi\rangle & = \int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})\left[g(|v_{n}|)v_{n}-g(|\theta_{n}|)\theta_{n} -g(|\tilde{v}_{n}|)\tilde{v}_{n}\right]\varphi\mathrm{d}t+o_{n}(1)\\ & = o_{n}(1)\; \hbox{uniformly in}\; \|\varphi\|\leq 1\; \hbox{as}\; n\rightarrow \infty, \end{aligned} \end{equation}

    which implies that \widehat{I}'_{\epsilon_{n}}(\theta_{n})\rightarrow 0 . Therefore

    \begin{equation} \nonumber o_{n}(1) = \widehat{I}'_{\epsilon_{n}}(\theta_{n})-\frac{1}{2}\langle\widehat{I}'_{\epsilon_{n}}(\theta_{n}), \theta_{n}\rangle = \int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})\left[\frac{1}{2}g(|\theta_{n}|)|\theta_{n}|^{2}-G(|\theta_{n}|)\right]\mathrm{d}t, \end{equation}

    from which together with (g_{3}) , we can infer that

    \begin{equation} \nonumber \int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})g(|\theta_{n}|)|\theta_{n}|^{2}\mathrm{d}t\rightarrow 0. \end{equation}

    Notice that \{\|v_{n}\|_{\infty}\} is bounded, thus,

    \begin{equation} \nonumber \int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})g(|\theta_{n}|)|\theta_{n}^{+}-\theta_{n}^{-}|^{2}\mathrm{d}t\leq C. \end{equation}

    As a consequence, we obtain

    \begin{equation} \nonumber \begin{aligned} \|\theta_{n}\|^{2}& = \langle\widehat{I}'_{\epsilon_{n}}(\theta_{n}), \theta_{n}^{+}-\theta_{n}^{-}\rangle +\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})g(|\theta_{n}|)\theta_{n}(\theta_{n}^{+}-\theta_{n}^{-})\mathrm{d}t\\ = &o_{n}(1)+\int_{\mathbb{R}}A^{\frac{1}{2}}(\epsilon_{n}t+\epsilon_{n}y_{n})g^{\frac{1}{2}}(|\theta_{n}|)|\theta_{n}| A^{\frac{1}{2}}(\epsilon_{n}t+\epsilon_{n}y_{n})g^{\frac{1}{2}}(|\theta_{n}|)|\theta_{n}^{+}-\theta_{n}^{-}|\mathrm{d}t\\ \leq& o_{n}(1)+\left(\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})g(|\theta_{n}|)|\theta_{n}|^{2}\mathrm{d}t\right)^{\frac{1}{2}}\\ &\left(\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})g(|\theta_{n}|)|\theta_{n}^{+}-\theta_{n}^{-}|^{2}\mathrm{d}t\right)^{\frac{1}{2}}\\ \leq& o_{n}(1)+C\left(\int_{\mathbb{R}}A(\epsilon_{n}t+\epsilon_{n}y_{n})g(|\theta_{n}|)|\theta_{n}|^{2}\mathrm{d}t\right)^{\frac{1}{2}}\\ = &o_{n}(1), \end{aligned} \end{equation}

    that is, \|\theta_{n}\|\rightarrow 0 , which together with (4.11) leads to v_{n}\rightarrow v in E as n\rightarrow \infty .

    Lemma 4.9. We have that v_{n}(t)\to0 uniformly in n\in \mathbb{N} as t\to\infty . Moreover, there exist c, C > 0 such that for all t\in\mathbb{R} , it holds that

    \begin{equation} \nonumber |v_{n}(t)|\leq C \exp(-c|t|). \end{equation}

    Proof. Firstly, we observe that if z is a homoclinic orbit of system (1.1), then it satisfies the following relation

    \begin{equation} \nonumber \frac{\mathrm{d}}{\mathrm{d}t}z = \mathscr{J}\bigg(Lz+A(\epsilon t)g(|z|)z\bigg). \end{equation}

    Computing directly, we obtain

    \begin{equation} \nonumber \frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}z = (\mathscr{J}L)^{2}z+Q(t, z) \end{equation}

    with

    \begin{equation} \begin{aligned} Q(t, z) = &\mathscr{J}\bigg[\bigg(\epsilon A'(\epsilon t)+L\mathscr{J}A(\epsilon t)\bigg)g(|z|)z+\bigg(g'_{z}(|z|)|z|+g(|z|)\bigg)A(\epsilon t)\mathscr{J}Lz\\ & \ +\bigg(g'_{z}(|z|)|z|+g(|z|)\bigg)A^{2}(\epsilon t)Lg(|z|)z\bigg]. \end{aligned} \end{equation} (4.15)

    Setting

    \begin{eqnarray*} \hbox{sgn}z \left\{ \begin{array}{ll} \frac{z}{|z|}&\hbox{if}\; z\neq 0, \\ 0, &\hbox{if}\; \; z = 0.\\ \end{array} \right. \end{eqnarray*}

    Applying Kato's inequality and (4.15), and using the real positivity of (\mathscr{J}L)^{2} , we can find some \rho > 0 such that

    \begin{equation} \frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}|z|\geq \frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}z (\hbox{sgn}z) = (\mathscr{J}L)^{2}z\frac{z}{|z|}+Q(t, z)\frac{z}{|z|}\geq \rho|z|-|Q(t, z)|. \end{equation} (4.16)

    Hence, using (2.1), (4.6), (4.15) and (4.16) we conclude that there exists \kappa > 0 such that

    \begin{equation} \nonumber \frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}|z|\geq -\kappa|z|\; \; \; \hbox{for all}\; \; t\in \mathbb{R}. \end{equation}

    Then by the sub-solution estimate [38], there is a \widehat{c}_{0} independent of t ; we have the following estimate

    \begin{equation} |z(t)|\leq \widehat{c}_{0}\int_{B_{1}(t)}|z(s)|\mathrm{d}s. \end{equation} (4.17)

    Now we claim that v_{n}(t)\to0 uniformly in n\in \mathbb{N} as t\to\infty . Indeed, if it is not true, then using (4.17) we can find that there exist c_{0} > 0 and t_{n}\in \mathbb{R} with |t_{n}|\rightarrow \infty such that

    \begin{equation} \nonumber c_{0}\leq |v_{n}(t_{n})|\leq \widehat{c}_{0}\int_{B_{1}(t_{n})}|v_{n}(t)|\mathrm{d}t, \end{equation}

    this is because v_n satisfies \widehat{I}'_{\epsilon_{n}}(v_{n}) = 0 , then, the above processes still hold for v_{n} . From Lemma 4.8, it follows that v_n\to v in E . Therefore, we get

    \begin{equation} \nonumber \begin{aligned} c_{0}&\leq |v_{n}(t_{n})|\leq \widehat{c}_{0}\int_{B_{1}(t_{n})}|v_{n}(t)|\mathrm{d}t\leq \widehat{c}_{0}\int_{B_{1}(t_{n})}|v_{n}-v|\mathrm{d}t+\widehat{c}_{0}\int_{B_{1}(t_{n})}|v|\mathrm{d}t\\ &\leq \overline{c}\bigg(\int_{\mathbb{R}}|v_{n}-v|^{2}\mathrm{d}t\bigg)^{\frac{1}{2}}+\widehat{c}_{0}\int_{B_{1}(t_{n})}|v|\mathrm{d}t\rightarrow 0, \end{aligned} \end{equation}

    which yields a contradiction. So, the claim holds.

    Note that g(s) = o(1) and g'_{s}(s)s = o(1) as s\rightarrow 0 ; then, we can find suitable constants 0 < \delta < \frac{\rho}{2} and R > 0 such that

    \begin{equation} \nonumber |Q(t, v_{n})|\leq \frac{\rho}{2}|v_{n}|, \; \; \; \forall|t|\geq R. \end{equation}

    Combining the above relation and (4.16), we get

    \begin{equation} \nonumber \frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}|v_{n}|\geq \delta |v_{n}|, \; \; \; \forall|t|\geq R. \end{equation}

    Let \Lambda(t) be a fundamental solution of the following equation

    \begin{equation} \nonumber -\frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}\Lambda+\delta \Lambda = 0. \end{equation}

    From the uniform boundedness, we may choose \Lambda(t) such that |v_{n}(t)|\leq\delta\Lambda(t) holds on |t| = R for all n\in\mathbb{N} . Let u_{n} = |v_{n}|-\delta\Lambda ; thus, we obtain

    \begin{equation} \nonumber \frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}u_{n} = \frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}|v_{n}|-\delta \frac{\mathrm{d}^{2}}{\mathrm{d}t^{2}}\Lambda\geq\delta(|v_{n}|-\delta\Lambda) = \delta u_{n}, \; \; \; \hbox{for all}\; \; |t|\geq R. \end{equation}

    The maximum principle yields that u_{n}(t)\leq 0 for |t|\geq R , i.e., |v_{n}(t)|\leq \delta \Lambda(t) for |t|\geq R . As we know that there exists c_{1} > 0 such that

    \begin{equation} \nonumber \Lambda(t)\leq c_{1}\exp(-\sqrt{\delta}|t|)\; \; \; \; \; \hbox{for all}\; \; \; |t|\geq 1. \end{equation}

    Therefore, there are constants C, c > 0 ; we obtain

    \begin{equation} \nonumber |v_{n}(t)|\leq C\exp(-c|t|)\; \; \; \hbox{for all}\; \; t\in\mathbb{R}. \end{equation}

    We complete the proof.

    Lemma 4.10. There exists \nu > 0 such that \|v_{n}\|_{\infty}\geq \nu for all n\in\mathbb{N} .

    Proof. According to Lemma 4.6, we can see that there exist r > 0 and \delta > 0 such that

    \begin{equation} \nonumber \int_{B_{r}(0)}|v_{n}|^{2}\mathrm{d}t\geq\delta. \end{equation}

    Suppose by contradiction that \|v_{n}\|_{\infty}\to0 as n\to\infty , then, it holds that

    \begin{equation} \nonumber 0 < \delta\leq\int_{B_{r}(0)}|v_{n}|^{2}\mathrm{d}t\leq |B_{r}|\|v_{n}\|_{\infty}^{2}\to0\; \hbox{as}\; n\to\infty, \end{equation}

    which is absurd. This ends the proof.

    Finally, based on the above facts, next we give the completed proof of Theorem 1.1.

    Proof of Theorem 1.1 (completed). Suppose that q_{n} is a global maximum point of |v_{n}(t)| for each n\in \mathbb{N} , then,

    \begin{equation} \nonumber |v_{n}(q_{n})| = \max\limits_{t\in \mathbb{R}}|v_{n}(t)|. \end{equation}

    Since v_{n}(t) = z_{n}(t+y_{n}) , we can see that p_{n} = q_{n}+y_{n} is a maximum point of |z_{n}(t)| . Lemma 4.10 shows that there exists \nu > 0 such that

    \begin{equation} \nonumber |v_{n}(q_{n})|\geq \nu \; \; \hbox{for all}\; \; n\in \mathbb{N}, \end{equation}

    then we know that \{q_{n}\} is bounded. So, we conclude from Lemma 4.7 that

    \begin{equation} \nonumber \epsilon_{n}p_{n} = \epsilon_{n}q_{n}+\epsilon_{n}y_{n}\rightarrow x_{0}\in \mathscr{A}. \end{equation}

    Consequently, it follows that

    \begin{equation} \nonumber \lim\limits_{n\rightarrow \infty}A(\epsilon_{n}p_{n}) = A(x_{0}) = A(0). \end{equation}

    Furthermore, from Lemma 4.7 and Lemma 4.8, it is easy to see that z_{n}(t+p_n) converges to a ground state homoclinic orbit v of the following limit system

    \begin{equation} \nonumber Sz = A(0)g(|z|)z, \; \; t\in \mathbb{R}. \end{equation}

    From Lemma 4.9 and the boundedness of \{q_{n}\} , we derive that

    \begin{equation} \nonumber \begin{aligned} |z_{n}(t)| = &|v_{n}(t-y_n)|\leq C\exp\left(-c|t-y_{n}|\right) = C\exp\left(-c|t-p_{n}+q_{n}|\right)\\ \leq & C\exp\left(-c|t-p_{n}|+c|q_{n}|\right)\leq \widetilde{C}\exp\left(-\tilde{c}|t-p_{n}|\right) \end{aligned} \end{equation}

    for some \tilde{c}, \widetilde{C} > 0 and all t\in\mathbb{R} .

    Finally, we observe that Lemma 4.2 shows that, there is \epsilon_{0} > 0 ; system (1.1) has a ground state homoclinic orbit z_{\epsilon} for each \epsilon \in(0, \epsilon_{0}) . So, the conclusion (a) of Theorem 1.1 holds. Moreover, according to the above discussions, we directly obtain the following conclusions:

    (b) let t_{\epsilon} be the maximum point of |z_{\epsilon}(t)| , then,

    \lim\limits_{\epsilon\to0}A(\epsilon t_{\epsilon}) = A(0);

    and z_{\epsilon}(t+t_{\epsilon})\to v in E , where v is a ground state homoclinic orbit of the limit system

    \begin{equation} \nonumber Sz = A(0)g(|z|)z, \; \; t\in \mathbb{R}; \end{equation}

    (c) there are two positive constants \tilde{c} , \widetilde{C} such that

    \begin{equation} \nonumber |z_{\epsilon}(t)|\leq \widetilde{C}\exp\left(-\tilde{c}|t-t_{\epsilon}|\right). \end{equation}

    We have finished the proof of all conclusions of Theorem 1.1.

    The authors declare that no artificial intelligence tools were used in the creation of this article.

    The research of Tianfang Wang was supported by the High Level Research Achievement Project-General Project of Natural Science of Baotou Teachers' College (BSYKJ2022-ZY09), the Youth Innovative Talent Project of Baotou City (2022). The research of Wen Zhang was supported by the Natural Science Foundation of Hunan Province (2022JJ30200), the Key project of Scientific Research Project of Department of Education of Hunan Province (22A0461), and Aid Program for Science and Technology Innovative Research Team in Higher Educational Institutions of Hunan Province.

    The authors declare that they have no competing interests.



    [1] D. Silver, A. Huang, C. J. Maddison, A. Guez, L. Sifre, G. Van Den Driessche, et al., Mastering the game of go with deep neural networks and tree search, Nature, 529 (2016), 484–489. https://doi.org/10.1038/nature16961 doi: 10.1038/nature16961
    [2] M. Moravcík, M. Schmid, N. Burch, V. Lisý, D. Morrill, N. Bard, et al., Deepstack: Expert-level artificial intelligence in no-limit poker, CoRR, abs/1701.01724.
    [3] M. V. Devarakonda, C. Tsou, Automated problem list generation from electronic medical records in IBM watson, in Proc. Twenty-Ninth AAAI Conf. Artif. Intell., (eds. B. Bonet, S. Koenig), (2015), 3942–3947.
    [4] D. Leech, Designing the voting system for the council of the european union, Public Choice, 113 (2002), 473–464. https://doi.org/10.1023/A:1020877015060 doi: 10.1023/A:1020877015060
    [5] O. Shehory, S. Kraus, Methods for task allocation via agent coalition formation, Artif. Intell., 101 (1998), 165–200. https://doi.org/10.1016/S0004-3702(98)00045-9 doi: 10.1016/S0004-3702(98)00045-9
    [6] Y. Zick, K. Gal, Y. Bachrach, M. Mash, How to form winning coalitions in mixed human-computer settings, in Proc. Twenty-Sixth Int. Joint Conf. Artif. Intell., IJCAI (ed. C. Sierra), (2017), 465–471. https://doi.org/10.24963/ijcai.2017/66
    [7] L. S. Shapley, M. Shubik, A method for evaluating the distribution of power in a committee system, Am. political Sci. Rev., 48 (1954), 787–792. https://doi.org/10.2307/1951053 doi: 10.2307/1951053
    [8] J. F. Banzhaf III, Weighted voting doesn't work: A mathematical analysis, Rutgers L. Rev., 19 (1964), 317.
    [9] L. Wu, S. Chen, X. Gao, Y. Zheng, J. Hao, Detecting and learning against unknown opponents for automated negotiations, in PRICAI 2021: Trends in Artificial Intelligence (eds. D. N. Pham, T. Theeramunkong, G. Governatori and F. Liu), (2021), 17–31. https://doi.org/10.1007/978-3-030-89370-5_2
    [10] X. Gao, S. Chen, Y. Zheng, J. Hao, A deep reinforcement learning-based agent for negotiation with multiple communication channels, in 2021 IEEE 33nd Int. Conf. Tools with Artif. Intell. (ICTAI), (2021), 868–872. https://doi.org/10.1109/ICTAI52525.2021.00139
    [11] C. Gao, J. Liu, Network-based modeling for characterizing human collective behaviors during extreme events, IEEE Trans. Syst. Man Cybern. Syst., 47 (2017), 171–183. https://doi.org/10.1109/TSMC.2016.2608658
    [12] H. Mao, Z. Zhang, Z. Xiao, Z. Gong, Y. Ni, Learning multi-agent communication with double attentional deep reinforcement learning, Auton. Agents Multi Agent Syst., 34 (2020), 32. https://doi.org/10.1007/s10458-020-09455-w doi: 10.1007/s10458-020-09455-w
    [13] J. N. Foerster, Y. M. Assael, N. De Freitas, S. Whiteson, Learning to communicate with deep multi-agent reinforcement learning, arXiv prints, arXiv: 1605.06676.
    [14] P. Peng, Y. Wen, Y. Yang, Q. Yuan, Z. Tang, H. Long, et al., Multiagent bidirectionally-coordinated nets: Emergence of human-level coordination in learning to play starcraft combat games, arXiv prints, arXiv: 1703.10069.
    [15] T. Eccles, Y. Bachrach, G. Lever, A. Lazaridou, T. Graepel, Biases for emergent communication in multi-agent reinforcement learning, in Advances in Neural Information Processing Systems 32: Annual Conference on Neural Information Processing Systems 2019, NeurIPS 2019, Vancouver, BC, Canada (eds. H. M. Wallach, H. Larochelle, A. Beygelzimer, F. d'Alché-Buc, E. B. Fox, R. Garnett), (2019), 13111–13121. https://dl.acm.org/doi/10.5555/3454287.3455463
    [16] E. Hughes, T. W. Anthony, T. Eccles, J. Z. Leibo, D. Balduzzi, Y. Bachrach, Learning to resolve alliance dilemmas in many-player zero-sum games, in Proceedings of the 19th International Conference on Autonomous Agents and Multiagent Systems, AAMAS '20, (eds. A. E. F. Seghrouchni, G. Sukthankar, B. An, N. Yorke{-}Smith), (2020), 538–547.
    [17] T. Matthews, S. Ramchurn, G. Chalkiadakis, Competing with humans at fantasy football: Team formation in large partially-observable domains, in Proc. AAAI Conf. Artif. Intell., 26 (2012). https://ojs.aaai.org/index.php/AAAI/article/view/8259
    [18] Y. Bachrach, R. Everett, E. Hughes, A. Lazaridou, J. Z. Leibo, M. Lanctot, et al., Negotiating team formation using deep reinforcement learning, Artif. Intell., 288 (2020), 103356. https://doi.org/10.1016/j.artint.2020.103356 doi: 10.1016/j.artint.2020.103356
    [19] J. W. Crandall, M. Oudah, F. Ishowo-Oloko, S. Abdallah, J. F. Bonnefon, M. Cebrian, et al., Cooperating with machines, Nat. Commun., 9 (2018), 1–12. https://doi.org/10.1038/s41467-017-02597-8
    [20] K. Cao, A. Lazaridou, M. Lanctot, J. Z. Leibo, K. Tuyls, S. Clark, Emergent communication through negotiation, in 6th Int. Conf. Learn. Represent., ICLR 2018, Vancouver, Conference Track Proceedings, 2018.
    [21] Y. Shoham, K. Leyton-Brown, Multiagent systems: Algorithmic, game-theoretic, and logical foundations, Cambridge University Press, 2008. https://dl.acm.org/doi/abs/10.1145/1753171.1753181
    [22] D. K. Kim, S. Omidshafiei, J. Pazis, J. P. How, Crossmodal attentive skill learner: learning in atari and beyond with audio–-video inputs, Auton. Agent. Multi Agent Syst., 34 (2020), 1–21. https://doi.org/10.1007/s10458-019-09439-5 doi: 10.1007/s10458-019-09439-5
    [23] R. Su, Y. Zhu, Q. Zou, L. Wei, Distant metastasis identification based on optimized graph representation of gene interaction patterns, Brief. Bioinform., 23, (2022), bbab468. http://doi.org/10.1093/bib/bbab468
    [24] J. Liu, R. Su, J. Zhang, L. Wei, Classification and gene selection of triple-negative breast cancer subtype embedding gene connectivity matrix in deep neural network, Brief. Bioinform., 22, (2021), bbaa395. https://doi.org/10.1093/bib/bbaa395
    [25] A. Rubinstein, Perfect equilibrium in a bargaining model, Econometrica, 50 (1982), 97–109.
    [26] S. Chen, G. Weiss, An intelligent agent for bilateral negotiation with unknown opponents in continuous-time domains, ACM Trans. Auton. Adapt. Syst., 9 (2014), 16: 1–16: 24. https://dl.acm.org/doi/10.1145/2629577
    [27] S. Chen, G. Weiss, An approach to complex agent-based negotiations via effectively modeling unknown opponents, Expert Syst. Appl., 42 (2015), 2287–2304. https://doi.org/10.1016/j.eswa.2014.10.048 doi: 10.1016/j.eswa.2014.10.048
    [28] K. Dautenhahn, Socially intelligent robots: dimensions of human–robot interaction, Philos. Trans. R. Soc. B Biol. Sci., 362 (2007), 679–704. https://doi.org/10.1098/rstb.2006.2004 doi: 10.1098/rstb.2006.2004
    [29] S. Chen, Y. Cui, C. Shang, J. Hao, G. Weiss, Onecg: Online negotiation environment for coalitional games, in Proc. 18th Int. Conf. Auton. Agent. MultiAg. Syst., (2019), 2348–2350. https://dl.acm.org/doi/10.5555/3306127.3332108
    [30] G. A. Rummery, M. Niranjan, On-line Q-learning using connectionist systems, University of Cambridge, Department of Engineering Cambridge, UK, 1994.
    [31] R. Su, X. Liu, L. Wei, Q. Zou, {Deep-Resp-Forest}: A deep forest model to predict anti-cancer drug response, Methods, 166 (2019) 91–102. https://doi.org/10.1016/j.ymeth.2019.02.009
    [32] R. Su, X. Liu, Q. Jin, X. Liu, L. Wei, Identification of glioblastoma molecular subtype and prognosis based on deep mri features, Knowl. Based Syst., 232 (2021), 107490. https://doi.org/10.1016/j.knosys.2021.107490 doi: 10.1016/j.knosys.2021.107490
    [33] R. Su, H. Wu, B. Xu, X. Liu, L. Wei, Developing a multi-dose computational model for drug-induced hepatotoxicity prediction based on toxicogenomics data, IEEE/ACM Trans. Comput. Biol. Bioinform., 16 (2018), 1231–1239. https://doi.org/10.1109/TCBB.2018.2858756 doi: 10.1109/TCBB.2018.2858756
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2717) PDF downloads(102) Cited by(9)

Figures and Tables

Figures(5)  /  Tables(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog