diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzelsx" "b/data_all_eng_slimpj/shuffled/split2/finalzzelsx" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzelsx" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction}\n\nThe goal of this paper is the computation of correlation functions in the six-vertex model directly by Markov chain Monte-Carlo simulations. This model was introduced by Pauling~\\cite{Pauling} who proposed it to describe the crystal where the oxygen groups form a square lattice with a hydrogen atom between each pair of lattice sites.\nHe proposed the ice rule: each lattice site has two hydrogen atoms close to it and two further apart; see a recent historical review~\\cite{Harris}.\nAnother crystal with such a structure is the potassium dihydrogen phosphate.\nSlater was the first who suggested that the two dimensional case, known as the six-vertex model, is important \nto understand universal thermodynamic properties of these structures~\\cite{Slater}. \nThe states in this model are configurations of arrows on edges which satisfy the ice rules, see Fig.~\\ref{6v}.\nAn arrow indicates to which of two sites (the oxygen atoms) the hydrogen atom (which is approximately in the middle of an edge) is closer.\nEquivalently, the configurations of arrows can be regarded as configurations of lattice paths such that paths may meet at a vertex, \nturn or pass, as it is shown in Fig.~\\ref{6v}. Boltzmann weight of a configuration is the product of Boltzmann weights assigned to vertices. The weight of a vertex depends on the configurations of paths on adjacent edges,\nsee Fig. \\ref{6v}. The probability of state $\\sigma$ is\n\\[\nProb(\\sigma)=\\frac{1}{Z} \\, w(\\sigma),\n\\]\nwhere $w(\\sigma)=\\prod_v w_v(\\sigma)$ is the weight of state $\\sigma$, $w_v(\\sigma)$ is the weight of the vertex $v$\nin the state $\\sigma$, and $Z=\\sum_\\sigma w(\\sigma)$ is the partition function. \n\nLocally, lattice paths of the six-vertex model on a planar lattice can be regarded as level curves of a step function\ndefined on faces~\\cite{R2010}. We assume it is increasing when we move to the right and up. This integer valued function is called \nthe {\\it height function} $\\chi(n,m)$. It is a random variable with values in integers $\\mathbb Z$ with the probability distribution given by Boltzmann weights described above.\nOn a planar simply connected lattice domain there is a bijection between configurations of paths with fixed positions on the boundary and height functions with corresponding boundary values\\footnote{We assume that the value of a height function is fixed at some reference point on the domain.}. \n\n\\begin{figure}[ht!]\n\\begin{center}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure01.pdf}\n\\end{center}\n\\caption{\\label{6v} Local configurations and weights of the six-vertex model. For the symmetric model, $\\tau=\\tau_{1}=\\tau_{2}$, where $\\tau=a,b,c$.}\n\\end{figure}\n\nThe first breakthrough in the study of the model came in works of Lieb, Yang, Sutherland and others\nwhere the Bethe ansatz was used for finding the spectrum of transfer matrices with periodic boundary conditions~\\cite{Lieb,Yang,Sutherland}.\nThen, came works of Baxter where the role of commuting transfer matrices became clear and the partition function of the eight-vertex model was obtained~\\cite{BaxterPaper}, see~\\cite{Baxter} for an overview of these developments.\nThen, many important algebraic structures came in the framework of the algebraic Bethe ansatz and the quantum inverse scattering method~\\cite{FT}, for an overview see~\\cite{BIK,BM,AGP}.\nIn the last decade, substantially better understanding of thermodynamic properties of the six-vertex model with domain wall boundary \nconditions (DW) on a square lattice has been achieved.\nThese boundary conditions correspond to paths coming through each edge on the top side of the square and leaving through edges on the right side only.\nThese particular boundary conditions are quite remarkable because the partition function in this case can be written as a determinant~\\cite{Izergin,BIK,PronkoTMF} as well as because of the relation to \nthe alternating sign matrices~\\cite{Ku}.\n\nIn the large volume limit (the thermodynamic limit, $N\\to \\infty$), properly normalized height function converges, as a random variable, \nto a deterministic function $h_{0}(x,y): D=[0,1]\\times [0,1]\\to \\mathbb R$ known as the limit shape\nheight function.\nSuch a behavior is known as the limit shape phenomenon, see~\\cite{CP,Ok} for an overview.\nThis phenomenon is an analogue of the central limit theorem in probability theory.\nIt predicts the following behavior of the height function as $N\\to \\infty$\\footnote{This means the convergence of $\\chi(n,m)\/N\\to h_0\\left(\\frac{n}{N}, \\frac{m}{N}\\right)$ and $\\chi(n,m)-Nh_0\\left(\\frac{n}{N}, \\frac{m}{N}\\right)\\to \\phi\\left(\\frac{n}{N}, \\frac{m}{N}\\right)$ in probability.}:\n\\begin{equation}\n\\label{eq1chi}\n\\chi \\left(n,m \\right) \\to N h_0\\left( \\frac{n}{N}, \\frac{m}{N} \\right)+\\phi \\left(\\frac{n}{N}, \\frac{m}{N} \\right).\n\\end{equation}\nHere $h_0(x,y)$ is the limit shape height function which can be computed using the variational principle.\nThe variational principle was developed and proved for dimer models in~\\cite{KOS,Ok}. It was adopted to the six-vertex model in~\\cite{ZJ,PR}.\nThe random variable $\\phi(x,y)$ is a free Gaussian quantum field in the Euclidean space time with the metric determined by the height function $h(x,y)$, see for example~\\cite{ZJ,GorinLectures}.\nFor generic values of parameters in the six-vertex model, mathematically, the variational principle is still a hypothesis~\\cite{PR,GBDJ}. It is proven in some special cases of stochastic weights in~\\cite{B-stoch} and\nit follows from~\\cite{CKP} for the free fermionic case $\\Delta=0$. \n\nAn important characteristic of the (symmetric) six-vertex model is the parameter~\\cite{BaxterPaper}\n\\begin{equation}\n\\label{delta}\n\\Delta=\\frac{a^2+b^2-c^2}{2ab}.\n\\end{equation}\nWhen $\\Delta=0$, the model can be mapped to a dimer model and the partition function and correlation functions\ncan be computed in terms of the determinant and the minors of the Kasteleyn matrix~\\cite{Ka}, respectively. \n\nIt has already been shown earlier~\\cite{R2005,AS} how to use configurations generated by a Markov chain Monte-Carlo simulations for calculating the limit shape of the height function of the six-vertex model with DW boundary conditions.\nIn this paper, we show how, based on the generated configurations, to compute numerically the two-point correlation function.\nWhen $\\Delta=0$ both the limit shape height function and the correlation functions are known from the exact solution because in this case the six-vertex can be mapped to a dimer model on a modified (decorated) square lattice, details can be found in~\\cite{PR,RS}. We demonstrate that in this case the usual averaging over time in Markov process gives an excellent agreement of numerical results with the exact ones.\n\nAfter that, we apply the same algorithm for other values of $\\Delta$.\nWhen $|\\Delta|\\leq 1$ the model is critical, i.e. the Gaussian field $\\phi(x,y)$ is a massless \nfield on the space time with the metric induced by the limit shape. The numerics confirms that \ncorrelation functions are conformal at short distances. Of course, we should not expect conformal invariance at all distances\nfor $\\Delta$ other than zero. \n\nWhen $\\Delta<-1$ an antiferroelectric diamond shape droplet forms\nin the middle of the limit shape. Because the antiferroelectric ground state is double degenerate,\nthe Markov process gets stuck in one of the ground states for exponentially long time. \nNumerical estimations for this case are given in the last section.\n\nThe paper is organized as follows. In section~\\ref{exact}, we outline a derivation of the\ntwo-point correlation function from the exact solution of the six-vertex model for $\\Delta=0$.\nIn section~\\ref{num}, we demonstrate the results of numerical Monte-Carlo simulations and comparisons with the exact solution.\nIn the appendices, we provide the technical details to specify the model to obtain the exact solution.\n \n{\\bf Acknowledgements.} We would like to thank A.~G.~Pronko for discussions and for sharing a draft of the manuscript~\\cite{Pronko},\nD. Keating and A. Sridhar for numerous discussions and the latest version of the Monte-Carlo code.\nWe also benefited from discussions with A.~A.~Nazarov.\nThe work of N.~Yu.~Reshetikhin was partly supported by the\nNSF grant DMS-1902226 and the RSF grant 18-11-00297.\nP.~A.~Belov is grateful to the Russian Science Foundation, grant no. 18-11-00297, for the financial support.\nThe calculations were carried out using the facilities of the SPbU Resource Center ``Computational Center of SPbU''.\n\n\\section{Correlation functions in the six-vertex model at the free fermionic point}\\label{exact}\n\n\\subsection{The free fermionic point of the six-vertex model and mapping to dimers}\\label{LSh}\nIn this paper, we focus on the symmetric six-vertex model with weights $a_{1}=a_{2}=a$, $b_{1}=b_{2}=b$, $c_{1}=c_{2}=c$.\nThe parameter $\\Delta$, Eq.~(\\ref{delta}), is an important characteristic of the model.\nIt determines the phases of the model on the $M\\times N$ torus when $M,\\, N \\to \\infty$.\nFor simplicity, in the following we assume that $M=N$.\nWhen $\\Delta>1$ the model develops a ferroelectric, totally ordered phase.\nFor $|\\Delta|<1$ it develops a disordered phase and for $\\Delta<-1$ it transitions to an antiferroelectric phase.\nWhen $\\Delta=\\pm 1$ the model undergoes phase transitions (in parameter $\\Delta$).\n\nWhen the weights of the six-vertex model satisfy the condition $\\Delta=0$\nthe six-vertex model can be mapped to the dimer model on a modified lattice, see for example\n\\cite{Wu,RS}.\nThe partition function of the dimer model is the sum of Pfaffians (the number of terms is determined by the topology of the lattice)~\\cite{Ka,FisherPR,McCoyWu}.\nEach Pfaffian can be regarded as the Gaussian Grassmann integral.\nBecause of this and because it implies that the multipoint correlators can be expressed as Pfaffians of the two-point correlation functions, the case of $\\Delta=0$ is also known as a free fermionic point of the six-vertex model.\n\nBecause the weights can be multiplied by an overall constant factor without changing the probability measure, we can put $c=1$.\nThen, we can parametrize weights $a$ and $b$ as\n$$\na=\\cos{(u)}, \\enskip b=\\sin{(u)}.\n$$\nThis is a particular case of Baxter's parametrization of \nweights of the six-vertex model \\cite{Baxter}. \nNote that the mapping $a\\mapsto b, \\ \\ b\\mapsto a$ is a symmetry of the \nprobability measure, see Appendix \\ref{Sym} for details. This is why we can assume,\nwithout loosing generality, that $b\/a\\leq 1$. \n\n\\subsection{The variational principle}\nHere we will recall the variational principle for deriving the limit shape.\nLet $\\sigma(s,t)$ be the free energy per site for the six-vertex model on a torus with $s$ and $t$ being fixed\ndensities of edges occupied vertical and horizontal paths respectively.\n\nThe limit shape height function $h_0(x,y)$ for the six-vertex model with DW boundary conditions is a real valued \nfunction on $\\mathcal D=[0,1]\\times [0,1]$ which minimizes the large deviation rate functional\n\\begin{equation}\\label{LDf}\nS[h]=-{\\int \\int}_{\\mathcal D} \\sigma(\\pa_xh,\\pa_yh) \\, dx \\, dy\n\\end{equation}\nin the space of functions with boundary conditions $h(0,y)=h(x,0)=0, \\ \\ h(1,y)=y, \\ \\ h(x,1)=x$ and the constraints\n\\[\n|h(x,y)-h(x',y)|\\leq |x-x'|, \\ \\ |h(x,y)-h(x,y')|\\leq |y-y'|.\n\\]\nFor dimer models it follows from~\\cite{CKP}.\n\nThe critical value $S[h_{0}]$ is the minus free energy of the model.\nIf $Z_{N}$ is the partition function of the six-vertex model with DW boundary conditions, then\n$$\nS[h_{0}] = \\lim_{N\\to\\infty} \\frac{1}{N^{2}} \\ln Z_{N}.\n$$\n\nThe six-vertex model at the free fermionic point~($\\Delta=0$) can be mapped to the dimer model \non a decorated square lattice, see for example~\\cite{RS} and references therein.\nAs it was already stated earlier, the corresponding dimer model can be solved by the Pfaffian method.\nThis method gives the formula for the partition function of the model as a Pfaffian (or a determinant) of certain $N\\times N$\nmatrix, called the Kasteleyn matrix~\\cite{Ka}.\nIn this case $\\sigma(s,t)$ can be computed explicitly as the Legendre transform of the free energy $f(H,V)$\nof the six-vertex model on a torus (with $\\Delta=0$) in the presence of electric fields $H$ and $V$:\n$$\n\\sigma(s,t)=\\min_{H,V} \\left( Hs + Vt - f(H,V) \\right).\n$$\nThe function $f(H,V)$ is given by the double integral\n\\[\nf(H,V)=\\frac{1}{(2\\pi i)^2}\\int_{|z|=\\exp(H)}\\int_{|w|=\\exp(V)} \\ln|P(z,w)| \\frac{dz}{z}\\frac{dw}{w},\n\\]\nwhere\n\\begin{equation}\n\\label{SpectralCurve}\nP(z,w)=a(wz-1)+b(z+w)\n\\end{equation}\nis the spectral polynomial of the Kasteleyn matrix, see \\cite{Ka,McCoyWu,KOS}. \nNote that $f(H,V)$ is convex, $\\text{det}(\\partial_{i}\\partial_{j} f)>0$, and $\\sigma(s,t)$ is concave, $\\text{det}(\\partial_{i}\\partial_{j} \\sigma)<0$.\n\nEuler-Lagrange equations for the large deviation rate functional (\\ref{LDf}) can be written explicitly as follows (see \\cite{KOS,KO} for details).\nConsider complex valued \nfunctions $z(x,y)$ and $w(x,y)$ such that \n\\begin{equation}\\label{arg}\narg(z(x,y))=\\pi \\pa_x h(x,y), \\ \\ arg(w(x,y))=-\\pi \\pa_y h(x,y).\n\\end{equation}\nThen the Euler-Lagrange equations for $h(x,y)$ can be written as a system of \nequations for $z(x,y)$ and $w(x,y)$ as\n\\begin{equation}\\label{DEhf}\n\\pa_y \\log (z)+\\pa_x \\log(w)=0, \\ \\ P(z,w)=0.\n\\end{equation}\n\n\\begin{figure}[t!]\n\\begin{center}\n\\includegraphics[width=1.0\\textwidth, angle=0.0, scale=0.6]{figure02.pdf}\n\\end{center}\n\\caption{\\label{Ell-1} The $xy$ plane with the denoted domains. The ellipse is the boundary of the limit shape. The height function smooth\ninside the ellipse and linear outside. Here, we also show the lines $n_x=0$, $n_y=0$ and areas $A$, $B$, $C$, and $D$. }\n\\end{figure}\n\n\\subsection{The limit shape}\nHere we describe the limit shape height function $h_0(x,y)$ for DW boundary \nconditions. We use the result of \\cite{Pronko} where the density of the horizontal edges \noccupied by the paths is derived.\n\nDefine the function $D(x,y)$ as\n$$\nD(x,y) = \\alpha (1-\\alpha) \\left[ \\frac{(y-x)^{2}}{\\alpha}+\\frac{(1-x-y)^{2}}{1-\\alpha}-1 \\right].\n$$\nHere, the parameter $\\alpha$ is determined by values of Boltzmann weights of the model as\n$$\n\\alpha=\\frac{b}{a}.\n$$\nIn Baxter's parametrization $ \\alpha=\\tan{(u)}$.\nDefine the region $E= \\{ (x,y) | D(x,y) \\le 0 \\}$ as the interior of the ellipse $\\partial E= \\{ (x,y) | D(x,y)= 0 \\}$\nwhich is inscribed in the square $0\\leq x \\leq 1, 0\\leq y \\leq 1$ as it is shown in Fig. \\ref{Ell-1}. The \nellipse is the boundary of the limit shape, or the ``arctic curve''~\\cite{Pronko2011}.\n\nThe following expression was derived in \\cite{Pronko}: \n\\begin{equation}\\label{DyHF}\n\\partial_{y} h_0(x,y) = \\left\\{\n \\begin{array}{cc}\n \\frac{1}{\\pi} \\text{arccot} \\left( \\frac{-n_{y}}{\\sqrt{-D(x,y)}} \\right), & (x,y) \\in E \\\\\n 0, & (x,y) \\in A \\cup B, \\enskip n_{y}<0 \\\\\n 1, & (x,y) \\in C \\cup D, \\enskip n_{y}>0 \\\\\n \\end{array}\n\\right.\n\\end{equation}\nRegions $A,B,C,D$ are shown in Fig. \\ref{Ell-1}\nHere $n_{y}=(1-\\alpha)(y-x)+\\alpha (1-x-y)=x+(2\\alpha-1) y -\\alpha$.\n\nIntegrating this expression, we obtain the following formula for \nthe limit shape height function itself:\n\\begin{equation}\\label{HF}\nh_0(x,y) = \\left\\{\n \\begin{array}{cc}\n \\frac{1}{\\pi} \\left( y \\, \\text{arccot} \\left[ \\frac{-n_{y}}{\\sqrt{-D(x,y)}} \\right] -\\frac{1}{2} \\, \\text{arctan}\\left[\\frac{-x^2+(y-\\alpha)(\\alpha-1)+x(1+y-2y\\alpha)}{(1-x-\\alpha)\\sqrt{-D(x,y)}} \\right] + \\right. & \\\\\n \\left. +(\\frac{1}{2}-x) \\, \\text{arctan} \\left[ \\frac{-n_{x}}{\\sqrt{-D(x,y)}} \\right] \\right) +\\frac{x}{2} & (x,y) \\in E \\mbox{ }\\& \\mbox{ } x<1-\\alpha \\\\\n\\frac{1}{\\pi} \\left( y \\, \\text{arccot} \\left[ \\frac{-n_{y}}{\\sqrt{-D(x,y)}} \\right] -\\frac{1}{2} \\, \\text{arctan}\\left[\\frac{-x^2+(y-\\alpha)(\\alpha-1)+x(1+y-2y\\alpha)}{(1-x-\\alpha)\\sqrt{-D(x,y)}} \\right] + \\right. & \\\\\n \\left. +(\\frac{1}{2}-x) \\, \\text{arctan} \\left[ \\frac{-n_{x}}{\\sqrt{-D(x,y)}} \\right] \\right) +\\frac{x}{2}-\\frac{1}{2} & (x,y) \\in E \\mbox{ }\\& \\mbox{ } x \\ge 1-\\alpha \\\\\n 0 & (x,y) \\in A \\\\\n x & (x,y) \\in B \\\\\n x+y-1 & (x,y) \\in C \\\\\n y & (x,y) \\in D\n \\end{array}\n\\right.\n\\end{equation}\nHere $n_{x}=(1-\\alpha)(y-x)+\\alpha (x+y-1)=y+(2\\alpha-1) x -\\alpha$.\n\\begin{figure}[b!]\n\\begin{center}\n\\vspace{1mm}\\includegraphics[width=0.4\\textwidth, angle=0.0]{figure03.pdf}\n\\end{center}\n\\caption{\\label{f1} The density of horizontal edges occupied by paths, or $\\partial_{y} h_0(x,y)$, for $\\alpha=9\/25$.}\n\\end{figure}\n\nDifferentiating this expression in $x$, we obtain the density of edges occupied with horizontal paths:\n\\begin{equation}\n\\label{partialX}\n\\partial_{x} h_0(x,y) = \\left\\{\n \\begin{array}{cc}\n -\\frac{1}{\\pi} \\text{arctan} \\left( \\frac{-n_{x}}{\\sqrt{-D(x,y)}} \\right)+\\frac{1}{2}, & (x,y) \\in E \\\\\n 0, & (x,y) \\in A \\cup D, \\enskip n_{x}<0 \\\\\n 1, & (x,y) \\in B \\cup C, \\enskip n_{x}>0 \\\\\n \\end{array}\n\\right.\n\\end{equation}\n\nHere we use the branch of the function $\\text{arctan}$ which behaves as\n\\begin{equation}\n-\\frac{1}{\\pi} \\text{arctan} \\left( \\frac{-n_{x}}{\\sqrt{-D(x,y)}} \\right) \\to \\left\\{\n \\begin{array}{cc}\n -\\frac{1}{2}, & n_{x}<0, \\enskip (x,y) \\in A \\cup D \\\\\n \\frac{1}{2}, & n_{x}>0, \\enskip (x,y) \\in B \\cup C \\\\\n \\end{array}\n\\right.\n\\end{equation}\nwhen $(x,y)$ approach to the boundary of $E$.\n\nAs an example, in Fig.~\\ref{f1} we show the partial derivative of the height function~(\\ref{HF}) for $\\alpha=9\/25$.\nInside the arctic curve, it is given by the nontrivial part of Eq.~(\\ref{DyHF}) and outside that one, it equals to zero or one. \nThe limit shape height function $h_{0}(x,y)$, Eq.~(\\ref{HF}), is shown in Fig.~\\ref{f2}.\n\\begin{figure}[t!]\n\\begin{center}\n\\vspace{1mm}\\includegraphics[width=0.4\\textwidth, angle=0.0]{figure04.pdf}\n\\end{center}\n\\caption{\\label{f2} The limit shape height function $h_{0}(x,y)$, Eq.~(\\ref{HF}), for $\\alpha=9\/25$.}\n\\end{figure}\n\n\\subsection{The function $z(x,y)$}\nAn important property of functions $z(x,y)$ and $w(x,y)$ is that $z$ maps \nthe inner part of the ellipse $D(x,y)=0$ (the arctic curve) to the upper half plane.\nIndeed, as we saw in the previous section, the $x$-derivative of the height function~(\\ref{partialX})\nis non-negative when $0\\leq x\\leq 1$ and, therefore, the imaginary part of the function $z(x,y)$\nis also non-negative.\n\nIn our case we already know the height function, so in order to find functions $z$ and $w$\nit is sufficient to solve the algebraic equation in (\\ref{DEhf}). \nMoreover, since we know the height function, we know the arguments of $z$ and $w$, therefore, we just have to solve\nthe equation $P(z,w)=0$ for absolute values of $z$ and $w$. This will give \nthe conformal mapping $z$ from the interior $E$ of our ellipse to the \nupper half of the complex plane.\n\nSolving the quadratic equation $P(z,w)=0$ for the absolute values of $z$ and $w$ and taking into account (\\ref{arg}),\nwe obtain:\n\\begin{eqnarray}\\label{zw}\n\\nonumber |z|=\\frac{1}{2 \\, a \\, b \\, \\sin[-\\pi \\pa_yh_0]}\\Bigg(a^2 \\sin[ \\pi (\\pa_xh_0-\\pa_yh_0) ]+b^2 \\sin[ \\pi (\\pa_xh_0+\\pa_yh_0)] \\\\\n\\hspace{1cm}\\mp \\sqrt{4 a^2 b^2 (\\sin[-\\pi \\pa_yh_0])^2+\\bigg(a^2 \\sin[\\pi (\\pa_xh_0-\\pa_yh_0)]+b^2 \\sin[\\pi (\\pa_xh_0+\\pa_yh_0)]\\bigg)^2}\\Bigg),\\label{z} \\\\\n\\nonumber |w|=\\frac{1}{2 \\, a \\, b \\, \\sin[\\pi \\pa_xh_{0}]}\\Bigg(a^2 \\sin[\\pi (\\pa_xh_0-\\pa_yh_0)]+b^2 \\sin[\\pi (\\pa_xh_0+\\pa_yh_0)] \\\\\n\\hspace{1cm}\\pm \\sqrt{4 a^2 b^2 (\\sin[-\\pi \\pa_yh_{0}])^2+\\bigg(a^2 \\sin[ \\pi (\\pa_xh_0-\\pa_yh_0)]+b^2 \\sin[ \\pi (\\pa_xh_0+\\pa_yh_0)])\\bigg)^2}\\Bigg) \\label{w}.\n\\end{eqnarray}\n\nWe almost constructed the mapping $z: E \\to H=\\{ z | Re(z) \\geq 0 \\}, \\pa E \\to \\mathbb R$. The last step is to determine the signs in (\\ref{z}).\nIn the Appendix~\\ref{A1}, we determine the signs and the mapping.\nIt maps the boundary of $E$ bijectively to the real line in the following way:\n\\begin{itemize}\n\\item $z: A\\cap \\pa E\\to (0, \\frac{a}{b})$, $A\\cap B\\mapsto 0$, $A\\cap D\\mapsto \\frac{a}{b}$\n\\item $z: B\\cap \\pa E\\to (-\\frac{b}{a}, 0)$, $B\\cap C\\mapsto -\\frac{b}{a}$, $B\\cap A\\mapsto 0$\n\\item $z: C\\cap \\pa E\\to (-\\infty, -\\frac{b}{a})$, $C\\cap D|_{C}\\mapsto -\\infty$, $C\\cap B\\mapsto -\\frac{b}{a}$\n\\item $z: D\\cap \\pa E\\to (\\frac{a}{b},\\infty)$, $D\\cap A\\mapsto \\frac{a}{b}$, $D\\cap C|_{D}\\mapsto \\infty$\n\\end{itemize}\n\n\\subsection{The two-point correlation function}\nIn the continuum limit, the fluctuations of the height function are described\nby the massless Euclidean quantum Bose field in the interior of the arctic curve with the metric determined\nby the second variation $S^{(2)}$ of the large deviation rate functional (\\ref{LDf}) computed at the limit shape height function.\nIt reads\n$$\nS^{(2)}[h_{0}]= \\frac{1}{2} \\iint_{\\mathcal D} \\left( \\partial_{1}^{2} \\sigma(\\vec{\\nabla} h_{0}) (\\partial_{x} \\phi)^{2}+\n2 \\partial_{1} \\partial_{2} \\sigma(\\vec{\\nabla} h_{0}) \\partial_{x} \\phi \\, \\partial_{y} \\phi +\n\\partial_{2}^{2} \\sigma(\\vec{\\nabla} h_{0}) (\\partial_{y} \\phi)^{2} \\right) \\, dx \\, dy.\n$$\n\nThe mapping $z$ brings the functional $S^{(2)}$ with the kernel defined on functions on the interior of $E$ to \nthe Dirichlet functional for the Laplace operator acting on functions on the upper half of the complex plane. \nThis defines the two-point correlation function for fluctuations of the height function on $E$\nas the Green's function for the Laplace operator on the upper half plane with Dirichlet boundary \nconditions on the real line:\n\\begin{equation}\n\\label{Corr2Pi}\n\\left \\langle \\phi(x_{1},y_{1}),\\phi(x_{2},y_{2}) \\right \\rangle = -\\frac{1}{2\\pi} \\ln \\Bigg| \\frac{z(x_{1},y_{1})-z(x_{2},y_{2})}{z(x_{1},y_{1})- \\overline{z(x_{2},y_{2})}} \\Bigg|.\n\\end{equation}\nHere, $\\phi$ is the fluctuation field from (\\ref{eq1chi}).\nThe formula (\\ref{Corr2Pi}) means that the two-point correlation function at the free fermionic point~($\\Delta=0$) has a logarithmic dependence on the distance between points $(x_{1},y_{1})$ and $(x_{2},y_{2})$, when this distance is small~\\cite{Kenyon}.\n\nFor free fermionic models, local correlation functions (multipoint correlation functions) $\\langle \\phi(\\vec{r}_{1}),\\ldots,\\phi(\\vec{r}_{n}) \\rangle$ of fluctuations of the height function are determined by the two-point correlation functions through the Wick's formula.\n\n\\begin{figure}[t!]\n\\begin{center}\n\\vspace{1mm}\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure05.pdf}\n\\end{center}\n\\caption{\\label{thermalizationFig} The behavior of the normalized volume under the height function values as a function of a number of flips for three different lattice sizes and three values of $\\Delta$: $0$, $1\/2$, and $-7\/2$ of the six-vertex model.}\n\\end{figure}\n\n\\section{Numerical results}\\label{num}\n\n\\subsection{Computation of observables and the thermalization.}\n\nAs it was mentioned in the introduction, we use the Markov chain sampling algorithm to generate a sequence of random states of the six-vertex model~\\cite{R2005}.\nThis method is known as the Markov chain Monte-Carlo simulation.\nIt is based on the special choice of the transition probabilities to transfer from an arbitrary distribution to the desired one.\nIt is also known as the Metropolis algorithm~\\cite{Metropolis}.\nAn overview of these numerical methods and their applications to statistical mechanics can be found in~\\cite{LandauBinder2014}.\nSee~\\cite{Zvonarev,AS,Keesman,Korepin1,Korepin2} for related numerical simulations.\n\nThe idea of Markov sampling is to create a random process that will follow the most likely states in the model.\nThis is guaranteed by the choice of the matrix of transition probabilities which is symmetrizable (detailed balanced condition) by the diagonal matrix with entries given by Boltzmann weights of the system.\nThis condition (plus an assumption of nondegeneracy of the largest eigenvalue) \nalso guarantees the asymptotical convergence of the process to the Boltzmann distribution \nstarting from any distribution. Also, in this case the Boltzmann distribution is the Perron-Frobenius \neigenvector of the matrix of transition probabilities\\footnote{These are all standard facts about Markov processes, for details see\nfor example \\cite{LandauBinder2014,Seneta,Markov}.}.\n\nWhen a random process is constructed, the expectation values of observables with respect to the \nBoltzmann distribution can be computed by averaging along the random process. \nThis procedure is especially effective when the Boltzmann distribution is concentrated in a small neighborhood of\nthe most likely state (the limit shape). In probability theory this is known as large deviations, in non-equilibrium \nstatistical physics this is known as a hydrodynamic limit. \n\nIn dimer models it was proven rigorously~\\cite{CKP} \nthat there exists a most probable state and the probability for any other state to be ``macroscopically distant'' from\nit is exponentially suppressed:\n\\begin{equation}\\label{localization}\nProb(h)\\propto \\exp\\left[N^2(S[h_{0}]-S[h])\\right].\n\\end{equation}\nHere, $h_0$ is the height function corresponding to the limit shape~(\\ref{HF}). It minimizes the large deviation rate functional.\nThe minimal value $S[h_{0}]$ is exactly (minus) the free energy of the system. \n\nThe six-vertex model with $\\Delta=0$ is equivalent to a dimer model.\nTherefore, in this case we can use the probability distribution and results from the corresponding dimer model.\nFor other values of $\\Delta$ in the six-vertex model the analysis is more complicated, but we expect a similar \nstructure of the distribution, suggesting the formation of the limit shape $h_{0}$.\n\nThe localization (concentration) of random states near the limit shape makes the numerical computation of observables\neasy once the Markov process is thermalized i.e. when it moves along the states in a vicinity of the limit shape.\nThus, the main challenge for computing observables is to know when the process is thermalized.\nUnfortunately, it is very hard to have an effective criterium for thermalization.\nInstead, we use a simple empirical technique: we monitor the fluctuations of the normalized volume under the height\nfunction\n\\[\nvol(h)=\\frac{1}{N^3}\\sum_{(n,m)} h(n,m).\n\\]\nAs it is clear from Fig.~\\ref{thermalizationFig}, the normalized volume ``drifts'', when the process is not yet thermalized.\nThen it starts to fluctuate around the normalized volume under the limit shape $h_{0}$.\nThus, we can start measurements to compute observables using the Markov chain simulations.\nFor example, Fig.~\\ref{thermalizationFig} shows that for the lattice of size $90\\times 90$ and $\\Delta=1\/2$\nit is safe to start averaging after about $10^{7}$ flips~\\cite{R2005}.\n\\begin{figure}[t!]\n\\begin{center}\n\\begin{tabular}{cc}\n\\begin{minipage}{0.5\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure06.pdf}\n\\end{minipage}\n&\n\\begin{minipage}{0.5\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure07.pdf}\n\\end{minipage}\n\\end{tabular}\n\\end{center}\n\\caption{\\label{FigCompare4040} (\\textit{left plot}) The numerical height function for $\\Delta=0$ and for the lattice size of $40\\times 40$. The parameter $\\alpha=9\/25$. The height function is the result of averaging over measurements. (\\textit{right plot}) The difference between the theoretical limit shape height function and the one obtained from numerical simulation.}\n\\end{figure}\n\nOnce the thermalization is achieved, we compute an observable by time averaging:\n\\begin{equation}\\label{O}\n\\left \\langle O\\right \\rangle \\simeq \\frac{O(s_1)+\\dots +O(s_K)}{K}.\n\\end{equation}\nHere $s_i$ is a random state at time $T_i$ counting from the first measurement, $K$ is the total number of measurements.\nThe right side depends on random states $s_i$ and is a random variable, but as $K\\to \\infty$\nit converges to the Boltzmann expectation value.\nOf course, numerically $K\\to \\infty$ simply means large values.\nWe will use this to compute the limit shape $h_{0}$\nand correlation functions.\nIn particular, the two-point correlation function of points $(x_{i},y_{i})$ and $(x_{j},y_{j})$ is calculated as\n\\begin{equation}\\label{ncf}\n\\left \\langle\\phi(x_{i},y_{i}),\\phi(x_{j},y_{j})\\right \\rangle = \\left \\langle \\chi(x_{i},y_{i}) \\chi(x_{j},y_{j}) \\right \\rangle-\\left \\langle \\chi(x_{i},y_{i})\\right \\rangle \\left \\langle \\chi(x_{j},y_{j}) \\right \\rangle,\n\\end{equation}\nwhere \n\\begin{equation}\\label{nhf}\n\\left \\langle \\chi(x_{i},y_{i})\\right \\rangle = \\frac{1}{K}\\sum_{k=1}^K \\chi_k(x_{i},y_{i}),\n\\end{equation}\nthe height function $\\chi$ is from Eq.~(\\ref{eq1chi}),\nand indices $i,j=1,\\ldots,N$ numerate the lattice sites.\nHere $\\chi_{k}$ are random variables, but the sum represents a deterministic quantity as $K\\to\\infty$.\n\\begin{figure}[t!]\n\\begin{center}\n\\begin{tabular}{cc}\n\\begin{minipage}{0.5\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure08.pdf}\n\\end{minipage}\n&\n\\begin{minipage}{0.5\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure09.pdf}\n\\end{minipage} \\\\\n\\begin{minipage}{0.5\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure10.pdf}\n\\end{minipage}\n&\n\\begin{minipage}{0.5\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure11.pdf}\n\\end{minipage}\n\\end{tabular}\n\\end{center}\n\\caption{\\label{CorrCompareTheor406090} The plots of (a) the exact limit shape two-point correlation function inside the arctic curve for $\\Delta=0$, (b) the difference between the numerical correlation function on the square lattice of $60\\times 60$ and the exact one, the numerical ones on the square lattices of sizes (c) $40\\times 40$ and (d) $90\\times 90$. The parameter $\\alpha=9\/25$.}\n\\end{figure}\n\n\\subsection{Numerical computation of the limit shape and correlation functions at the free fermionic point.}\nWe start by comparison of the calculated height function with the exact one, the limit shape $h_{0}$ given by Eq.~(\\ref{HF}), for $\\Delta=0$ and $\\alpha=9\/25$.\nThe difference between the exact height function and the numerical one is shown in Fig.~\\ref{FigCompare4040}.\nIt should be noted that the numerical height function is smooth since it was averaged over a number of measurements, as described above.\nThe difference between $h_{0}$ and the numerical height function reveals the Airy asymptotic near the boundary of the limit shape.\nThe difference vanishes as the lattice size $N\\to \\infty$.\n\\begin{figure}[t!]\n\\begin{center}\n\\vspace{1mm}\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure12.pdf}\n\\end{center}\n\\caption{\\label{DetailedCorrCompare6060} The values of the two-point correlation function along the slice $y=0.5$ for $\\Delta=0$. The curve shows the theoretical limit shape correlation function~(\\ref{Corr2Pi}). The points are the calculated values for different lattice sizes. The inset shows the values of numerically obtained coefficient against logarithm in Eq.~(\\ref{Corr2Pi}) with respect to the lattice size as well as a fit of these coefficients.}\n\\end{figure}\n\nThe results of computations of two-point correlation functions $\\left \\langle\\phi(x_{i},y_{i}),\\phi(x_{j},y_{j})\\right \\rangle$ are presented in Fig.~\\ref{CorrCompareTheor406090}.\nWe show plots of data~(\\ref{ncf}) for one point $(x_i,y_i)=(0.5,0.5)$ and another point $(x_j,y_j)$ running through the square $[0,1]\\times [0,1]$ of the lattice domain with step $1\/N$.\nThe plot (a) shows the theoretical limit shape correlation function, given by Eq.~(\\ref{Corr2Pi}).\nThe plots (c),(d) represent the computed values itself\nobtained in a ``single run'' of the Markov process described above for different linear sizes $N$ of the system.\nThe measurements are taken after thermalization after each several hundred thousand iterations of the process.\nThe total number of measurements is $10^{5}$.\nThe plot (b) shows the difference between the theoretical exact values and the numerical computation.\nAgain, the difference shows the Airy waves propagating from the boundary of the limit shape.\nAs in the case of the height function, one can see the Airy waves which decrease with increasing of $N$.\n\nThe agreement of theoretical values of the correlation function and the corresponding numerical values can\nbe seen qualitatively by comparing pictures in Fig.~\\ref{CorrCompareTheor406090}.\nWhen $\\Delta=0$ the six-vertex model maps to a dimer model and therefore in the limit $N\\to \\infty$\ncorrelation functions converge to conformally invariant correlation functions (\\ref{Corr2Pi}).\nOne can see the logarithmic behavior in the two-point correlation function\nin the vicinity of $(x_i,y_i)$, where $\\left \\langle\\phi(\\vec{r_{i}}),\\phi(\\vec{r}_{j})\\right \\rangle \\sim -1\/(2\\pi)\\, \\ln{|\\vec{r}_{i}-\\vec{r}_{j}|}$.\n\nBy plotting the results of numerics along the slice $y=0.5$ we can examine the logarithmic behavior carefully, see Fig.~\\ref{DetailedCorrCompare6060}.\nThere is a good agreement between the theoretical result and numerical data for different lattice sizes:\nthe calculated values converge to the theoretical prediction as the lattice size increases.\nThe numerical values of the coefficient against the logarithm in Eq.~(\\ref{Corr2Pi}) have been obtained from the fit to data.\nThey are shown in the inset of Fig.~\\ref{DetailedCorrCompare6060}.\nWe see that, as the lattice size increases, $N\\to \\infty$, the value of the coefficient against the logarithm tends to the exact one $-1\/(2\\pi)\\approx -0.159155$.\nFor example, a fit by logarithm for the lattice size $90\\times 90$ yields the coefficient $-0.167$.\nA fit of the values of the coefficient (red line in the inset), in turn, gives the approximate value for the infinite lattice to be $-0.1584\\pm0.0082$, which is close to the exact one.\n\n\\subsection{Numerical results for $\\Delta=1\/2$}\n\nThe agreement of theoretical and numerical results at the free fermionic point, $\\Delta=0$, suggests that the numerics should work equally well for other values of $\\Delta$, where the analytical results are still unknown.\nHere, we present numerical results for $\\Delta=1\/2$.\nWe choose this value of $\\Delta$ randomly, but note that it is also known as the combinatorial point where the model has many extra interesting features~\\cite{ZJdisser}.\n\nThe results of numerical computation of the two-point correlation function for $a\/c=b\/c=1$\nare shown in Fig.~\\ref{f1Delta05}.\nThree plots correspond to three lattice sizes: $40\\times 40$, $60\\times 60$, and $90\\times 90$.\nThe behavior of the correlation function at short distances, as expected, is very similar to that for the free fermionic point in Fig.~\\ref{CorrCompareTheor406090}.\n\n\n\\begin{figure}[t!]\n\\begin{center}\n\\begin{tabular}{ccc}\n\\begin{minipage}{0.33\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure13.pdf}\n\\end{minipage}\n&\n\\begin{minipage}{0.33\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure14.pdf}\n\\end{minipage}\n&\n\\begin{minipage}{0.33\\linewidth}\n\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure15.pdf}\n\\end{minipage}\n\\end{tabular}\n\\end{center}\n\n\\caption{\\label{f1Delta05} The correlation function for the case $\\Delta=1\/2$, $a\/c=b\/c=1$. The shown data are for lattice sizes (a) $40\\times 40$, (b) $60\\times 60$, and (c) $90\\times 90$.}\n\\end{figure}\n\nThe numerical values of the two-point correlation function along the slice $y=0.5$ for $\\Delta=1\/2$ are shown in Fig.~\\ref{f2Delta05}.\nThe short distance asymptotic of the correlation function is again logarithmic. This is \nin agreement with the fact that the model is in the disordered phase. The difference with the \nfree fermionic case is that the global correlation function is not given by a conformal \nmapping anymore, but is given by an effective Gaussian field theory, see for example the discussion in~\\cite{GBDJ}.\nHowever, as in any disordered phase, at distances which are larger than the lattice step, but much smaller than the characteristic size of the lattice, the correlation functions are still given by an effective conformal field theory.\nIn the case of the six-vertex model, this is $c=1$ Gaussian CFT model with logarithmic correlators.\n\nThe fitted coefficient against logarithm in Eq.~(\\ref{Corr2Pi}) approaches the exact value $-1\/(2\\pi)$ as the lattice size $N \\to \\infty$ in this case as well.\nHowever, the numerical values of this coefficient when $\\Delta=1\/2$ are notably worse than the same values for $\\Delta=0$. For $\\Delta=1\/2$, the values for smaller lattices are systematically smaller than those for $\\Delta=0$.\nThey are naturally expected to converge to the exact value, but the rate of a convergence is less than for $\\Delta=0$.\nFor example, when the lattice size is $90\\times 90$, the fit by logarithm gives the value $-0.182$ for the coefficient against $\\log$.\nThe convergence is shown in the inset of Fig.~\\ref{f2Delta05} with the extrapolated value for the infinite lattice being $-0.1588\\pm0.0058$, which is still close to the expected $-1\/(2\\pi)$.\n\n\\begin{figure}[t!]\n\\begin{center}\n\\vspace{1mm}\\includegraphics[width=1.0\\textwidth, angle=0.0]{figure16.pdf}\n\\end{center}\n\\caption{\\label{f2Delta05}The values of the two-point correlation function along the slice $y=0.5$ for $\\Delta=1\/2$. The calculated values for different lattice sizes are shown. The inset shows the values of numerically obtained coefficient against logarithm in the logarithm-like fit with respect to the lattice size as well as a fit of these coefficients.}\n\\end{figure}\n\nAs in the case $\\Delta=0$ one can see the Airy waves near the boundary of the limit shape. They disappear when $N$ is increasing.\n\n\\begin{figure}[t!]\n\\begin{center}\n\\vspace{1mm}\\includegraphics[width=0.750\\textwidth, angle=0.0]{figure17.pdf}\n\\end{center}\n\\caption{\\label{af}The values of the two-point correlation function for the antiferroelectric phase $\\Delta=-7\/2$. The lattice size is $N=90$. The sharp peak is given by the exponential fall which is fitted as $0.484+0.516 \\exp{\\left((-99\\pm 4) |\\vec{r}_{i}-\\vec{r}_{j}|\\right)}$.}\n\\end{figure}\n\n\\subsection{Numerical results for $\\Delta=-7\/2$}\n\nFor $\\Delta<-1$, the antiferroelectric phase of the six-vertex model is opened in the form of a diamond shape droplet.\nThis droplet has already been observed earlier~\\cite{R2005,Zvonarev,Cugliandolo,Korepin1}.\nThe Gaussian field $\\phi(x,y)$ in this region is massive which predicts the exponential decay of correlation functions at short distances.\nThe numerical observation of the exponential decay of correlation function in this phase is challenging. \nIn order to carry out such computations of correlation functions at distances deep in the antiferroelectric droplet, \nthe characteristic length of the droplet should be much larger than the correlation length. But for these values of $N$ and $\\Delta$ \nthe thermalization is expected exponentially long~\\cite{Ra}.\n\nWe carried out calculations of two-point correlation functions for $\\Delta=-7\/2$ and the lattice size $N=90$.\nThe result is given in Fig.~\\ref{af}, and the thermalization was presented in Fig.~\\ref{thermalizationFig}.\nThe numerics are in qualitative agreement with the theoretical prediction that the correlation function should exponentially decrease.\nOne can see a sharp peak over a relatively flat background.\nThe sharp peak is given by the exponential fall and can be fitted as $0.484+0.516 \\exp{\\left((-99\\pm 4) |\\vec{r}_{i}-\\vec{r}_{j}|\\right)}$.\n\n\nThe background ``pillow'' in Fig.~\\ref{af} is expected to be a result of ``mesoscopic'' effects. The lattice size $N=90$ is relatively small and \ncorrelation functions get affected by the Airy processes on the boundaries of the disordered region. \nThe parallel GPU computations on large lattices may resolve this issue~\\cite{AS}.\nMore careful analysis of comparative values of the linear size of the droplet and of the correlation length\nwill be given in a separate publication both numerically and from the exact solution.\n\n\\section{Conclusion}\nIn this paper, we numerically calculated the two-point correlation functions for the six-vertex model with the domain wall boundary conditions.\nThe disordered ($|\\Delta|<1$) phase has mainly been studied.\nParticular attention was paid to the free fermionic point ($\\Delta=0$), for which the correlation function has been also obtained analytically in the thermodynamic limit, $N\\to\\infty$.\nThe logarithm-like behavior of correlation functions at the small scales has been confirmed.\nFor antiferroelectric phase, the exponential decrease of the correlator has been observed.\nThe numerics for $N=90$ and $\\Delta=-7\/2$ show that it might be interesting to study correlation functions in the mesoscopic region where the size of the antiferroelectric droplet is comparable to the correlation length in the antiferroelectric phase.\nWe plan to continue studies of correlation functions and, in particular, their asymptotics in the limit $\\Delta \\to -1-0$\nwhen the relatively small characteristic size of the droplet requires computations on large lattices.\nFor such lattices, the implementation of Markov sampling on GPU may be of great practical significance.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\nMost success in machine learning has been booked in the domain of supervised learning, when large quantities of labeled data are available.\nHowever it is commonly agreed that this is not a sustainable way forward, as we want algorithms that can learn new tasks and generalize to new domains even when there is no or very little labeled data available.\nOne succesful approach to unsupervised and semi-supervised learning are Generative Adversarial Networks (GANs) \\cite{goodfellow2014generative}, which are formulated as a min-max game between a generator $g_\\theta(z)$ which defines an implicit density $\\mathbb{Q}_\\theta$ over $\\X \\subset \\mathbb R^d$ from which we can sample, and a discriminator $D$ (or critic $f$) which measures a distance between the ``real'' $\\mathbb{P}_r$ and ``fake'' $\\mathbb{Q}_\\theta$ distribution.\nIn order to use GANs for semi-supervised learning \\cite{springenberg2015unsupervised,salimans2016improved} or unsupervised learning \\cite{chen2016infogan}, typically the discriminator network is in some sense shared with a learned classifier $p(y|x)$, with $y \\in \\{1 \\dots K\\}$ the discrete label set. \nThe underlying assumption is that features that are helpful towards telling real\/fake apart, are also helpful towards classification.\nSpecifically the idea of the discriminator having ``$K+1$'' output directions, with the $K+1$\\textsuperscript{th} ``fake'' direction competing with the K ``real'' directions \\cite{salimans2016improved} has lead to strong empirical results \\cite{dai2017good}.\n\nRecently the distance metric between distributions became a topic of interest \\cite{arjovsky2017towards,nowozin2016f,kaae2016amortised,mao2016least,arjovsky2017wasserstein,gulrajani2017improved,mroueh2017mcgan,mroueh2017fisher,li2017mmd,anon2017sobolev}.\nWe will focus on GANs with one class of distance metrics, Integral Probability Metrics or IPMs \\cite{muller1997integral,sriperumbudur2009integral,sriperumbudur2012empirical}.\nIPMs were first introduced in the GAN framework in Wasserstein GAN (WGAN) \\cite{arjovsky2017wasserstein,gulrajani2017improved},\nand also formed the basis for McGan \\cite{mroueh2017mcgan}, Fisher GAN \\cite{mroueh2017fisher} and Sobolev GAN \\cite{anon2017sobolev}.\nThe IPM distance between $\\mathbb{P}$ and $\\mathbb{Q}$ is defined as:\n\\[ \\sup_{f \\in \\mathcal{F}} \\mathbb{E}_{x\\sim \\mathbb{P}} f(x)- \\mathbb{E}_{x\\sim \\mathbb{Q}} f(x). \\]\nThe IPM definition looks for a witness function (or ``critic'') $f(x)$, which maximally discriminates between samples coming from the two distributions.\nThis naturally fits the GAN idea: we can parametrize the critic with a neural network which takes the place of the discriminator in the GAN framework.\nA crucial ingredient of the IPM metric is the function class $\\mathcal{F}$, which defines how the critic $f(x)$ is bounded,\nwhich in its turn \\emph{defines the metric} being measured between distributions.\n\nIn WGAN \\cite{arjovsky2017wasserstein,gulrajani2017improved} we approximate $\\mathcal{F}$ the class of Lipschitz functions.\nWGAN-GP \\cite{gulrajani2017improved} uses a pointwise gradient norm penalty.\nFisher GAN \\cite{mroueh2017fisher} introduces a tractable constraint on $\\mathbb{E}_{x \\sim \\mu} f^2(x)$ which is enforced on samples from $\\mu=\\frac{\\mathbb{P}+\\mathbb{Q}}{2}$.\nSobolev GAN \\cite{anon2017sobolev} introduces the tractable constraint on $\\mathbb{E}_{x \\sim \\mu } \\nor{\\nabla_x f(x)}^2$ on the same $\\mu$.\nIn this paper, we investigate which IPM formulations are amenable towards semi-supervised learning,\nand whether we can leverage the $K+1$ formulation of classical JSD-based GANs \\cite{salimans2016improved}.\n\n\\begin{figure}[t]\n\\centering\n\\includegraphics[width=0.75\\linewidth]{critic_ssl.pdf}\n\\caption{``K+1'' parametrization of the IPM critic for semi-supervised learning.}\n\\label{fig:sslcritic}\n\\end{figure}\n\nWe will investigate the ``$K+1$ parametrization\" of the critic as introduced in \\cite{mroueh2017fisher} (See Figure \\ref{fig:sslcritic}): \n\\[ \nf(x) =\\underbrace{ \\sum_{y=1}^K p(y|x)\\scalT{S_y}{\\Phi_{\\omega}(x)}}_{\\bm{f_{+}}: \\text{ ``real\" critic }} - \\underbrace{\\scalT{v}{\\Phi_{\\omega}(x)}}_{\\bm{f_-} :\\text{``fake\" critic}} \n\\]\n\\vskip -1.0em\nThe training objective for critic and classifier now becomes:\n\\[\n\n \\max_{S, \\Phi_{\\omega}, f } \\L_{D} = \\frac{1}{N} \\left(\\sum_{x \\in \\text{unl} } f(x) - \\sum_{z \\sim p_z} f(g_\\theta(z)) \\right) + \\L^{\\text{C}} (f_+, f_-, g_\\theta) - \\lambda_{CE} \\sum_{(x,y) \\in \\text{lab} } CE (p(y|x),y )\n\\]\nWhere the constraint term $\\L^\\text{C}$ can contain the real and fake terms of the critic $f_+$ and $f_-$ separately.\nNote that $p(y|x)=\\rm{Softmax}(\\scalT{S}{\\Phi_{\\omega}(x)})_y$ appears both in the critic formulation and in the Cross-Entropy term.\nIntuitively this critic uses the $K$ class directions of the classifier $S_y$ to define the ``real'' direction, which competes with another $K+1$\\textsuperscript{th} direction $v$ that indicates fake samples.\n\nWe empirically investigate the merit of this $K+1$ formulation in IPM-based GANs in Section \\ref{sec:Kp1}.\nIn Section \\ref{sec:constraint} we investigate on which parts of the critic $f, f_+, f_-$ to apply the constraint terms.\nFinally, in Section \\ref{sec:normalization}, we investigate the influence of activation normalization in the critic $f$, such as batch normalization (BN) \\cite{ioffe2015batch} and layer normalization (LN) \\cite{ba2016layer}.\n\n\\section{Experiments}\nWe will provide experimental results on CIFAR-10 \\cite{cifar10} and SVHN \\cite{netzer2011reading} using the CNN architectures as in \\cite{mroueh2017fisher} for $g_\\theta$ and $f$,\nwhich is very close to the discriminator architecture used in \\cite{salimans2016improved,dumoulin2016adversarially}.\nUnless noted otherwise this CNN will have no normalization (LN or BN) in the critic, but $g_\\theta$ always includes BN.\nSimilar to the standard procedure in other GAN papers we use 4k labeled samples for CIFAR-10, 1k labeled samples for SVHN, and do hyperparameter and model selection on the standard CIFAR-10 and SVHN validation split.\nHyperparameter details are in Appendix \\ref{sec:hypers}.\nWe provide purely supervised baseline results of the critic architecture in Appendix \\ref{sec:supervised}.\n\n\n\\subsection{$K+1$ formulations.} \\label{sec:Kp1}\nWe show in Table \\ref{tab:Kp1} the results for plain vs $K+1$ critic formulation.\n``Plain $\\scal{v}{\\Phi_\\omega(x)}$'' indicates the plain critic $f=\\scal{v}{\\Phi_\\omega(x)}$ not interacting with the classifier.\nFor the $K+1$ formulations, the constraints act on the full critic, except for the combination where Fisher acts on $f$ but Sobolev is applied to $f_-$ (see next Section).\n\nThe third and sixth column of Table \\ref{tab:Kp1} present results for a modified $K+1$ formulation $f^H = f_+^H - f_-$, where \n\\[ \nf_+^{H}(x) \n= \\sum_{y=1}^K p(y|x) \\log( p(y|x) )\n= \\sum_{y=1}^K p(y|x)\\scalT{S_y}{\\Phi_{\\omega}(x)} - \\log Z_Y(x)\n= f_+(x) - \\log Z_Y(x)\n\\]\n\nIn this formulation, $f_+^H(x)$ is the negative entropy $-H[p(y|x)]$ of the classifier for a given $x$, which is what is being optimized as the sole objective in \\cite{springenberg2015unsupervised}.\nThis $f_+^H$ is not sensitive to the magnitude of, or an additive bias to, the $\\scalT{S_y}{\\Phi_\\omega(x)}$ because of the normalization constant $\\log Z_Y(x) = \\log( \\sum_{y'=1}^K \\exp \\scalT{S_{y'}}{\\Phi_\\omega(x)})$.\nThis could either stabilize the training and shift more focus on the $K+1$\\textsuperscript{th} direction $v$,\nor could limit the effectiveness of $f_+$ in contrasting between real and fake samples - which makes it interesting to investigate.\n\nWe see that in the two succesful IPMs (Fisher and Fisher+Sobolev), the $K+1$ formulations give a strong gain\nover the plain formulation.\nBest results are obtained for the (unmodified) $K+1$ formulation and Sobolev + Fisher constraint.\n\n\\begin{takeaway}\nThe $K+1$ formulation is superior over the plain $f(x)=\\scal{v}{\\Phi_\\omega(x)}$ formulation.\n\\end{takeaway}\n\n\\begin{table}[t]\n\\centering\n\\caption{Results for different critic formulations. Note that the $K+1$ formulation is better across the board (except for WGAN-GP because of the harmful gradient penalty on the full critic $f$, see Section \\ref{sec:constraint}).\nIn gray we indicate all settings which failed to improve over straight-up supervised training with the small set of labeled samples (Appendix \\ref{sec:supervised}).\n}\n\\label{tab:Kp1}\n\\resizebox{\\textwidth}{!}{\n\\begin{tabular}{| l | l |l | l | l | l | l | }\n \\toprule\n & \\multicolumn{3}{| c |}{CIFAR-10 (4k)} & \\multicolumn{3}{| c |}{SVHN (1k)} \\\\\n & Plain $\\scal{v}{\\Phi_\\omega(x)}$ & $K+1$ & $K+1, f_+^H$ & Plain $\\scal{v}{\\Phi_\\omega(x)} $ & $K+1$ & $K+1, f_+^H$ \\\\\n\\midrule\nWGAN clip & \\cellcolor[gray]{0.9} $40.85$ & & & \\cellcolor[gray]{0.9} $28.65$ & & \\\\\nWGAN-GP & \\cellcolor[gray]{0.9} $44.64 \\pm 0.61$ & \\cellcolor[gray]{0.9} $48.92 \\pm 0.50$ & \\cellcolor[gray]{0.9} $48.81 \\pm 0.41$ & $17.51 \\pm 0.23$ & \\cellcolor[gray]{0.9} $31.59 \\pm 1.75$ & \\cellcolor[gray]{0.9} $33.85 \\pm 1.46$ \\\\\nFisher & $18.95 \\pm 0.32$ & $17.82 \\pm 0.43$ & $17.12 \\pm 0.11$ & $15.59 \\pm 0.67$ & $9.46 \\pm 0.32$ & $11.93 \\pm 0.26$ \\\\\nSobolev & \\cellcolor[gray]{0.9} $80.45 \\pm 0.59$ & \\cellcolor[gray]{0.9} $45.35 \\pm 0.62$ & \\cellcolor[gray]{0.9} $45.27 \\pm 0.33$ & \\cellcolor[gray]{0.9} $80.46 \\pm 0.62$ & \\cellcolor[gray]{0.9} $28.46 \\pm 1.04$ & \\cellcolor[gray]{0.9} $26.00 \\pm 0.43$ \\\\\nFisher + Sobolev & $21.44 \\pm 0.74$ & $\\bm{16.29 \\pm 0.42}$ & $22.68 \\pm 0.31$ & $13.23 \\pm 0.53$ & $\\bm{8.88 \\pm 0.84}$ & $12.01 \\pm 0.19$ \\\\\n\\bottomrule\n\\end{tabular}\n}\n\\vskip -0.13in\n\\end{table}\n\n\\subsection{$K+1$: constraints on $f$, $f_+$ or $f_-$} \\label{sec:constraint}\n\n\\begin{table}[b]\n\\centering\n\\vskip -0.13in\n\\caption{ Four constraint combinations (GP, S, F, F+S), acting on either $f$, $f_-$, $f_+$.\n Failed settings are color-marked:\n blue indicates experiments that only include a penalty on $f_-$,\n gray marks the experiments where a gradient constraint was acting on the full critic $f$.\n}\n\\label{tab:constraint}\n\\resizebox{0.40\\textwidth}{!}{\n\\begin{tabular}{| l | r | r |}\n\\toprule\n & CIFAR-10 (4k) & SVHN (1k) \\\\\n\\midrule\n\\rowcolor[gray]{0.9}\n$\\om_{GP}(f)$ & $49.28$ & $34.05$ \\\\\n\\rowcolor{myblue}\n$\\om_{GP}(f_{-})$ & $73.78$ & $70.61$ \\\\\n\\rowcolor[gray]{0.9}\n$\\om_S(f)$ & $45.55$ & $26.15$ \\\\\n\\rowcolor{myblue}\n$\\om_S(f_{-})$ & $71.38$ & $20.04$ \\\\\n$\\om_F(f)$ & $17.49$ & $9.16$ \\\\\n\\rowcolor{myblue}\n$\\om_F(f_{-})$ & $72.08$ & $38.33$ \\\\\n\\rowcolor[gray]{0.9}\n$\\om_F(f) , \\om_S(f)$ & $45.55$ & $29.30$ \\\\\n$\\om_F(f) , \\om_S(f_-)$ & $\\bm{17.08}$ & $8.27$ \\\\\n\\rowcolor[gray]{0.9}\n$\\om_F(f_-) , \\om_S(f)$ & $47.15$ & $26.31$ \\\\\n\\rowcolor{myblue}\n$\\om_F(f_-) , \\om_S(f_-)$ & $68.62$ & $37.50$ \\\\\n\\rowcolor[gray]{0.9}\n$\\om_F(f_+) , \\om_S(f)$ & $47.79$ & $27.99$ \\\\\n$\\om_F(f_+) , \\om_S(f_-)$ & $17.12$ & $\\bm{8.01}$ \\\\\n\\bottomrule\n\\end{tabular}\n}\n\\end{table}\n\nSimilar to the notation in \\cite{mroueh2017fisher,anon2017sobolev}\nwe write the Fisher and Sobolev constraint respectively as\n$\\om_{F}(f) = \\frac{1}{N} \\sum_{\\tilde{x} \\sim \\mu} f^2(\\tilde{x}) = 1$,\nand\n$\\om_{S}(f) = \\frac{1}{N} \\sum_{\\tilde{x} \\sim \\mu} \\nor{\\nabla_x f(\\tilde{x})}^2 = 1$.\nThe WGAN-GP constraint is\n$\\nor{\\nabla_x f(\\tilde{x})}=1$ for points interpolated between real and fake,\nthis is enforced through\n$\\om_{GP}(f) = \\sum_{\\tilde{x} \\sim \\mu_{GP}} (1 - \\nor{\\nabla_x f(\\tilde{x})})^2$.\nBoth $\\om_{F}$ and $\\om_{S}$ are enforced with augmented lagrange multipliers with hyperparameters $\\rho_F$ and $\\rho_S$,\nwhile $\\om_{GP}$ is enforced with a large fixed penalty weight $\\lambda_{GP}=10.0$.\n\nNote now that the constraints can be enforced on either $f_+$, $f_-$, or the full critic $f = f_+ - f_-$.\nTo ensure that the critic is bounded, at least some constraint has to be acting on $f_-$ directly or through $f$,\nwhile the boundedness of $f_+$ could in principle be ensured through the CE term on the small labeled set.\nNote there will be non-trivial interaction between $\\lambda_{GP}$, $\\rho_F$, $\\rho_{S}$ and $\\lambda_{CE}$.\n\nIn Table \\ref{tab:constraint} are results for the four different constraints and combinations acting on different parts of the critic.\nWe see that formulations with constraints only action on $f_-$ failed: clearly the CE term alone is not enough to constrain $f_+$.\nAnother important conclusion is that any combination where a gradient-norm constraint (Sobolev or WGAN-GP) is acting on the full critic $f$, the classifier is compromised:\nin these settings it is impossible for the network to fit even the small labeled training set (heavy underfitting), causing bad SSL performance.\n\n\\begin{takeaway}\nWe need some form of constraint acting on both $f_-$ and $f_+$; the CE term alone is not enough to control $f_+$.\n\\end{takeaway}\n\\begin{takeaway}\nConstraints including the gradient norm (WGAN-GP, Sobolev) should only act on $f_-$, otherwise the network underfits.\n\\end{takeaway}\n\n\n\\subsection{How to normalize the critic} \\label{sec:normalization}\n\\begin{table}[t]\n\\centering\n\\caption{How to normalize the critic. We see BN performs signficantly worse than other options for Fisher GAN (and is incompatible with GP\/Sobolev).\nThe layernorm formulation with singleton $\\mu,\\sigma^2$ statistics are superior to statistics per feature map.\nFisher GAN benefits from layernorm, while in Sobolev+Fisher no normalization is prefered - this gives the overall best result.\n}\n\\label{tab:normalization}\n\\resizebox{\\textwidth}{!}{\n\\begin{tabular}{lllll}\n\\toprule\n & CIFAR-10 (4k) & & SVHN (1k) & \\\\\nIPM Definition & Fisher & Fisher + Sobolev & Fisher & Fisher + Sobolev \\\\\n$f$ normalization & & & & \\\\\n\\midrule\nBatch Normalization & $24.14 \\pm 0.26$ & & $10.76 \\pm 0.28$ & \\\\\nLN ($\\mu,\\sigma^2 \\in \\mathbb R^{1\\times 1\\times 1}$) ($\\bm{g}, \\bm{b} \\in \\mathbb R^{C\\times 1 \\times 1}$)\n& $\\bm{16.45 \\pm 0.42}$ & $16.79 \\pm 0.11$ & $8.78 \\pm 0.68$ & $9.02 \\pm 0.29$ \\\\\nLN ($\\mu,\\sigma^2 \\in \\mathbb R^{1\\times 1\\times 1}$) ($\\bm{g}, \\bm{b} \\in \\mathbb R^{1\\times H \\times W}$)\n& $16.81 \\pm 0.36$ & $17.24 \\pm 0.34$ & $\\bm{8.55 \\pm 0.29}$ & $8.70 \\pm 0.67$ \\\\\nLN ($\\mu,\\sigma^2 \\in \\mathbb R^{C\\times 1\\times 1}$) ($\\bm{g}, \\bm{b} \\in \\mathbb R^{C\\times 1 \\times 1}$)\n & $20.44 \\pm 0.24$ & $20.09 \\pm 0.43$ & $9.18 \\pm 0.12$ & $9.19 \\pm 0.21$ \\\\\nLN ($\\mu,\\sigma^2 \\in \\mathbb R^{C\\times 1\\times 1}$) ($\\bm{g}, \\bm{b} \\in \\mathbb R^{1\\times H \\times W}$)\n & $19.93 \\pm 0.34$ & $19.85 \\pm 0.29$ & $9.69 \\pm 0.16$ & $9.29 \\pm 0.16$ \\\\\n No Normalization & $17.73 \\pm 0.56$ & $\\bm{16.33 \\pm 0.15}$ & $9.06 \\pm 0.46$ & $\\bm{8.31 \\pm 0.60}$ \\\\\n\\bottomrule\n\\end{tabular}\n}\n\\end{table}\n\nIn the original DCGAN \\cite{radford2015unsupervised} architecture, batch normalization (BN) \\cite{ioffe2015batch} was a crucial ingredient,\nboth in the generator and discriminator.\nEven though the original WGAN still relies on BN,\nboth WGAN-GP, Fisher GAN, and Sobolev GAN report strong results without any layerwise normalization.\nWhen the constraint involves a norm of the gradient, BN is problematic since it couples the different samples in the batch.\nHere we investigate as alternative to BN either layer normalization (LN) \\cite{ba2016layer} or having no normalization in the critic.\n\nOne important detail which usually glossed over, is how layernorm exactly is extended to the convolutional setting,\nwhere the activations at a given layer are $\\in \\mathbb R^{C \\times H \\times W}$.\nSpecifically we need to decide whether we will accumulate the statistics mean $\\mu$ and variance $\\sigma^2$ either into a singleton ($\\in \\mathbb R^{1\\times 1\\times 1}$), or separate per feature map ($\\in \\mathbb R^{C\\times 1\\times 1}$).\nSimilarly, we need to decide whether we will parametrize the scale $\\bm{g}$ and bias $\\bm{b}$ separate per feature map ($\\in \\mathbb R^{C\\times 1\\times 1}$), or separate per pixel ($\\in \\mathbb R^{1\\times H\\times W}$).\nIn the above, we used broadcasting notation, meaning that the singleton dimensions will be expanded to perform the elementwise operations.\nFor reference, in batch normalization for convolutional networks both mean $\\mu$, variance $\\sigma^2$, scale $\\bm{g}\/\\gamma$ and bias $\\bm{b}\/\\beta$ are collected separately per feature map, i.e. $\\in \\mathbb R^{C\\times 1\\times 1}$.\nImplementation-wise we follow \\cite{ren2016normalizing} in adding a small $\\epsilon$ \\emph{inside} the square root in $\\sqrt{\\sigma^2+\\epsilon}$.\nThe results in Table \\ref{tab:normalization} lead us to conclude:\n\n\\begin{takeaway}\nAvoid batchnorm, definitely when constraining the gradient norm in objective, but it also hurts for Fisher GAN!\n\\end{takeaway}\n\\begin{takeaway}\nThe layer normalization formulation with singleton stats ($\\mu,\\sigma^2 \\in \\mathbb R^{1\\times 1\\times 1}$) and parameters per feature map ($\\bm{g}, \\bm{b} \\in \\mathbb R^{C\\times 1 \\times 1}$) is superior.\nFisher GAN benefits from this LN, while for Sobolev+Fisher no normalization is better.\n\\end{takeaway}\n\n\\iffalse\n\\subsection{Class-conditional generator.}\nACGAN \\cite{odena2016conditional} defines a simple extension to GANs where a one-hot encoded class label is appended to the noise vector, which allows the generator to be conditioned on class $k$: $g_\\theta(z, y=k)$.\nAs described in \\cite{mroueh2017fisher}, we need to add a CE term to the objetive for the generator $\\L_G$ (with weight $\\lambda_{CE,G}=0.1$).\nIn Table \\ref{tab:conditional} we investigate whether a class-conditional generator helps with SSL.\n\n\\begin{table}[htb]\n\\centering\n\\caption{Class-conditional generator $g_\\theta(z, y=k)$ vs unconditional generator $g_\\theta(z)$.\nExperiments on CIFAR-10 with different number of labeled samples. All critics with $K+1$ formulation.\nIn almost all settings, class-conditioning slightly hurts rather than helps performance.\n}\n\\label{tab:conditional}\n\\resizebox{\\textwidth}{!}{\n\\begin{tabular}{lllll}\n\\toprule\nlabeledSamples & 1000 & 2000 & 4000 & 8000 \\\\\n\\midrule\nFisher BN Cond & $35.52$ & $29.79$ & $25.40$ & $20.71$ \\\\\nFisher BN Uncond & $36.80$ & $30.43$ & $24.35$ & $21.35$ \\\\\nFisher LN Cond & $20.99 \\pm 0.66$ & $19.01 \\pm 0.21$ & $17.41 \\pm 0.38$ & $15.50 \\pm 0.41$ \\\\\nFisher LN Uncond & $19.74 \\pm 0.21$ & $17.87 \\pm 0.38$ & $16.13 \\pm 0.53$ & $14.81 \\pm 0.16$ \\\\\nFisher No Norm Cond & $22.58 \\pm 0.43$ & $20.16 \\pm 0.27$ & $18.16 \\pm 0.53$ & $16.22 \\pm 0.18$ \\\\\nFisher No Norm Uncond & $21.49 \\pm 0.18$ & $19.20 \\pm 0.46$ & $17.30 \\pm 0.30$ & $15.57 \\pm 0.33$ \\\\\nSobolev+Fisher No Norm Cond & $21.40 \\pm 0.60$ & $18.65 \\pm 0.59$ & $16.96 \\pm 0.16$ & $15.13 \\pm 0.22$ \\\\\nSobolev+Fisher No Norm Uncond & $20.14 \\pm 0.21$ & $17.38 \\pm 0.10$ & $15.77 \\pm 0.19$ & $14.20 \\pm 0.08$ \\\\\n\\bottomrule\n\\end{tabular}\n}\n\\end{table}\n\\fi\n\n\\section{Conclusion}\nWe empirically investigated how different types of IPM-based Generative Adversarial Networks \ncan be used for semi-supervised learning.\nA comparison with literature results is given in Appendix~\\ref{sec:literature}.\nOur main conclusions are (1) the $K+1$ formulation works, (2) batch normalization should be avoided, also in Fisher GAN,\nand (3) gradient penalty constraints should act on $f_-$ only, not on the full critic which includes the classifier $p(y|x)$.\n\n\\bibliographystyle{unsrt}\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\nWe are concerned with the following nonlinear Kirchhoff type equations\n\\begin{equation}\n\\left\\{\n\\begin{array}{ll}\n-\\left( a\\int_{\\mathbb{R}^{N}}|\\nabla u|^{2}dx+b\\right) \\Delta u+V(x)u=h(x,u)\n& \\text{ in }\\mathbb{R}^{N}, \\\\\nu\\in H^{1}(\\mathbb{R}^{N}), &\n\\end{array\n\\right. \\label{1-0}\n\\end{equation\nwhere $N\\geq 1,a,b>0,$ $V\\in C(\\mathbb{R}^{N},\\mathbb{R})$ and $h\\in C\n\\mathbb{R}\\times \\mathbb{R}^{N},\\mathbb{R}).$\n\nKirchhoff type equations, of the form similar to Eq. $(\\ref{1-0}),$ are\nanalogous to the stationary case of equations that arise in the study of\nstring or membrane vibrations, namely,\n\\begin{equation}\nu_{tt}-\\left( a\\int_{\\Omega }|\\nabla u|^{2}dx+b\\right) \\Delta u=h(x,u)\\text{\nin }\\Omega , \\label{1-1}\n\\end{equation\nwhere $\\Omega $ is a bounded domain in $\\mathbb{R}^{N}.$ As an extension of\nthe classical D'Alembert's wave equation, Eq. $(\\ref{1-1})$ was first\npresented by Kirchhoff \\cite{K} in 1883 to describe the transversal\noscillations of a stretched string, particularly, taking into account the\nsubsequent change in string length caused by oscillations, where $u$ denotes\nthe displacement, $h$ is the external force and $b$ is the initial tension\nwhile $a$ is related to the intrinsic properties of the string, such as\nYoung's modulus. Equations of this type are often referred to as being\nnonlocal because of the presence of the integral.\n\nAfter the pioneering work of Pohozaev \\cite{P} and Lions \\cite{L}, the\nsolvability of the Kirchhoff type equation $(\\ref{1-1})$ has been\nwell-studied in general dimension by various authors, see for examples,\nD'Ancona-Shibata \\cite{DS}, D'Ancona-Spagnolo \\cite{DS1} and Nishihara \\cit\n{N}. More recently, the corresponding elliptic version like Eq. $(\\ref{1-0})\n$ has begun to receive much attention via variational methods. We refer the\nreader to \\cite{Az1,Az2,CKW,DPS,G,HL,HZ,LLS,LY,LLS1,N1,SZ,SW,SW1,TC,Y} and\nthe references therein.\n\nMost of researchers have of late years focused on the existence of positive\nsolutions, ground states, radial solutions and semiclassical states for Eq. \n(\\ref{1-0})$ in lower dimensions, i.e., $N=1,2,3.$ The typical way to deal\nwith such problem is to apply the mountain-pass theorem or the Nehari\nmanifold method. Owing to the fourth power of the nonlocal term, one usually\nassumes that the nonlinearity $h(x,u)$ is either $4$-superlinear at infinity\non $u$ in the sense tha\n\\begin{equation*}\n\\lim_{|u|\\rightarrow \\infty }\\frac{\\int_{0}^{u}h(x,s)ds}{u^{4}}=\\infty \\text{\nuniformly in }x\\in \\mathbb{R}^{N},\n\\end{equation*\nor satisfies the following (AR)-condition\n\\begin{equation}\n\\exists \\mu >4\\text{ such that }0<\\mu \\int_{0}^{u}h(x,s)ds\\leq h(x,u)u\\text{\nfor }u\\neq 0. \\label{1-10}\n\\end{equation\nFor example, $h(x,u)=f(x)\\left\\vert u\\right\\vert ^{p-2}u$ with $40;$\\newline\n$(ii)$ $N=4:$ one radial ground state solution exists if and only if \na<\\left( \\int_{\\mathbb{R}^{N}}|\\nabla \\overline{u}|^{2}dx\\right) ^{-1};\n\\newline\n$(iii)$ $N\\geq 5:$ one radial solution exists if and only if\n\\begin{equation*}\na\\leq \\left( \\frac{N-4}{N-2}\\right) ^{\\frac{N-2}{2}}\\frac{2}{(N-4)b^{\\frac\nN-4}{2}}\\int_{\\mathbb{R}^{N}}|\\nabla \\overline{u}|^{2}dx}.\n\\end{equation*}\n\nMotivated by these findings mentioned above, in the present paper we are\nlikewise interested in looking for positive solutions of Kirchhoff type\nequations. The problem we consider is thu\n\\begin{equation}\n\\left\\{\n\\begin{array}{ll}\n-\\left( a\\int_{\\mathbb{R}^{N}}|\\nabla u|^{2}dx+b\\right) \\Delta\nu+u=f(x)\\left\\vert u\\right\\vert ^{p-2}u & \\text{ in }\\mathbb{R}^{N}, \\\\\nu\\in H^{1}(\\mathbb{R}^{N}), &\n\\end{array\n\\right. \\tag{$E_{a}$}\n\\end{equation\nwhere $N\\geq 1,a,b>0,20.$\n\\end{itemize}\n\nEq. $(E_{a})$ is variational, and its solutions correspond to critical\npoints of the energy functional $J_{a}:H^{1}(\\mathbb{R}^{N})\\rightarrow\n\\mathbb{R}$ given by\n\\begin{equation*}\nJ_{a}\\left( u\\right) =\\frac{a}{4}\\left( \\int_{\\mathbb{R}^{N}}|\\nabla\nu|^{2}dx\\right) ^{2}+\\frac{1}{2}\\int_{\\mathbb{R}^{N}}(b|\\nabla\nu|^{2}+u^{2})dx-\\frac{1}{p}\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx.\n\\end{equation*\nFurthermore, one can see that $J_{a}$ is a $C^{1}$ functional with the\nderivative given b\n\\begin{equation*}\n\\left\\langle J_{a}^{\\prime }(u),\\varphi \\right\\rangle =\\left( a\\int_{\\mathbb\nR}^{N}}|\\nabla u|^{2}dx+b\\right) \\int_{\\mathbb{R}^{N}}\\nabla u\\nabla \\varphi\ndx+\\int_{\\mathbb{R}^{N}}(u\\varphi -f(x)|u|^{p-2}u\\varphi )dx\n\\end{equation*\nfor all $\\varphi \\in H^{1}(\\mathbb{R}^{N})$, where $J_{a}^{\\prime }$ denotes\nthe Fr\\'{e}chet derivative of $J_{a}.$\n\nDistinguishing from the existing literature, this paper is devoted to study\na series of questions as follows:\n\n\\begin{itemize}\n\\item[$(I)$] In spite of the amount of papers dealing with Eq. $(\\ref{1-0})\n, the geometric properties of the energy functional $J_{a}$ have not been\ndescribed in detail. One objective of this study is to shed some light on\nthe behavior of $J_{a}$. We will study whether $J_{a}$ is bounded below or\nnot, depending on the parameter $a$ and the dimension $N.$\n\n\\item[$(II)$] As we can see, the Nehari-Pohozaev manifold can help to find\npositive solutions with positive energy for Eq. $(\\ref{1-0})$ when the\nnonlinearity $h(x,u)$ does not satisfy the (AR)-condition (\\ref{1-10}) (see\n\\cite{G,LY,TC,Y}). However, to our knowledge, such approach is only valid\nfor the case of $N=3$. In our study, since the nonlinearity $f(x)\\left\\vert\nu\\right\\vert ^{p-2}u(20$ is a sharp constant of Gagliardo-Nirenberg inequality. Thus, there\nexist two positive numbers $\\overline{C}_{0}(N,p,f)$ and $\\underline{C\n_{0}(p,f)$ such tha\n\\begin{equation*}\n0<\\overline{\\mathbf{A}}_{f}\\leq \\overline{C}_{0}(N,p,f)\\text{ for }N\\geq 4\n\\end{equation*\nand\n\\begin{equation*}\n0<\\underline{\\mathbf{A}}_{f}\\leq \\underline{C}_{0}(p,f)\\text{ for }N=4.\n\\end{equation*}\n\nLe\n\\begin{equation}\n\\overline{a}_{\\ast }=\\frac{2(p-2)}{4-p}\\left( \\frac{4-p}{p}\\right) ^{2\/(p-2)\n\\overline{\\mathbf{A}}_{f}\\text{ for }N\\geq 4 \\label{15-3}\n\\end{equation\nand\n\\begin{equation*}\n\\underline{a}_{\\ast }=\\frac{2(p-2)}{4-p}\\left( \\frac{4-p}{p}\\right)\n^{2\/(p-2)}\\underline{\\mathbf{A}}_{f}\\text{ for }N=4.\n\\end{equation*}\n\nWe now summarize the first part of our main results as follows.\n\n\\begin{theorem}\n\\label{t0-1}Suppose that $20;$\\newline\n$(ii)$ If $N=4,$ then for each $0\\overline{a\n_{\\ast },$ $J_{a}$ is bounded below on $H^{1}(\\mathbb{R}^{N})$ and \n\\inf_{u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\{0\\}}J_{a}(u)>0;$\\newline\n$(iii)$ If $N\\geq 5,$ then $J_{a}$ is bounded below on $H^{1}(\\mathbb{R\n^{N}) $ for all $a>0.$ More precisely, for each $0\\overline{a}_{\\ast },$ there holds \n\\inf_{u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\{0\\}}J_{a}(u)>0.$\n\\end{theorem}\n\nFor brevity, we sum up the main result of Theorem \\ref{t0-1} with the table\nbelow\n\\begin{equation*}\n\\begin{tabular}{|c|c|c|c|c|}\n\\hline\n& $a>0$ & $00$ \\\\ \\hline\n$N\\geq 5$ & $\\inf J_{a}(u)>-\\infty $ & - & $\\inf J_{a}(u)<0$ & $\\inf\nJ_{a}(u)>0$ \\\\ \\hline\n\\end{tabular\n\\end{equation*}\n\nNote that $\\frac{p^{2\/\\left( p-2\\right) }}{2^{p\/\\left( p-2\\right) }}>1,$\nsince $20\n$ for $a>\\overline{a}_{\\ast }$, if $N\\geq 4$. However, we obtain the\nfollowing nonexistence result.\n\n\\begin{theorem}\n\\label{t0-2}Suppose that $N\\geq 4$ and condition $(D1)$ holds. Then for each\n$a>\\frac{p^{2\/\\left( p-2\\right) }}{2^{p\/\\left( p-2\\right) }}\\overline{a\n_{\\ast },$ Eq. $(E_{a})$ does not admit any nontrivial solutions.\n\\end{theorem}\n\nNext, we need the following assumption on $f.$\n\n\\begin{itemize}\n\\item[$\\left( D2\\right) $] $\\lim_{\\left\\vert x\\right\\vert \\rightarrow \\infty\n}f\\left( x\\right) =f_{\\infty }>0\\ $uniformly on$\\ \\mathbb{R}^{N}.$\n\\end{itemize}\n\n\\begin{theorem}\n\\label{t0-3}$(i)$ Suppose that $N\\geq 5$ and $f(x)\\equiv f_{\\infty }>0.$\nThen for each $00,$\nwhere $v_{a}^{+}$ is the positive solution as described in part $(i)$.\n\\end{itemize}\n\nThen for each $00,\n\\end{equation*\nwhere $\\mathbf{M}_{a}=\\{u\\in H^{1}(\\mathbb{R}^{N})\\backslash\n\\{0\\}:\\left\\langle J_{a}^{\\prime }(u),u\\right\\rangle =0\\}$ is the Nehari\nmanifold. We will show that $\\mathbf{M}_{a}(c)$ can be divided into two\nparts\n\\begin{equation*}\n\\mathbf{M}_{a}^{(1)}(c)=\\{u\\in \\mathbf{M}_{a}(c):\\Vert u\\Vert\n_{H^{1}}C_{2}\\},\n\\end{equation*\nin which each local minimizer of the functional $J_{a}$ is a critical point\nof $J_{a}$ in $H^{1}(\\mathbb{R}^{N})$. Our approach is to minimize the\nenergy functional $J_{a}$ on $\\mathbf{M}_{a}^{(1)}(c)$, where $J_{a}$ is\nbounded below and the minimizing sequence is bounded. In fact, such approach\nhas been applied in the study of Schrodinger-Poisson systems in $\\mathbb{R\n^{3}$ by us (see \\cite{SWF1,SWF2}).\n\nWe assume that $f$ satisfies the following condition:\n\n\\begin{itemize}\n\\item[$\\left( D4\\right) $] $f_{\\max }=\\sup_{x\\in \\mathbb{R}^{N}}f\\left(\nx\\right) <\\frac{f_{\\infty }}{D(p)^{(p-2)\/2}},$ wher\n\\begin{equation*}\nD(p)=\\left\\{\n\\begin{array}{ll}\n\\left( \\frac{4-p}{2}\\right) ^{1\/(p-2)}, & \\text{ if }21.\n\\end{equation*}\n\\end{remark}\n\nLet\n\\begin{equation}\n\\Lambda _{0}=\\left[ 1-D(p)\\left( \\frac{f_{\\max }}{f_{\\infty }}\\right)\n^{2\/(p-2)}\\right] \\left( \\frac{f_{\\infty }}{S_{p}^{p}}\\right) ^{2\/(p-2)},\n\\label{1-7}\n\\end{equation\nwhere $S_{p}$ is the best Sobolev constant for the embedding of $H^{1}\n\\mathbb{R}^{N})$ in $L^{p}(\\mathbb{R}^{N}).$ In particular, if $f(x)\\equiv\nf_{\\infty },$ then equality (\\ref{1-7}) becomes\n\\begin{equation*}\n\\Lambda _{0}=(1-D(p))\\left( \\frac{f_{\\infty }}{S_{p}^{p}}\\right) ^{2\/(p-2)}.\n\\end{equation*\nSe\n\\begin{equation*}\n\\Lambda =\\left\\{\n\\begin{array}{ll}\n\\frac{4-p}{2}\\left( \\frac{f_{\\infty }(4-p)}{2pS_{p}^{p}}\\right) ^{2\/(p-2)} &\n\\text{ if }N=1,2,3, \\\\\n\\min \\left\\{ \\frac{p-2}{2(4-p)}\\left( \\frac{4-p}{p}\\right) ^{2\/(p-2)}\\Lambda\n_{0},\\overline{a}_{\\ast }\\right\\} & \\text{ if }N\\geq 4\n\\end{array\n\\right.\n\\end{equation*\nLet $w_{0}$ be the unique positive solution of the following Schr\\\"{o}dinger\nequatio\n\\begin{equation}\n\\begin{array}{ll}\n-\\Delta u+u=f_{\\infty }|u|^{p-2}u & \\text{ in }\\mathbb{R}^{N}\n\\end{array}\n\\tag*{$\\left( E_{0}^{\\infty }\\right) $}\n\\end{equation\nFrom \\cite{K}, we see that\n\\begin{equation}\n\\left\\Vert w_{0}\\right\\Vert _{H^{1}}^{2}=\\int_{\\mathbb{R}^{N}}f_{\\infty\n}|w_{0}|^{p}dx=\\left( \\frac{S_{p}^{p}}{f_{\\infty }}\\right) ^{2\/(p-2)},\n\\label{1-8}\n\\end{equation\nand\n\\begin{equation*}\nJ_{0}^{\\infty }(w_{0})=\\frac{p-2}{2p}\\left( \\frac{S_{p}^{p}}{f_{\\infty }\n\\right) ^{2\/(p-2)},\n\\end{equation*\nwhere $J_{0}^{\\infty }$ is the energy functional of equation $(E_{0}^{\\infty\n})$ in $H^{1}(\\mathbb{R}^{N})$ in the form\n\\begin{equation*}\nJ_{0}^{\\infty }(u)=\\frac{1}{2}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-\\frac{1}{\n}\\int_{\\mathbb{R}^{N}}f_{\\infty }|u|^{p}dx.\n\\end{equation*}\n\nWe now summarize the second part of our main results as follows.\n\n\\begin{theorem}\n\\label{t1}Assume that $f(x)\\equiv f_{\\infty }>0.$ Then the following\nstatements are true.\\newline\n$\\left( i\\right) $ If $N\\geq 1,$ then for each $0\\frac{p-2}{2\n}\\left( \\frac{S_{p}^{p}}{f_{\\infty }}\\right) ^{2\/(p-2)}>0;\n\\end{equation*\n\\newline\n$\\left( ii\\right) $ If $1\\leq N\\leq 4,$ then for each $00,$\nwhere $v_{a}^{-}$ is the positive solution as described in Theorem \\ref{t1}.\n\\end{itemize}\n\nThen for each $0\\frac{p-\n}{4p}\\left( \\frac{S_{p}^{p}}{f_{\\max }}\\right) ^{2\/(p-2)}.\n\\end{equation*}\n\\end{theorem}\n\n\\begin{theorem}\n\\label{t3}Suppose that $N\\geq 5$ and conditions $(D1)-(D5)$ hold. Then for\neach $00$ when $N=3$ and for $a>0$ sufficiently small when $N=4.$ In the\nfollowing, we shall further describe some characteristics of such solution\ndepending on $a$ and $f_{\\infty }$, which are not concerned in \\cite{Az1,Az2\n.\n\nDefine the fibering map $h_{a,u}:t\\rightarrow J_{a}\\left( tu\\right) $ a\n\\begin{equation*}\nh_{a,u}\\left( t\\right) =\\frac{t^{2}}{2}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}\n\\frac{at^{4}}{4}\\left( \\int_{\\mathbb{R}^{N}}\\left\\vert \\nabla u\\right\\vert\n^{2}dx\\right) ^{2}-\\frac{t^{p}}{p}\\int_{\\mathbb{R}^{N}}f(x)\\left\\vert\nu\\right\\vert ^{p}dx\\text{ for }t>0.\n\\end{equation*\nAbout its theory and application, we refer the reader to \\cite{BZ,DP}. Note\nthat for $u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\left\\{ 0\\right\\} $ and $t>0,$\n$h_{a,u}^{\\prime }\\left( t\\right) =0$ holds if and only if $tu\\in \\mathbf{M\n_{a}$. In particular, $h_{a,u}^{\\prime }\\left( 1\\right) =0$ holds if and\nonly if $u\\in \\mathbf{M}_{a}.$ It is natural to split $\\mathbf{M}_{a}$ into\nthree parts corresponding to the local minima, local maxima and points of\ninflection. As a consequence, following \\cite{T}, we can define\n\\begin{eqnarray*}\n\\mathbf{M}_{a}^{+} &=&\\{u\\in \\mathbf{M}_{a}:h_{a,u}^{\\prime \\prime }\\left(\n1\\right) >0\\}, \\\\\n\\mathbf{M}_{a}^{0} &=&\\{u\\in \\mathbf{M}_{a}:h_{a,u}^{\\prime \\prime }\\left(\n1\\right) =0\\}, \\\\\n\\mathbf{M}_{a}^{-} &=&\\{u\\in \\mathbf{M}_{a}:h_{a,u}^{\\prime \\prime }\\left(\n1\\right) <0\\}.\n\\end{eqnarray*}\n\nFor $20, \\label{1-3} \\\\\n\\overline{A}_{0} &=&\\frac{p^{2}}{16}\\left( \\frac{f_{\\infty }}{S_{p}^{p}\n\\right) ^{2\/(p-2)}>0, \\label{1-4} \\\\\nA_{0}^{\\ast } &=&\\frac{p-2}{2}\\left( \\frac{4-p}{p}\\right) ^{\\left(\n4-p\\right) \/(p-2)}(f_{\\infty }C_{p}^{p})^{2\/(p-2)}>0. \\label{1-5}\n\\end{eqnarray}\n\nIt is clearly that $\\overline{A}_{0}>\\Lambda .$ We now state the last part\nof our main results as follows.\n\n\\begin{theorem}\n\\label{t5}Let $u_{0}$ be a nontrivial solution of Eq. $(E_{a})$ with \nf(x)\\equiv f_{\\infty }$. Then the following statements are true.\\newline\n$(i)$ When $N=3,$ for each $a>0$ with $\\sqrt{a^{2}+4}+\\frac{2}{a}\\geq A_{0},$\nthere holds $u_{0}\\in \\mathbf{M}_{a}^{-}.$ In particular, $v_{a}^{-}$ is a\nground state solution as in Theorem \\ref{t1} $\\left( i\\right) $ for $N=3.\n\\newline\n$(ii)$ When $N=4,$ for each $0A_{0}^{\\ast },$ there\nholds $u_{0}\\in \\mathbf{M}_{a}^{+}.$ In particular, $v_{a}^{-}$ is a ground\nstate solution as in Theorem \\ref{t1} $\\left( i\\right) $ for $N=4.$\n\\end{theorem}\n\n\\begin{remark}\n$(i)$ Note that $\\inf_{a>0}\\left( \\sqrt{a^{2}+4}+\\frac{2}{a}\\right) >0.$\nThen whe\n\\begin{equation*}\nf_{\\infty }\\geq \\left[ \\frac{3\\left( p-1\\right) (-p^{2}+2p+12)}\np^{2}(p-2)\\inf_{a>0}\\left( \\sqrt{a^{2}+4}+\\frac{2}{a}\\right) }\\right]\n^{\\left( p-2\\right) \/2}S_{p}^{p},\n\\end{equation*\nthere holds $\\sqrt{a^{2}+4}+\\frac{2}{a}\\geq A_{0}$ for all $a>0.$ This shows\nthat $u_{0}\\in \\mathbf{M}_{a}^{-}$ for all $a>0.$\\newline\n$(ii)$ If there is a number $a_{0}>0$ such that $\\sqrt{a_{0}^{2}+4}+\\frac{2}\na_{0}}\\left( \\frac{p}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)$ such tha\n\\begin{equation}\n\\inf_{t\\geq 0}J_{a}(tu)=\\inf_{\\left( \\frac{p}{4-p}\\right)\n^{1\/(p-2)}T_{f}(u)0.$\n\\end{lemma}\n\n\\begin{proof}\n$\\left( i\\right) $ For $u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\{0\\}$ and $t>0\n, it has\n\\begin{eqnarray*}\nh_{a,u}(t) &=&J_{a}(tu)=\\frac{t^{2}}{2}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}\n\\frac{at^{4}}{4}\\left( \\int_{\\mathbb{R}^{N}}|\\nabla u|^{2}dx\\right) ^{2}\n\\frac{t^{p}}{p}\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx \\\\\n&=&t^{4}\\left[ g(t)+\\frac{a}{4}\\left( \\int_{\\mathbb{R}^{N}}|\\nabla\nu|^{2}dx\\right) ^{2}\\right] ,\n\\end{eqnarray*\nwhere\n\\begin{equation*}\ng(t)=\\frac{t^{-2}}{2}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-\\frac{t^{p-4}}{p\n\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx.\n\\end{equation*\nClearly, $J_{a}(tu)=0$ if and only if\n\\begin{equation*}\ng(t)+\\frac{a}{4}\\left( \\int_{\\mathbb{R}^{N}}|\\nabla u|^{2}dx\\right) ^{2}=0.\n\\end{equation*\nIt is easy to see that\n\\begin{equation*}\ng(\\hat{t}_{a})=0,\\ \\lim_{t\\rightarrow 0^{+}}g(t)=\\infty \\ \\text{and \n\\lim_{t\\rightarrow \\infty }g(t)=0,\n\\end{equation*\nwhere $\\hat{t}_{a}=\\left( \\frac{p}{2}\\right) ^{1\/(p-2)}T_{f}(u).$ By\ncalculating the derivative of $g(t)$, we obtain\n\\begin{equation*}\ng^{\\prime }(t)=t^{-3}\\left[ -\\left\\Vert u\\right\\Vert _{H^{1}}^{2}+\\frac\n(4-p)t^{p-2}}{p}\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx\\right] ,\n\\end{equation*\nwhich implies that $g(t)$ is decreasing when $0\\left( \\frac{p}{4-p\n\\right) ^{1\/(p-2)}T_{f}(u).$ This indicates tha\n\\begin{eqnarray}\n\\inf_{t>0}g(t) &=&g\\left( \\left( \\frac{p}{4-p}\\right)\n^{1\/(p-2)}T_{f}(u)\\right) \\notag \\\\\n&=&-\\frac{p-2}{2(4-p)}\\left( \\frac{p\\left\\Vert u\\right\\Vert _{H^{1}}^{2}}\n(4-p)\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx}\\right) ^{-2\/(p-2)}\\Vert u\\Vert\n_{H^{1}}^{2}. \\label{3-5}\n\\end{eqnarray\nNote that\n\\begin{equation*}\n0a.\n\\end{equation*\nUsing the above inequality, together with $(\\ref{3-5}),$ leads t\n\\begin{equation*}\n\\inf_{t>0}g(t)<-\\frac{a}{4}\\left\\Vert u\\right\\Vert _{D^{1,2}}^{4}.\n\\end{equation*\nThis implies that there exist two numbers $\\widehat{t}_{a}^{(0)}$ and \n\\widehat{t}_{a}^{(1)}$ satisfying\n\\begin{equation*}\n0<\\widehat{t}_{a}^{(1)}<\\left( \\frac{p}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)\n\\widehat{t}_{a}^{(0)}\n\\end{equation*\nsuch that\n\\begin{equation*}\ng\\left( \\widehat{t}_{a}^{(j)}\\right) +\\frac{a}{4}\\left\\Vert u\\right\\Vert\n_{D^{1,2}}^{4}=0\\text{ for }j=0,1.\n\\end{equation*\nThat is,\n\\begin{equation*}\nJ_{a}\\left( \\widehat{t}_{a}^{(j)}u\\right) =0\\text{ for }j=0,1.\n\\end{equation*\nThus,\n\\begin{eqnarray*}\nJ_{a}\\left( \\left( \\frac{p}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)u\\right)&=&\\left[\n\\left( \\frac{p}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)\\right]^{4}\\left[ g\\left(\n\\left( \\frac{p}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)\\right) +\\frac{a}{4}\\left\\Vert\nu\\right\\Vert _{D^{1,2}}^{4}\\right] \\\\\n&<&0,\n\\end{eqnarray*\nand s\n\\begin{equation*}\n\\inf_{t\\geq 0}J_{a}(tu)<0.\n\\end{equation*\nNote tha\n\\begin{equation*}\nh_{a,u}^{\\prime }(t)=4t^{3}\\left( g(t)+\\frac{a}{4}\\left\\Vert u\\right\\Vert\n_{D^{1,2}}^{4}\\right) +t^{4}g^{\\prime }(t).\n\\end{equation*\nThen we have\n\\begin{equation*}\nh_{a,u}^{\\prime }(t)<0\\text{ for all }t\\in \\left( \\widehat{t\n_{a}^{(1)},\\left( \\frac{p}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)\\right]\n\\end{equation*\nan\n\\begin{equation*}\nh_{a,u}^{\\prime }\\left( \\widehat{t}_{a}^{(0)}\\right) >0.\n\\end{equation*\nConsequently, we arrive at inequality (\\ref{eqq17}).\\newline\n$\\left( ii\\right) $ For each $u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\{0\\},$\nwe can find a unique $t_{0}:=t_{0}(u)>0$ such that $h_{a,u}(t_{0})=0$ and \nh_{a,u}^{\\prime }(t_{0})=0.$ In fact, we only need to solve the system with\nrespect to the variables $t,a\n\\begin{equation*}\n\\left\\{\n\\begin{array}{c}\nh_{a,u}(t)=t^{2}\\left( \\frac{1}{2}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}+\\frac\nat^{2}}{4}\\left\\Vert u\\right\\Vert _{D^{1,2}}^{4}-\\frac{t^{p-2}}{p}\\int_\n\\mathbb{R}^{N}}f(x)|u|^{p}dx\\right) =0, \\\\\nh_{a,u}^{\\prime }(t)=t\\left( \\left\\Vert u\\right\\Vert\n_{H^{1}}^{2}+at^{2}\\left\\Vert u\\right\\Vert _{D^{1,2}}^{4}-t^{p-2}\\int_\n\\mathbb{R}^{N}}f(x)|u|^{p}dx\\right) =0\n\\end{array\n\\right.\n\\end{equation*\nA direct calculation shows tha\n\\begin{equation*}\nt_{0}(u)=\\left( \\frac{2(p-2)\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx}{ap\\left\\Vert\nu\\right\\Vert _{D^{1,2}}^{4}}\\right) ^{1\/(4-p)},\n\\end{equation*\nand accordingly,\n\\begin{eqnarray*}\na_{0}(u) &=&\\frac{2(p-2)(4-p)^{(4-p)\/(p-2)}\\left( \\int_{\\mathbb{R\n^{N}}f(x)|u|^{p}dx\\right) ^{2\/(p-2)}}{p^{2\/(p-2)}\\left\\Vert u\\right\\Vert\n_{D^{1,2}}^{4}\\left\\Vert u\\right\\Vert _{H^{1}}^{2(4-p)\/(p-2)}} \\\\\n&=&\\frac{2(p-2)(4-p)^{(4-p)\/(p-2)}}{p^{2\/(p-2)}}\\overline{A}_{f}(u).\n\\end{eqnarray*\nSince\n\\begin{equation*}\n\\overline{a}_{\\ast }=\\frac{2(p-2)(4-p)^{(4-p)\/(p-2)}}{p^{2\/(p-2)}}\\sup_{u\\in\nH^{1}(\\mathbb{R}^{N})\\backslash \\{0\\}}\\overline{A}_{f}(u)=\\frac\n2(p-2)(4-p)^{(4-p)\/(p-2)}}{p^{2\/(p-2)}}\\overline{\\mathbf{A}}_{f},\n\\end{equation*\nwe have for each $a\\geq \\overline{a}_{\\ast }$ and $u\\in H^{1}(\\mathbb{R\n^{N})\\backslash \\{0\\},$\n\\begin{equation*}\nh_{a,u}(t)=J_{a}(tu)\\geq 0\\text{ for all }t>0.\n\\end{equation*\nThis completes the proof.\n\\end{proof}\n\n\\begin{corollary}\n\\label{g17}Suppose that $N=4$ and condition $(D1)$ holds. Then for all \na\\geq \\overline{a}_{\\ast }$, $J_{a}$ is bounded below on $H^{1}(\\mathbb{R\n^{N})$ and $\\inf_{u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\{0\\}}J_{a}(u)\\geq 0.$\n\\end{corollary}\n\n\\begin{lemma}\n\\label{g11}Suppose that $N\\geq 5$ and condition $(D1)$ holds. Then for all \na>0,$ $J_{a}$ is bounded below on $H^{1}(\\mathbb{R}^{N})$ and there exist\nnumbers $\\widehat{r},\\widehat{R}_{a}>0$ such tha\n\\begin{equation*}\nJ_{a}(u)>0\\text{ for all }u\\in H^{1}(\\mathbb{R}^{N})\\text{ with \n0<\\left\\Vert u\\right\\Vert _{H^{1}}<\\widehat{r}\\text{ or }\\left\\Vert\nu\\right\\Vert _{H^{1}}\\geq \\widehat{R}_{a}.\n\\end{equation*\nFurthermore, for each $00,\n\\end{eqnarray*\nwhere $\\alpha =\\frac{2^{\\ast }(p-2)}{p(2^{\\ast }-2)}$ and $0<\\beta <\\frac\np(2^{\\ast }-p)}{2^{\\ast }-2}.$ This implies that $J_{a}(u)$ is bounded below\non $H^{1}(\\mathbb{R}^{N})$ for all $a>0.$ Moreover, for each $a>0,$ there\nexist\n\\begin{equation*}\nR_{a}:=\\left[ \\frac{42^{\\ast }}{a\\alpha p^{2}}\\left( f_{\\max }C_{p}^{p}\\beta\n^{-\\frac{(1-\\alpha )p}{2}}\\right) ^{\\frac{2^{\\ast }}{\\alpha p}}\\right]\n^{1\/(4-2^{\\ast })}>0\n\\end{equation*\nsuch tha\n\\begin{equation*}\nJ_{a}(u)>0\\text{ for all }u\\in H^{1}(\\mathbb{R}^{N})\\text{ with }\\left\\Vert\nu\\right\\Vert _{D^{1,2}}>R_{a}.\n\\end{equation*}\n\nLet\n\\begin{equation*}\n\\widehat{R}_{a}=\\left[ R_{a}+\\left( \\frac{1}{2}-\\frac{1}{(1-\\alpha )p\n\\right) ^{-1}\\frac{2^{\\ast }}{\\alpha p}R_{a}^{2^{\\ast }}\\right] ^{1\/2}.\n\\end{equation*\nWe now prove that\n\\begin{equation*}\nJ_{a}(u)>0\\text{ for all }u\\in H^{1}(\\mathbb{R}^{N})\\text{ with }\\left\\Vert\nu\\right\\Vert _{H^{1}}>\\widehat{R}_{a}.\n\\end{equation*\nLet $u\\in H^{1}(\\mathbb{R}^{N})$ with $\\left\\Vert u\\right\\Vert _{H^{1}}\\geq\n\\widehat{R}_{a}.$ If $\\left\\Vert u\\right\\Vert _{D^{1,2}}>R_{a},$ then the\nresult is done clearly. If $\\left\\Vert u\\right\\Vert _{D^{1,2}}\\left( \\frac{1}{2}-\\frac{\\beta }{\\left(\n1-\\alpha \\right) p^{2}}\\right) ^{-1}\\frac{2^{\\ast }}{\\alpha p}\\left( f_{\\max\n}C_{p}^{p}\\beta ^{-\\frac{(1-\\alpha )p}{2}}\\right) ^{\\frac{2^{\\ast }}{\\alpha \n}}R_{a}^{2^{\\ast }}.\n\\end{equation*\nIndeed, note tha\n\\begin{equation}\n\\frac{f_{\\max }}{p}\\int_{\\mathbb{R}^{N}}|u|^{p}dx\\leq \\frac{2^{\\ast }}\n\\alpha p^{2}}\\left( f_{\\max }C_{p}^{p}\\beta ^{-\\frac{(1-\\alpha )p}{2\n}\\right) ^{\\frac{2^{\\ast }}{\\alpha p}}R_{a}^{2^{\\ast }}+\\frac{\\beta }\n(1-\\alpha )p^{2}}\\int_{\\mathbb{R}^{N}}u^{2}dx. \\notag\n\\end{equation\nThen we hav\n\\begin{eqnarray*}\nJ_{a}(u) &\\geq &\\frac{a}{4}\\left\\Vert u\\right\\Vert _{D^{1,2}}^{4}+\\frac{1}{2\n\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-\\frac{f_{\\max }}{p}\\int_{\\mathbb{R\n^{N}}|u|^{p}dx \\\\\n&>&\\left( \\frac{1}{2}-\\frac{\\beta }{\\left( 1-\\alpha \\right) p^{2}}\\right)\n\\int_{\\mathbb{R}^{N}}u^{2}dx-\\frac{2^{\\ast }}{\\alpha p^{2}}\\left( f_{\\max\n}C_{p}^{p}\\beta ^{-\\frac{(1-\\alpha )p}{2}}\\right) ^{\\frac{2^{\\ast }}{\\alpha \n}}R_{a}^{2^{\\ast }} \\\\\n&>&0.\n\\end{eqnarray*\nThus, we obtain that there exists a positive number $\\widehat{R}_{a}>R_{a}$\nsuch tha\n\\begin{equation*}\nJ_{a}(u)>0\\text{ for all }u\\in H^{1}(\\mathbb{R}^{N})\\text{ with }\\left\\Vert\nu\\right\\Vert _{H^{1}}\\geq \\widehat{R}_{a}.\n\\end{equation*}\n\nMoreover, using the Sobolev inequality give\n\\begin{equation*}\nJ_{a}(u)\\geq \\left\\Vert u\\right\\Vert _{H^{1}}^{2}\\left( \\frac{1}{2}-\\frac\nf_{\\max }}{pS_{p}^{p}}\\left\\Vert u\\right\\Vert _{H^{1}}^{p-2}\\right) ,\n\\end{equation*\nwhich implies that there exists a numbe\n\\begin{equation*}\n\\widehat{r}:=\\left( \\frac{pS_{p}^{p}}{2f_{\\max }}\\right) ^{1\/(p-2)}>0\n\\end{equation*\nsuch tha\n\\begin{equation*}\nJ_{a}(u)>0\\text{ for all }u\\in H^{1}(\\mathbb{R}^{N})\\text{ with \n0<\\left\\Vert u\\right\\Vert _{H^{1}}<\\widehat{r}.\n\\end{equation*\nHence, we hav\n\\begin{equation*}\nJ_{a}(u)>0\\text{ for all }u\\in H^{1}(\\mathbb{R}^{N})\\text{ with \n0<\\left\\Vert u\\right\\Vert _{H^{1}}<\\widehat{r}\\text{ or }\\left\\Vert\nu\\right\\Vert _{H^{1}}>\\widehat{R}_{a}.\n\\end{equation*}\n\nIt follows from Lemma \\ref{g4}$(i)$ that for each $00\\text{\nfor all }u\\in \\mathbf{M}_{a}. \\label{2}\n\\end{equation\nMoreover, for all $u\\in \\mathbf{M}_{a},$ we have\n\\begin{eqnarray}\nh_{a,u}^{\\prime \\prime }(1) &=&\\left\\Vert u\\right\\Vert _{H^{1}}^{2}+3a\\left(\n\\int_{\\mathbb{R}^{N}}|\\nabla u|^{2}dx\\right) ^{2}-(p-1)\\int_{\\mathbb{R\n^{N}}f(x)|u|^{p}dx \\notag \\\\\n&=&-(p-2)\\left\\Vert u\\right\\Vert _{H^{1}}^{2}+a(4-p)\\left( \\int_{\\mathbb{R\n^{N}}|\\nabla u|^{2}dx\\right) ^{2} \\notag \\\\\n&=&-2\\left\\Vert u\\right\\Vert _{H^{1}}^{2}+(4-p)\\int_{\\mathbb{R\n^{N}}f(x)|u|^{p}dx. \\label{2-2}\n\\end{eqnarray\nThus, using $\\left( \\ref{2}\\right) $ and $(\\ref{2-2})$ gives\n\\begin{equation*}\nJ_{a}(u)=\\frac{1}{4}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-\\frac{(4-p)}{4p\n\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx>\\frac{p-2}{4p}\\left( \\frac{S_{p}^{p}}\nf_{\\max }}\\right) ^{2\/(p-2)}\\text{ for all }u\\in \\mathbf{M}_{a}^{-}\n\\end{equation*\nWe need the following conclusion.\n\n\\begin{lemma}\n\\label{g1}Suppose that $N\\geq 1$ and $2\\frac{p-2}{4p}\\left( \\frac{S_{p}^{p}}{f_{\\max }}\\right) ^{2\/(p-2)\n\\text{ for all }u\\in \\mathbf{M}_{a}^{-}.\n\\end{equation*}\n\\end{lemma}\n\n\\begin{lemma}\n\\label{g6}Suppose that $N=1,2,3$ and condition $(D1)$ holds$.$ Then for each\n$a>0$ and $u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\{0\\}$ satisfying\n\\begin{equation*}\n\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx>\\frac{p}{4-p}\\left( \\frac{2a(4-p)}{p-2\n\\right) ^{\\left( p-2\\right) \/2}\\left\\Vert u\\right\\Vert _{H^{1}}^{p},\n\\end{equation*\nthen there exist two numbers $t_{a}^{+}$ and $t_{a}^{-}$ satisfying\n\\begin{equation*}\nT_{f}(u)0.\n\\end{equation*\nClearly, $tu\\in \\mathbf{M}_{a}$ if and only if $m(t)+a\\left( \\int_{\\mathbb{R\n^{N}}|\\nabla u|^{2}dx\\right) ^{2}=0.$ A straightforward evaluation gives\n\\begin{equation*}\nm(T_{f}(u))=0,\\ \\lim_{t\\rightarrow 0^{+}}m(t)=\\infty \\text{ and \n\\lim_{t\\rightarrow \\infty }m(t)=0.\n\\end{equation*\nSince $2\\left( \\frac{2}{4-p}\\right)\n^{1\/(p-2)}T_{f}(u).$ This indicates that\n\\begin{equation}\n\\inf_{t>0}m(t)=m\\left( \\left( \\frac{2}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)\\right)\n. \\label{2-4}\n\\end{equation\nFor each $a>0$ and $u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\left\\{ 0\\right\\} $\nsatisfying\n\\begin{equation*}\n\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}dx>\\frac{p}{4-p}\\left( \\frac{2a(4-p)}{p-2\n\\right) ^{(p-2)\/2}\\left\\Vert u\\right\\Vert _{H^{1}}^{p},\n\\end{equation*\nwe can conclude that\n\\begin{equation*}\nm\\left( \\left( \\frac{2}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)\\right) <-2a\\left(\n\\frac{p}{2}\\right) ^{2\/(p-2)}\\left\\Vert u\\right\\Vert\n_{H^{1}}^{4}<-a\\left\\Vert u\\right\\Vert _{D^{1,2}}^{4},\n\\end{equation*\nwhere we have used the fact of $\\left( \\frac{p}{2}\\right) ^{2\/(p-2)}>1$.\nMoreover, by Remark $\\ref{r-1}$ we have\n\\begin{equation}\nT_{f}(u)<\\sqrt{D(p)}\\left( \\frac{2}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)<\\left(\n\\frac{2}{4-p}\\right) ^{1\/(p-2)}T_{f}(u), \\label{2-3}\n\\end{equation\nand a direct calculation shows that\n\\begin{equation}\n\\frac{\\left( \\frac{2}{4-p}\\right) D(p)^{(p-2)\/2}-1}{D(p)\\left( \\frac{2}{4-p\n\\right) ^{2\/(p-2)}}>\\frac{p-2}{2(4-p)}\\left( \\frac{4-p}{p}\\right) ^{2\/(p-2)}.\n\\label{2-1}\n\\end{equation\nIt follows from $(\\ref{2-4})-(\\ref{2-1})$ tha\n\\begin{equation*}\nm\\left( \\sqrt{D(p)}\\left( \\frac{2}{4-p}\\right) ^{1\/(p-2)}T_{f}(u)\\right)\n<-a\\left\\Vert u\\right\\Vert _{D^{1,2}}^{4}.\n\\end{equation*\nThus, there exist two numbers $t_{a}^{+},t_{a}^{-}>0$ which satisfy\n\\begin{equation*}\nT_{f}(u)0.\n\\end{eqnarray*\nThese imply that $t_{a}^{\\pm }u\\in \\mathbf{M}_{a}^{\\pm }.$ Note that\n\\begin{equation*}\nh_{a,u}^{\\prime }(t)=t^{3}\\left( m(t)+a\\left( \\int_{\\mathbb{R}^{N}}|\\nabla\nu|^{2}dx\\right) ^{2}\\right) .\n\\end{equation*\nThen one can see that $h_{a,u}^{\\prime }(t)>0$ for all $t\\in\n(0,t_{a}^{-})\\cup (t_{a}^{+},\\infty )$ and $h_{a,u}^{\\prime }(t)<0$ for all \nt\\in (t_{a}^{-},t_{a}^{+})$. It leads to\n\\begin{equation*}\nJ_{a}(t_{a}^{-}u)=\\sup_{0\\leq t\\leq t_{a}^{+}}J_{a}(tu)\\text{ and \nJ_{a}(t_{a}^{+}u)=\\inf_{t\\geq t_{a}^{-}}J_{a}(tu),\n\\end{equation*\nand so $J_{a}\\left( t_{a}^{+}u\\right) &J_{a}(u) \\\\\n&\\geq &\\frac{p-2}{2p}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-\\frac{a(4-p)}{4p\n\\left\\Vert u\\right\\Vert _{H^{1}}^{4},\n\\end{eqnarray*\nwhich implies that for $0D_{2}.\n\\end{equation*\nThus, we obtain that\n\\begin{eqnarray}\n&&\\mathbf{M}_{a}\\left( \\frac{D(p)(p-2)}{2p}\\left( \\frac{2S_{p}^{p}}\nf_{\\infty }(4-p)}\\right) ^{\\frac{2}{p-2}}\\right) \\notag \\\\\n&=&\\left\\{ u\\in \\mathbf{M}_{a}:J_{a}(u)<\\frac{D(p)(p-2)}{2p}\\left( \\frac\n2S_{p}^{p}}{f_{\\infty }(4-p)}\\right) ^{\\frac{2}{p-2}}\\right\\} \\notag \\\\\n&=&\\mathbf{M}_{a}^{(1)}\\cup \\mathbf{M}_{a}^{(2)}, \\label{4-4}\n\\end{eqnarray\nwhere\n\\begin{equation*}\n\\mathbf{M}_{a}^{(1)}:=\\left\\{ u\\in \\mathbf{M}_{a}\\left( \\frac{D(p)(p-2)}{2p\n\\left( \\frac{2S_{p}^{p}}{f_{\\infty }(4-p)}\\right) ^{2\/(p-2)}\\right)\n:\\left\\Vert u\\right\\Vert _{H^{1}}D_{2}\\right\\} .\n\\end{equation*\nWe further have\n\\begin{equation}\n\\left\\Vert u\\right\\Vert _{H^{1}}D_{2}>\\sqrt{2}\\left( \\frac{2S_{p}^{p}}\nf_{\\infty }(4-p)}\\right) ^{\\frac{1}{p-2}}\\text{ for all }u\\in \\mathbf{M\n_{a}^{(2)}. \\label{4-2}\n\\end{equation\nUsing the Sobolev inequality, $(\\ref{2-2})$ and $(\\ref{4-1})$ gives\n\\begin{equation*}\nh_{a,u}^{\\prime \\prime }(1)\\leq -2\\left\\Vert u\\right\\Vert\n_{H^{1}}^{2}+(4-p)S_{p}^{-p}f_{\\max }\\left\\Vert u\\right\\Vert _{H^{1}}^{p}<\n\\text{ for all }u\\in \\mathbf{M}_{a}^{(1)}.\n\\end{equation*\nBy $(\\ref{4-2}),$ we derive that\n\\begin{eqnarray*}\n\\frac{1}{4}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-\\frac{(4-p)}{4p}\\int_\n\\mathbb{R}^{N}}f(x)|u|^{p}dx &=&J_{a}(u)<\\frac{D(p)(p-2)}{2p}\\left( \\frac\n2S_{p}^{p}}{f_{\\infty }(4-p)}\\right) ^{2\/(p-2)} \\\\\n&<&\\frac{p-2}{2p}\\left( \\frac{2S_{p}^{p}}{f_{\\infty }(4-p)}\\right) ^{2\/(p-2)}\n\\\\\n&<&\\frac{p-2}{4p}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}\\text{ for all }u\\in\n\\mathbf{M}_{a}^{(2)},\n\\end{eqnarray*\nwhich implies that\n\\begin{equation*}\n2\\left\\Vert u\\right\\Vert _{H^{1}}^{2}<(4-p)\\int_{\\mathbb{R}^{N}}f(x)|u|^{p}d\n\\text{ for all }u\\in \\mathbf{M}_{a}^{(2)}.\n\\end{equation*\nUsing the above inequality, together with $(\\ref{2-2})$, yields\n\\begin{equation*}\nh_{a,u}^{\\prime \\prime }\\left( 1\\right) =-2\\left\\Vert u\\right\\Vert\n_{H^{1}}^{2}+\\left( 4-p\\right) \\int_{\\mathbb{R}^{N}}f(x)\\left\\vert\nu\\right\\vert ^{p}dx>0\\text{ for all }u\\in \\mathbf{M}_{a}^{(2)}.\n\\end{equation*\nHence, we have the following result.\n\n\\begin{lemma}\n\\label{g7}For $N\\geq 1$ and $04k-4+2N.$ This implies that $J_{a}$ is not bounded below on \nH^{1}(\\mathbb{R}^{N})\\ $for $N=1,2,3.$\\newline\n$(ii)$ It follows from Corollary \\ref{g17} that for each $a>\\overline{a\n_{\\ast },$ the energy functional $J_{a}$ is bounded below on $H^{1}(\\mathbb{\n}^{4})$ and $\\inf_{u\\in H^{1}(\\mathbb{R}^{4})\\backslash \\{0\\}}J_{a}(u)>0.$\nNext, we claim that for each $00,$ we hav\n\\begin{equation*}\nI(su)=s^{4}\\left( \\frac{a}{4}\\left\\Vert u\\right\\Vert _{D^{1,2}}^{4}\n\\overline{g}(s)\\right) ,\n\\end{equation*\nwhere\n\\begin{equation*}\n\\overline{g}(s)=\\frac{s^{-2}}{2}\\int_{\\mathbb{R}^{4}}u^{2}dx-\\frac{f_{\\min\n}s^{p-4}}{p}\\int_{\\mathbb{R}^{4}}|u|^{p}dx.\n\\end{equation*\nClearly, $I(su)=0$ if and only if $\\overline{g}(s)+\\frac{a}{4}\\left\\Vert\nu\\right\\Vert _{D^{1,2}}^{4}=0.$ It is not difficult to observe that \n\\overline{g}\\left( s_{a}\\right) =0,\\ \\lim_{s\\rightarrow 0^{+}}\\overline{g\n(s)=\\infty \\ $and$\\ \\lim_{s\\rightarrow \\infty }\\overline{g}(s)=0,$ where\n\\begin{equation*}\ns_{a}=\\left( \\frac{p\\int_{\\mathbb{R}^{4}}u^{2}dx}{2f_{\\min }\\int_{\\mathbb{R\n^{4}}\\left\\vert u\\right\\vert ^{p}dx}\\right) ^{1\/(p-2)}>0.\n\\end{equation*\nConsidering the derivative of $\\overline{g}(s)$, we find\n\\begin{equation*}\n\\overline{g}^{\\prime }(s)=s^{-3}\\left[ \\frac{s^{p-2}f_{\\min }(4-p)}{p}\\int_\n\\mathbb{R}^{4}}|u|^{p}dx-\\int_{\\mathbb{R}^{4}}u^{2}dx\\right] ,\n\\end{equation*\nwhich implies that $\\overline{g}(s)$ is decreasing when $0\\left( \\frac{p\\int_{\\mathbb{R\n^{4}}u^{2}dx}{(4-p)f_{\\min }\\int_{\\mathbb{R}^{4}}|u|^{p}dx}\\right)\n^{1\/(p-2)},$ and s\n\\begin{eqnarray}\n\\inf_{s>0}\\overline{g}(t) &=&\\overline{g}\\left( \\left( \\frac{p\\int_{\\mathbb{\n}^{4}}u^{2}dx}{(4-p)f_{\\min }\\int_{\\mathbb{R}^{4}}|u|^{p}dx}\\right)\n^{1\/(p-2)}\\right) \\notag \\\\\n&=&-\\frac{p-2}{2}\\left( \\frac{f_{\\min }\\int_{\\mathbb{R}^{4}}|u|^{p}dx}{p\n\\right) ^{2\/(p-2)}\\left( \\frac{4-p}{\\int_{\\mathbb{R}^{4}}u^{2}dx}\\right)\n^{\\left( 4-p\\right) \/(p-2)}<0. \\label{10-0}\n\\end{eqnarray\nSince $00}\\overline{g}(s)<-\\frac{a}{4}\\left\\Vert u\\right\\Vert\n_{D^{1,2}}^{4}. $ Se\n\\begin{equation*}\ns_{0}(u)=\\left( \\frac{p\\int_{\\mathbb{R}^{4}}u^{2}dx}{(4-p)f_{\\min }\\int_\n\\mathbb{R}^{4}}|u|^{p}dx}\\right) ^{1\/(p-2)}.\n\\end{equation*\nThen we obtain\n\\begin{equation}\nI(s_{0}(u)u)=s_{0}^{4}(u)\\left[ \\overline{g}(s_{0}(u))+\\frac{a}{4}\\left\\Vert\nu\\right\\Vert _{D^{1,2}}^{4}\\right] <0. \\label{10-11}\n\\end{equation\nLet $u_{0}=s_{0}\\left( u\\right) u$ and $v_{t}(x)=u_{0}(t^{-1}x).$ Then we\nhave\\newline\n$(ii-A)$ $\\int_{\\mathbb{R}^{4}}|\\nabla v_{t}(x)|^{2}dx=t^{2}\\int_{\\mathbb{R\n^{4}}|\\nabla u_{0}(x)|^{2}dx;$\\newline\n$(ii-B)$ $\\int_{\\mathbb{R}^{4}}|v_{t}(x)|^{2}dx=t^{4}\\int_{\\mathbb{R\n^{4}}|u_{0}(x)|^{2}dx;$\\newline\n$(ii-C)$ $\\int_{\\mathbb{R}^{4}}|v_{t}(x)|^{p}dx=t^{4}\\int_{\\mathbb{R\n^{4}}|u_{0}(x)|^{p}dx.$\\newline\nCombining the above conclusions with $(\\ref{10-11})$ give\n\\begin{eqnarray*}\nJ_{a}(v_{t}(x)) &\\leq &t^{4}\\left( \\frac{a}{4}\\left\\Vert u_{0}\\right\\Vert\n_{D^{1,2}}^{4}+\\frac{1}{2}\\int_{\\mathbb{R}^{N}}u_{0}^{2}dx-\\frac{f_{\\min }}{\n}\\int_{\\mathbb{R}^{4}}|u_{0}|^{p}dx\\right) +\\frac{t^{2}}{2}\\left\\Vert\nu_{0}\\right\\Vert _{D^{1,2}}^{2} \\\\\n&=&t^{4}I(u_{0})+\\frac{t^{2}}{2}\\left\\Vert u_{0}\\right\\Vert _{D^{1,2}}^{2} \\\\\n&\\rightarrow &-\\infty \\text{ as }t\\rightarrow \\infty ,\n\\end{eqnarray*\nwhich implies that for each $00$ for which the fibering map $h_{a,u}$ has a\ncritical point with second derivative zero at $t(u)$. Moreover, if $a>a(u)$,\nthen $h_{a,u}$ is increasing on $(0,\\infty )$ and has no critical point.\nNote that $\\sup_{u\\in H^{1}(\\mathbb{R}^{N})\\backslash \\left\\{ 0\\right\\}\n}a(u)=\\frac{p^{2\/\\left( p-2\\right) }}{2^{p\/\\left( p-2\\right) }}\\overline{a\n_{\\ast }$ by (\\ref{15-3}). Hence, the energy functional $J_{a}$ has no any\nnontrivial critical points for $a>\\frac{p^{2\/\\left( p-2\\right) }}\n2^{p\/\\left( p-2\\right) }}\\overline{a}_{\\ast }.$ Consequently, we complete\nthe proof.\n\nTo prove that Theorem \\ref{t0-3}\\textbf{, }we need the following result.\n\n\\begin{lemma}\n\\label{g12}Suppose that $N\\geq 5$ and condition $(D1)$ holds. Let $00$ such that\n\\begin{equation*}\n\\left\\Vert u_{n}\\right\\Vert _{H^{1}}\\leq \\widehat{R}_{a}\\text{ for }n\\text{\nlarge enough.}\n\\end{equation*\nConsequently, we complete the proof.\n\\end{proof}\n\n\\textbf{At the end of this section, we begin to prove Theorem \\ref{t0-3}: }\n(i)$ By Lemma \\ref{g12} and the Ekeland variational principle, for each $00$\nthere exist a number $R=R(\\theta )>0$ and a sequence $\\{z_{n}\\}\\subset\n\\mathbb{R}^{N}$ such tha\n\\begin{equation}\n\\int_{\\lbrack B_{R}(z_{n})]^{c}}(|\\nabla\nu_{n}(x)|^{2}+u_{n}^{2}(x))dx<\\theta \\text{ uniformly for }n\\geq 1.\n\\label{10-2}\n\\end{equation\nDefine a new sequence of functions $v_{n}:=u_{n}(\\cdot +z_{n})\\in H^{1}\n\\mathbb{R}^{N}).$ Clearly, $\\left\\langle (J_{a}^{\\infty })^{\\prime\n}(v_{n}),v_{n}\\right\\rangle =o(1)$ and $J_{a}^{\\infty }(v_{n})=\\inf_{u\\in\nH^{1}(\\mathbb{R}^{N})\\backslash \\{0\\}}J_{a}^{\\infty }(u)+o(1).$ By virtue of\n$(\\ref{10-2})$, for each $\\theta >0$ there exists a number $R=R(\\theta )>0$\nsuch tha\n\\begin{equation}\n\\int_{\\lbrack B_{R}(0)]^{c}}(|\\nabla v_{n}(x)|^{2}+v_{n}^{2}(x))dx<\\theta\n\\text{ uniformly for }n\\geq 1. \\label{10-3}\n\\end{equation\nSince $\\{v_{n}\\}$ is bounded in $H^{1}(\\mathbb{R}^{N}),$ one can assume that\nthere exist a subsequence $\\{v_{n}\\}$ and $v_{a}^{+}\\in H^{1}(\\mathbb{R\n^{N}) $ such tha\n\\begin{eqnarray}\nv_{n} &\\rightharpoonup &v_{a}^{+}\\text{ weakly in }H^{1}(\\mathbb{R}^{N}),\n\\label{10-4} \\\\\nv_{n} &\\rightarrow &v_{a}^{+}\\text{ strongly in }L_{loc}^{r}(\\mathbb{R}^{N}\n\\text{ for }2\\leq r<2^{\\ast }, \\label{10-5} \\\\\nv_{n} &\\rightarrow &v_{a}^{+}\\text{ a.e. in }\\mathbb{R}^{N}. \\notag\n\\end{eqnarray\nBy $(\\ref{10-3})-(\\ref{10-5})$ and Fatou's Lemma, for any $\\theta >0$ and\nsufficiently large $n$, there exists a number $R>0$ such tha\n\\begin{eqnarray*}\n&&\\int_{\\mathbb{R}^{3}}|v_{n}-v_{a}^{+}|^{p}dx \\\\\n&\\leq\n&\\int_{B_{R}(0)}|v_{n}-v_{a}^{+}|^{p}dx\n\\int_{[B_{R}(0)]^{c}}|v_{n}-v_{a}^{+}|^{p}dx \\\\\n&\\leq &\\theta +S_{p}^{-p}\\left[ \\int_{[B_{R}(0)]^{c}}(|\\nabla\nv_{n}|^{2}+v_{n}^{2})dx+\\int_{[B_{R}(0)]^{c}}(|\\nabla\nv_{a}^{+}|^{2}+(v_{a}^{+})^{2})dx\\right] ^{\\frac{p}{2}} \\\\\n&\\leq &\\theta +S_{p}^{-p}(2\\theta )^{\\frac{p}{2}},\n\\end{eqnarray*\nwhich implies that for every $p\\in (2,2^{\\ast }),\n\\begin{equation}\nv_{n}\\rightarrow v_{a}^{+}\\text{ strongly in }L^{p}(\\mathbb{R}^{N}).\n\\label{10-6}\n\\end{equation\nSince $\\left\\langle (J_{a}^{\\infty })^{\\prime }(v_{n}),v_{n}\\right\\rangle\n=o\\left( 1\\right) $ and $\\widehat{r}<\\left\\Vert v_{n}\\right\\Vert _{H^{1}}\n\\widehat{R}_{a},$ using $(\\ref{10-6})$ gives\n\\begin{equation*}\n\\int_{\\mathbb{R}^{N}}f_{\\infty }|v_{a}^{+}|^{p}dx\\geq \\left( \\frac{S_{p}^{p\n}{f_{\\infty }}\\right) ^{2\/(p-2)}>0,\n\\end{equation*\nwhich indicates that $v_{a}^{+}\\not\\equiv 0.$\n\nNext, we show that $v_{n}\\rightarrow v_{a}^{+}$ strongly in $H^{1}(\\mathbb{R\n^{N}).$ Suppose on the contrary. Then we have\n\\begin{equation}\n\\left\\Vert v_{a}^{+}\\right\\Vert _{H^{1}}<\\liminf_{n\\rightarrow \\infty }\\Vert\nv_{n}\\Vert _{H^{1}}. \\label{10-7}\n\\end{equation\nSimilar to the argument of Lemma \\ref{g6}, there exists a unique $t_{a}>0$\nsuch that\n\\begin{equation}\n(h_{a,v_{a}^{+}}^{\\infty })^{\\prime }(t_{a})=0, \\label{10-8}\n\\end{equation\nwhere $h_{a,u}^{\\infty }(t)=h_{a,u}\\left( t\\right) $ with $f(x)\\equiv\nf_{\\infty }$. Since $\\left\\langle (J_{a}^{\\infty })^{\\prime\n}(v_{n}),v_{n}\\right\\rangle =o(1)$, it follows from $(\\ref{10-6})-(\\ref{10-7\n)$ that\n\\begin{equation}\n(h_{a,v_{a}^{+}}^{\\infty })^{\\prime }(1)<0. \\label{10-9}\n\\end{equation\nCombining $(\\ref{10-8})-(\\ref{10-9})$ with the profile of \nh_{a,v_{a}^{+}}^{\\infty }(t)$ gives $t_{a}<1$. By $\\left( \\ref{10-6}\\right)\n-\\left( \\ref{10-7}\\right) $ again, we see $(h_{a,v_{n}}^{\\infty })^{\\prime\n}(t_{a})>0$ for sufficiently large $n$. Note that\n\\begin{equation*}\n\\left( h_{a,v_{n}}^{\\infty }\\right) ^{\\prime }(1)=o(1),\n\\end{equation*\nbecause of $\\left\\langle (J_{a}^{\\infty })^{\\prime\n}(v_{n}),v_{n}\\right\\rangle =o(1)$. Similar to the proof of Lemma \\ref{g6},\nwe obtain\n\\begin{equation*}\n\\left( h_{a,v_{n}}^{\\infty }\\right) ^{\\prime }(t)=t^{3}\\left( m^{\\infty\n}(t)+a\\left( \\int_{\\mathbb{R}^{N}}|\\nabla v_{n}|^{2}dx\\right) ^{2}\\right)\n\\text{ for }t>0,\n\\end{equation*\nwhere\n\\begin{equation*}\nm^{\\infty }(t):=t^{-2}\\left\\Vert v_{n}\\right\\Vert _{H^{1}}^{2}-t^{p-4}\\int_\n\\mathbb{R}^{N}}f_{\\infty }|v_{n}|^{p}dx.\n\\end{equation*\nOne can see that $m^{\\infty }(t)$ is decreasing for\n\\begin{equation*}\n01.\n\\end{equation*\nThis indicates that $\\left( h_{a,v_{n}}^{\\infty }\\right) ^{\\prime }(t)>0$\nfor $0\\frac{p}{4-p}\\left(\n\\frac{2a(4-p)}{p-2}\\right) ^{(p-2)\/2}\\left\\Vert w_{0}\\right\\Vert\n_{H^{1}}^{p},\n\\end{equation*\nfor all $00$ satisfyin\n\\begin{equation*}\n10$\nthere exist a positive constant $R=R\\left( \\theta \\right) $ and a sequence \n\\{z_{n}\\}\\subset \\mathbb{R}^{N}$ such tha\n\\begin{equation}\n\\int_{\\left[ B_{R}(z_{n})\\right] ^{c}}(|\\nabla\nu_{n}(x)|^{2}+u_{n}^{2}(x))dx<\\theta \\text{ uniformly for }n\\geq 1.\n\\label{18-4}\n\\end{equation\nDefine a new sequence of functions $v_{n}:=u_{n}(\\cdot +z_{n})\\in H^{1}\n\\mathbb{R}^{N}).$ Then we have $\\{v_{n}\\}\\subset \\mathbf{M}_{a}^{\\infty\n,(1)} $ and $J_{a}^{\\infty }(v_{n})=\\alpha _{a}^{\\infty ,-}+o(1).$ By virtue\nof $(\\ref{18-4})$, for each $\\theta >0$ there exists a constant $R=R(\\theta\n)>0$ such tha\n\\begin{equation*}\n\\int_{\\left[ B_{R}(z_{n})\\right] ^{c}}(|\\nabla\nv_{n}(x)|^{2}+v_{n}^{2}(x))dx<\\theta \\text{ uniformly for }n\\geq 1.\n\\end{equation*\nSince $\\{v_{n}\\}$ is bounded in $H^{1}(\\mathbb{R}^{N}),$ one can assume that\nthere exist a subsequence $\\{v_{n}\\}$ and $v_{a}^{-}\\in H^{1}(\\mathbb{R\n^{N}) $ such tha\n\\begin{eqnarray*}\nv_{n} &\\rightharpoonup &v_{a}^{-}\\text{ weakly in }H^{1}(\\mathbb{R}^{N}); \\\\\nv_{n} &\\rightarrow &v_{a}^{-}\\text{ strongly in }L_{loc}^{r}(\\mathbb{R}^{N}\n\\text{ for }2\\leq r<2^{\\ast }; \\\\\nv_{n} &\\rightarrow &v_{a}^{-}\\text{ a.e. in }\\mathbb{R}^{N}.\n\\end{eqnarray*\nIn the following, by adapting the argument of Theorem \\ref{t0-3} $(i),$ we\nobtain\n\\begin{equation*}\nv_{n}\\rightarrow v_{a}^{-}\\text{ strongly in }H^{1}(\\mathbb{R}^{N})\n\\end{equation*\nand\n\\begin{equation*}\nJ_{a}^{\\infty }(v_{n})\\rightarrow J_{a}^{\\infty }(v_{a}^{-})=\\alpha\n_{a}^{\\infty ,-}\\text{ as }n\\rightarrow \\infty .\n\\end{equation*\nThus, $v_{a}^{-}$ is a minimizer for $J_{a}^{\\infty }$ on $\\mathbf{M\n_{a}^{\\infty ,-}$ for each $0J_{a}^{\\infty }(T_{f_{\\infty\n}}(v_{a}^{-})v_{a}^{-})>J_{0}^{\\infty }(T_{f_{\\infty\n}}(v_{a}^{-})v_{a}^{-})\\geq \\alpha _{0}^{\\infty }.\n\\end{equation*}\n\n$(ii)$ Following the argument of Theorem 2.1 in \\cite{SZ}. By $(i),$ we\nobtain that Eq. $(E_{a})$ admits a positive solution $v_{a,1}^{-}\\in H^{1}\n\\mathbb{R}^{N}).$ Applying Theorem 4.1 in \\cite{HL1} gives \nv_{a,1}^{-}\\rightarrow 0$ as $|x|\\rightarrow \\infty .$ Then after\ntranslation, we can make $v_{a,1}^{-}$ satisf\n\\begin{equation}\nv_{a,1}^{-}>0,v_{a,1}^{-}(\\infty )=0,v_{a,1}^{-}(0)=\\max v_{a,1}^{-}(x).\n\\label{18-5}\n\\end{equation}\n\nNow we show that $v_{a,1}^{-}$ is unique under $(\\ref{18-5})$. Otherwise, we\nassume that $v_{a,2}^{-}$ is another positive solution satisfying $(\\re\n{18-5})$. Le\n\\begin{equation*}\nK_{1}=b+a\\int_{\\mathbb{R}^{N}}|\\nabla v_{a,1}^{-}|^{2}dx\\text{ and \nK_{2}=b+a\\int_{\\mathbb{R}^{N}}|\\nabla v_{a,2}^{-}|^{2}dx.\n\\end{equation*\nThen $v_{a,i}^{-}(i=1,2)$ is a solution of the proble\n\\begin{equation*}\n-\\Delta u+\\frac{1}{K_{i}}u=\\frac{1}{K_{i}}|u|^{p-2}u\\text{ in }\\mathbb{R\n^{N}.\n\\end{equation*\nLet $w_{i}(x)=v_{a,i}^{-}(\\sqrt{K_{i}}x).$ Then $w_{i}(x)$ is a solution o\n\\begin{equation}\n\\left\\{\n\\begin{array}{ll}\n-\\Delta w+w=|w|^{p-2}w\\text{ } & \\text{in }\\mathbb{R}^{N}, \\\\\nw>0,w(\\infty )=0,w(0)=\\max w(x). &\n\\end{array\n\\right. \\label{18-6}\n\\end{equation\nIt follows from \\cite{K} that the solution of problem $(\\ref{18-6})$ is\nunique. So $w_{1}(x)\\equiv w_{2}(x)$ i.e., $v_{a,1}^{-}(\\sqrt{K_{1}\nx)=v_{a,2}^{-}(\\sqrt{K_{2}}x).$ Thus, we have\n\\begin{equation}\nv_{a,2}^{-}(x)=v_{a,1}^{-}\\left( \\sqrt{\\frac{K_{1}}{K_{2}}}x\\right) .\n\\label{18-8}\n\\end{equation\nThus, we hav\n\\begin{eqnarray*}\nK_{2} &=&b+a\\int_{\\mathbb{R}^{N}}|\\nabla v_{a,2}^{-}|^{2}dx \\\\\n&=&b+a\\left( \\frac{K_{2}}{K_{1}}\\right) ^{\\frac{N-2}{2}}\\int_{\\mathbb{R\n^{N}}|\\nabla v_{a,1}^{-}|^{2}dx \\\\\n&=&b+\\left( \\frac{K_{2}}{K_{1}}\\right) ^{\\frac{N-2}{2}}(K_{1}-b),\n\\end{eqnarray*\nwhich implies tha\n\\begin{equation}\n\\frac{(K_{2}-b)^{2}}{K_{2}^{N-2}}=\\frac{(K_{1}-b)^{2}}{K_{1}^{N-2}}.\n\\label{18-7}\n\\end{equation\nDefin\n\\begin{equation*}\ny(x)=\\frac{(x-b)^{2}}{x^{N-2}}\\text{ for }x>b.\n\\end{equation*\nA direct calculation shows that\n\\begin{equation*}\ny^{\\prime }(x)=\\frac{\\left[ (4-N)x+b(N-2)\\right] (x-b)}{x^{N-1}}>0\\text{ for\n}x>b\\text{ and }1\\leq N\\leq 4,\n\\end{equation*}\nwhich implies that $y(x)$ is strictly increasing when $x>b$ and $1\\leq N\\leq\n4.$ This indicates that $K_{1}=K_{2},$since $K_{1},K_{2}>b.$ So by $(\\re\n{18-8})$ one have $v_{a,1}^{-}=v_{a,2}^{-}.$ Since the unique solution \nw_{1}(x)$ of problem $(\\ref{18-6})$ is radially symmetric by \\cite{K} and \nw_{1}(x)=v_{a,1}^{-}(\\sqrt{K_{1}}x)$, we obtain that $v_{a,1}^{-}$ is also\nradially symmetric.\n\n$\\left( iii\\right) $ Note that $\\Lambda \\leq \\overline{a}_{\\ast }$ for \nN\\geq 5.$ Then by virtue of Theorem \\ref{t0-3} $(i),$ for each $0\\sqrt{2}\\left( \\frac{2S_{p}^{p}}\nf_{\\infty }(4-p)}\\right) ^{1\/(p-2)}.\n\\end{equation*\nMoreover, from $(i)$ it follows that for each $0\\frac{p-2}{4p}\\left( \\frac{S_{p}^{p}}{f_{\\infty }\n\\right) ^{2\/(p-2)}\\text{ and }\\left\\Vert v_{a}^{-}\\right\\Vert\n_{H^{1}}<\\left( \\frac{2S_{p}^{p}}{f_{\\infty }(4-p)}\\right) ^{1\/(p-2)}.\n\\end{equation*\nConsequently, we complete the proof of Theorem \\ref{t1}.\n\n\\section{Proofs of Theorems \\protect\\ref{t2} and \\protect\\ref{t3}}\n\nBy virtue of Theorem \\ref{t1} $(i)$, we know that Eq. $(E_{a}^{\\infty })$\nadmits a positive solution $v_{a}^{-}\\in \\mathbf{M}_{a}^{\\infty ,-}$ such\nthat\n\\begin{equation*}\nJ_{a}^{\\infty }(v_{a}^{-})=\\alpha _{a}^{\\infty ,-}\\text{ and }(4-p)\\int_\n\\mathbb{R}^{N}}f_{\\infty }|v_{a}^{-}|^{p}dx<2\\Vert v_{a}^{-}\\Vert\n_{H^{1}}^{2}.\n\\end{equation*\nAccording to $(\\ref{2-0}),$ one has\n\\begin{equation}\nT_{f_{\\infty }}(v_{a}^{-})=\\left( \\frac{\\left\\Vert v_{a}^{-}\\right\\Vert\n_{H^{1}}^{2}}{\\int_{\\mathbb{R}^{N}}f_{\\infty }|v_{a}^{-}|^{p}dx}\\right)\n^{1\/(p-2)}>\\left( \\frac{4-p}{2}\\right) ^{1\/(p-2)}. \\label{5-1}\n\\end{equation\nMoreover, by Theorem \\ref{t1} $(ii)$, we obtain that Eq. $(E_{a}^{\\infty })$\nadmits a positive solution $v_{a}^{+}\\in \\mathbf{M}_{a}^{\\infty ,+}$ such\nthat\n\\begin{equation*}\nJ_{a}^{\\infty }(v_{a}^{+})=\\alpha _{a}^{\\infty ,+}\\ \\text{and }(4-p)\\int_\n\\mathbb{R}^{N}}f_{\\infty }|v_{a}^{+}|^{p}dx>2\\Vert v_{a}^{+}\\Vert\n_{H^{1}}^{2}.\n\\end{equation*\nSimilar to $(\\ref{5-1})$, we have\n\\begin{equation*}\nT_{f_{\\infty }}(v_{a}^{+})=\\left( \\frac{\\left\\Vert v_{a}^{+}\\right\\Vert\n_{H^{1}}^{2}}{\\int_{\\mathbb{R}^{N}}f_{\\infty }|v_{a}^{+}|^{p}dx}\\right)\n^{1\/(p-2)}<\\left( \\frac{4-p}{2}\\right) ^{1\/(p-2)}.\n\\end{equation*\nThen we have the following results.\n\n\\begin{lemma}\n\\label{m5}$(i)$ Suppose that $N\\geq 1.$ Then for each $00. \\label{5-3}\n\\end{equation\nClearly, there holds\n\\begin{equation}\nb_{a}^{\\infty }(1)+a\\left( \\int_{\\mathbb{R}^{N}}|\\nabla\nv_{a}^{-}|^{2}dx\\right) ^{2}=0 \\label{5-4}\n\\end{equation\nfor all $0\\left( \\frac{2}{4-p}\\right) ^{1\/(p-2)}T_{f_{\\infty }}(v_{a}^{-}).$ This\nindicates that\n\\begin{equation}\n\\inf_{t>0}b_{a}^{\\infty }(t)=b_{a}^{\\infty }\\left( \\left( \\frac{2}{4-p\n\\right) ^{1\/(p-2)}T_{f_{\\infty }}(v_{a}^{-})\\right) . \\label{5-2}\n\\end{equation\nMoreover, we notice that\n\\begin{equation}\n\\left( \\frac{2}{4-p}\\right) ^{1\/(p-2)}T_{f_{\\infty }}(v_{a}^{-})>1.\n\\label{5-5}\n\\end{equation\nThus, it follows from $(\\ref{5-4})-(\\ref{5-5})$ that\n\\begin{equation}\n\\inf_{t>0}b_{a}^{\\infty }(t)0. \\label{eqq41}\n\\end{equation\nApparently, $tv_{a}^{-}\\in \\mathbf{M}_{a}$ if and only if\n\\begin{equation*}\nb_{a}(t)+a\\left( \\int_{\\mathbb{R}^{N}}|\\nabla v_{a}^{-}|^{2}dx\\right) ^{2}=0.\n\\end{equation*\nBy analyzing $(\\ref{eqq41}),$ we obtain\n\\begin{equation*}\nb_{a}(T_{f}(v_{a}^{-}))=0,\\ \\lim_{t\\rightarrow 0^{+}}b_{a}(t)=\\infty \\text{\nand }\\lim_{t\\rightarrow \\infty }b_{a}(t)=0.\n\\end{equation*\nA direct calculation shows that\n\\begin{equation*}\nb_{a}^{\\prime }(t)=t^{-3}\\left( -2\\left\\Vert v_{a}^{-}\\right\\Vert\n_{H^{1}}^{2}+(4-p)t^{p-2}\\int_{\\mathbb{R}^{N}}f(x)|v_{a}^{-}|^{p}dx\\right) ,\n\\end{equation*\nwhich implies that $b_{a}(t)$ is decreasing on $0\\left( \\frac{2}\n4-p}\\right) ^{1\/(p-2)}T_{f}(v_{a}^{-}).$ By virtue of condition $(D5)$ one\nhas\n\\begin{equation*}\nT_{f}(v_{a}^{-})\\leq T_{f_{\\infty }}(v_{a}^{-})<1\\text{ and }b_{a}(t)\\leq\nb_{a}^{\\infty }(t),\n\\end{equation*\nwhere $b_{a}^{\\infty }(t)$ is given in $(\\ref{5-3}).$ Using condition $(D5)$\nand $(\\ref{5-6})$ gives\n\\begin{eqnarray*}\n\\inf_{t>0}b_{a}(t) &=&b_{a}\\left( \\left( \\frac{2}{4-p}\\right)\n^{1\/(p-2)}T_{f}(v_{a}^{-})\\right) \\\\\n&\\leq &-\\frac{p-2}{4-p}\\left( \\frac{4-p}{2}\\right) ^{2\/(p-2)}\\left\\Vert\nv_{a}^{-}\\right\\Vert _{H^{1}}^{2}\\left( \\frac{\\left\\Vert\nv_{a}^{-}\\right\\Vert _{H^{1}}^{2}}{\\int_{\\mathbb{R}^{N}}f_{\\infty\n}|v_{a}^{-}|^{p}dx}\\right) ^{-2\/(p-2)} \\\\\n&=&\\inf_{t>0}b_{a}^{\\infty }(t)<-a\\left( \\int_{\\mathbb{R}^{N}}|\\nabla\nv_{a}^{-}|^{2}dx\\right) ^{2}.\n\\end{eqnarray*\nThis explicitly tells us that there are two constants $t_{a}^{(1),-}$ and \nt_{a}^{(2),-}$ satisfyin\n\\begin{equation*}\nT_{f}(v_{a}^{-})0.\n\\end{eqnarray*\nSo, we get $t_{a}^{(1),-}v_{a}^{-}\\in \\mathbf{M}_{a}^{-}$ and \nt_{a}^{(2),-}v_{a}^{-}\\in \\mathbf{M}_{a}^{+}.$\n\nNote that\n\\begin{equation*}\nt_{a}^{(1),-}<\\left( \\frac{2}{4-p}\\right) ^{\\frac{1}{p-2}}T_{f}(v_{a}^{-}\n\\leq \\left( \\frac{2}{4-p}\\right) ^{\\frac{1}{p-2}}T_{f_{\\infty\n}}(v_{a}^{-})0$ for all $t\\in \\left(\n0,t_{a}^{(1),-}\\right) \\cup \\left( t_{a}^{(2),-},\\infty \\right) $ and \nh_{a,v_{a}^{-}}^{\\prime }(t)<0$ for all $t\\in \\left(\nt_{a}^{(1),-},t_{a}^{(2),-}\\right) $. Consequently, we arrive at\n\\begin{equation*}\nJ_{a}\\left( t_{a}^{(1),-}v_{a}^{-}\\right) =\\sup_{0\\leq t\\leq\nt_{a}^{(2),-}}J_{a}(tv_{a}^{-})\\text{ and }J_{a}\\left(\nt_{a}^{(2),-}v_{a}^{-}\\right) =\\inf_{t\\geq t_{a}^{\\left( 1\\right)\n,-}}J_{a}(tv_{a}^{-}).\n\\end{equation*\nThat is, $J_{a}\\left( t_{a}^{(2),-}v_{a}^{-}\\right) \\leq J_{a}\\left(\nt_{a}^{(1),-}v_{a}^{-}\\right) <\\alpha _{a}^{\\infty ,-}$, and so \nt_{a}^{(2),-}v_{a}^{-}\\in \\mathbf{M}_{a}^{(2),-}.$ Consequently, this\ncompletes the proof.\n\\end{proof}\n\n\\begin{lemma}\n\\label{m3-2}Suppose that $N\\geq 5$ and conditions ${(D1)}-{(D4)}$ hold. Then\nfor each $00$ and a differentiable\nfunction $t^{\\ast }:B_{\\sigma }(0)\\subset H^{1}(\\mathbb{R}^{N})\\rightarrow\n\\mathbb{R}^{+}$ such that $t^{\\ast }(0)=1\\ $and$\\ t^{\\ast }(v)(u-v)\\in\n\\mathbf{M}_{a}^{(j)}$ for all $v\\in B_{\\sigma }(0),$ and\n\\begin{eqnarray*}\n&&\\left\\langle (t^{\\ast })^{\\prime }(0),\\varphi \\right\\rangle \\\\\n&=&\\frac{2\\int_{\\mathbb{R}^{N}}(\\nabla u\\nabla \\varphi +u\\varphi\n)dx+4a\\left( \\int_{\\mathbb{R}^{N}}|\\nabla u|dx\\right) ^{2}\\int_{\\mathbb{R\n^{N}}\\nabla u\\nabla \\varphi dx-p\\int_{\\mathbb{R}^{N}}f(x)|u|^{p-2}u\\varphi d\n}{\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-(p-1)\\int_{\\mathbb{R\n^{N}}f(x)|u|^{p}dx}\n\\end{eqnarray*\nfor all $\\varphi \\in H^{1}(\\mathbb{R}^{N}).$\n\\end{lemma}\n\n\\begin{proof}\nFor any $u\\in \\mathbf{M}_{a}^{(j)}$, we define the function $F_{u}:\\mathbb{R\n\\times H^{1}(\\mathbb{R}^{N})\\rightarrow \\mathbb{R}$ b\n\\begin{eqnarray*}\nF_{u}(t,v) &=&\\left\\langle J_{a}^{\\prime }(t(u-v)),t(u-v)\\right\\rangle \\\\\n&=&t^{2}\\left\\Vert u-v\\right\\Vert _{H^{1}}^{2}+at^{4}\\left( \\int_{\\mathbb{R\n^{N}}|\\nabla u-\\nabla v|^{2}dx\\right) ^{2}-t^{p}\\int_{\\mathbb{R\n^{N}}f(x)|u-v|^{p}dx.\n\\end{eqnarray*\nClearly, $F_{u}(1,0)=\\left\\langle J_{a}^{\\prime }(u),u\\right\\rangle =0$ an\n\\begin{eqnarray*}\n\\frac{d}{dt}F_{u}(1,0) &=&2\\left\\Vert u\\right\\Vert _{H^{1}}^{2}+4a\\left(\n\\int_{\\mathbb{R}^{N}}|\\nabla u|^{2}dx\\right) ^{2}-p\\int_{\\mathbb{R\n^{N}}f(x)|u|^{p}dx \\\\\n&=&-2\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-\\left( p-4\\right) \\int_{\\mathbb{R\n^{N}}f(x)|u|^{p}dx \\\\\n&\\neq &0.\n\\end{eqnarray*\nApplying the implicit function theorem, there exist a constant $\\sigma >0$\nand a differentiable function $t^{\\ast }:B_{\\sigma }(0)\\subset H^{1}(\\mathbb\nR}^{N})\\rightarrow \\mathbb{R}$ such that $t^{\\ast }(0)=1$ and\n\\begin{eqnarray*}\n&&\\left\\langle (t^{\\ast })^{\\prime }(0),\\varphi \\right\\rangle \\\\\n&=&\\frac{2\\int_{\\mathbb{R}^{N}}(\\nabla u\\nabla \\varphi +u\\varphi\n)dx+4a\\left( \\int_{\\mathbb{R}^{N}}|\\nabla u|dx\\right) ^{2}\\int_{\\mathbb{R\n^{N}}\\nabla u\\nabla \\varphi dx-p\\int_{\\mathbb{R}^{N}}f(x)|u|^{p-2}u\\varphi d\n}{\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-(p-1)\\int_{\\mathbb{R\n^{N}}f(x)|u|^{p}dx}\n\\end{eqnarray*\nfor all $\\varphi \\in H^{1}(\\mathbb{R}^{N}),$ and $F_{u}(t^{\\ast }(v),v)=0$\nfor all $v\\in B_{\\sigma }(0),$ which is equivalent t\n\\begin{equation*}\n\\left\\langle J_{a}^{\\prime }(t^{\\ast }(v)(u-v)),t^{\\ast\n}(v)(u-v)\\right\\rangle =0\\text{ for all }v\\in B_{\\sigma }(0).\n\\end{equation*\nAccording to the continuity of the map $t^{\\ast }$, for $\\sigma $\nsufficiently small we have\n\\begin{eqnarray*}\nh_{a,t^{\\ast }(v)(u-v)}^{\\prime \\prime }(1) &=&-2\\left\\Vert t^{\\ast }\\left(\nv\\right) \\left( u-v\\right) \\right\\Vert _{H^{1}}^{2}-(p-4)\\int_{\\mathbb{R\n^{N}}f(x)|t^{\\ast }(v)(u-v)|^{p}dx \\\\\n&<&0,\n\\end{eqnarray*\nan\n\\begin{equation*}\nJ_{a}(t^{\\ast }(v)(u-v))<\\frac{(p-2)D(p)}{2p}\\left( \\frac{2S_{p}^{p}}\nf_{\\infty }(4-p)}\\right) ^{2\/(p-2)}.\n\\end{equation*\nHence, $t^{\\ast }(v)(u-v)\\in \\mathbf{M}_{a}^{(j)}$ for all $v\\in B_{\\sigma\n}(0).$ This completes the proof.\n\\end{proof}\n\nBy $(\\ref{4-4})$ and Lemma $\\ref{g7}$, we define\n\\begin{equation*}\n\\alpha _{a}^{-}=\\inf_{u\\in \\mathbf{M}_{a}^{(1)}}J_{a}(u)=\\inf_{u\\in \\mathbf{\n}_{a}^{-}}J_{a}(u)\\text{ for }N\\geq 1.\n\\end{equation*}\n\n\\begin{proposition}\n\\label{g8}Suppose that $N\\geq 1.$ Then for each $00$\nsuch that $t_{n}^{\\ast }(w)(u_{n}-w)\\in \\mathbf{M}_{a}^{(1)}.$ For $0<\\delta\n<\\epsilon _{n}$ and $u\\in H^{1}(\\mathbb{R}^{N})$ with $u\\not\\equiv 0,$ we\nset\n\\begin{equation*}\nw_{\\delta }=\\frac{\\delta u}{\\left\\Vert u\\right\\Vert _{H^{1}}}\\text{ and \nz_{\\delta }=t_{n}^{\\ast }(w_{\\delta })(u_{n}-w_{\\delta }).\n\\end{equation*\nSince $z_{\\delta }\\in \\mathbf{M}_{a}^{(1)},$ it follows from $(\\ref{22})$\ntha\n\\begin{equation*}\nJ_{a}(z_{\\delta })-J_{a}(u_{n})\\geq -\\frac{1}{n}\\Vert z_{\\delta }-u_{n}\\Vert\n_{H^{1}}.\n\\end{equation*\nUsing the mean value theorem gives\n\\begin{equation*}\n\\left\\langle J_{a}^{\\prime }(u_{n}),z_{\\delta }-u_{n}\\right\\rangle +o\\left(\n\\left\\Vert z_{\\delta }-u_{n}\\right\\Vert _{H^{1}}\\right) \\geq -\\frac{1}{n\n\\Vert z_{\\delta }-u_{n}\\Vert _{H^{1}}\n\\end{equation*\nand\n\\begin{eqnarray}\n&&\\left\\langle J_{a}^{\\prime }(u_{n}),-w_{\\delta }\\right\\rangle\n+(t_{n}^{\\ast }(w_{\\delta })-1)\\left\\langle J_{a}^{\\prime\n}(u_{n}),u_{n}-w_{\\delta }\\right\\rangle \\notag \\\\\n&\\geq &-\\frac{1}{n}\\Vert z_{\\delta }-u_{n}\\Vert _{H^{1}}+o(\\Vert z_{\\delta\n}-u_{n}\\Vert _{H^{1}}). \\label{23}\n\\end{eqnarray\nNote that $t_{n}^{\\ast }(w_{\\delta })(u_{n}-w_{\\delta })\\in \\mathbf{M\n_{a}^{(1)}$. From $(\\ref{23})$ it leads to\n\\begin{eqnarray*}\n&&-\\delta \\left\\langle J_{a}^{\\prime }(u_{n}),\\,\\frac{u}{\\left\\Vert\nu\\right\\Vert _{H^{1}}}\\right\\rangle +\\frac{(t_{n}^{\\ast }(w_{\\delta })-1)}\nt_{n}^{\\ast }(w_{\\delta })}\\left\\langle J_{a}^{\\prime }(z_{\\delta\n}),t_{n}^{\\ast }(w_{\\delta })(u_{n}-w_{\\delta })\\right\\rangle \\\\\n&&+(t_{n}^{\\ast }(w_{\\delta })-1)\\left\\langle J_{a}^{\\prime\n}(u_{n})-J_{a}^{\\prime }(z_{\\delta }),u_{n}-w_{\\delta }\\right\\rangle \\\\\n&\\geq &-\\frac{1}{n}\\left\\Vert z_{\\delta }-u_{n}\\right\\Vert\n_{H^{1}}+o(\\left\\Vert z_{\\delta }-u_{n}\\right\\Vert _{H^{1}}).\n\\end{eqnarray*\nWe rewrite the above inequality as\n\\begin{eqnarray}\n\\left\\langle J_{a}^{\\prime }(u_{n}),\\frac{u}{\\left\\Vert u\\right\\Vert _{H^{1}\n}\\right\\rangle &\\leq &\\frac{\\left\\Vert z_{\\delta }-u_{n}\\right\\Vert _{H^{1}\n}{\\delta n}+\\frac{o(\\left\\Vert z_{\\delta }-u_{n}\\right\\Vert _{H^{1}})}\n\\delta } \\notag \\\\\n&&+\\frac{(t_{n}^{\\ast }(w_{\\delta })-1)}{\\delta }\\langle J_{a}^{\\prime\n}(u_{n})-J_{a}^{\\prime }(z_{\\delta }),u_{n}-w_{\\delta }\\rangle . \\label{24}\n\\end{eqnarray\nThere exists a constant $C>0$ independent of $\\delta $ such tha\n\\begin{equation*}\n\\left\\Vert z_{\\delta }-u_{n}\\right\\Vert _{H^{1}}\\leq \\delta +C(\\left\\vert\nt_{n}^{\\ast }\\left( w_{\\delta }\\right) -1\\right\\vert )\n\\end{equation*\nan\n\\begin{equation*}\n\\lim_{\\delta \\rightarrow 0}\\frac{\\left\\vert t_{n}^{\\ast }(w_{\\delta\n})-1\\right\\vert }{\\delta }\\leq \\left\\Vert (t_{n}^{\\ast })^{\\prime\n}(0)\\right\\Vert \\leq C.\n\\end{equation*\nLetting $\\delta \\rightarrow 0$ in $(\\ref{24})$ and using the fact that \n\\lim_{\\delta \\rightarrow 0}\\left\\Vert z_{\\delta }-u_{n}\\right\\Vert\n_{H^{1}}=0,$ we ge\n\\begin{equation*}\n\\left\\langle J_{a}^{\\prime }(u_{n}),\\frac{u}{\\left\\Vert u\\right\\Vert _{H^{1}\n}\\right\\rangle \\leq \\frac{C}{n},\n\\end{equation*\nwhich leads to $(\\ref{eqq20}).$\n\\end{proof}\n\nBefore proving Theorem \\ref{t2}, we also need the following compactness\nlemma which is an immediate conclusion of Proposition \\ref{g10}.\n\n\\begin{lemma}\n\\label{lem2} Suppose that $N\\geq 1$ and conditions ${(D1)}-{(D2),(D4)}-{(D5)}\n$ hold. Let $\\{u_{n}\\}\\subset \\mathbf{M}_{a}^{(1)}$ be a $(PS)_{\\beta }\n--sequence in $H^{1}(\\mathbb{R}^{N})$ for $J_{a}$ with $0<\\beta <\\alpha\n_{a}^{\\infty ,-}.$ Then there exist a subsequence $\\{u_{n}\\}$ and a nonzero \nu_{0}$ in $H^{1}(\\mathbb{R}^{N})$ such that $u_{n}\\rightarrow u_{0}$\nstrongly in $H^{1}(\\mathbb{R}^{N})$ and $J_{a}(u_{0})=\\beta .$ Furthermore, \nu_{0}$ is a nonzero solution of Eq. $(E_{a}).$\n\\end{lemma}\n\n\\textbf{We are now ready to prove Theorems \\ref{t2} and \\ref{t3}:} By\nProposition \\ref{g8}, there exists a sequence $\\{u_{n}\\}\\subset \\mathbf{M\n_{a}^{(1)}$ satisfying\n\\begin{equation*}\nJ_{a}(u_{n})=\\alpha _{a}^{-}+o(1)\\text{ and }J_{a}^{\\prime }(u_{n})=o(1\n\\text{ in }H^{-1}(\\mathbb{R}^{N}).\n\\end{equation*\nIt follows from Lemmas \\ref{lem2} and \\ref{m3} that Eq. $(E_{a})$ has a\nnontrivial solution $u_{a}^{-}\\in \\mathbf{M}_{a}^{-}$ such that \nJ_{a}(u_{a}^{-})=\\alpha _{a}^{-}.$ Thus, $u_{a}^{-}$ is a minimizer for \nJ_{a}$ on $\\mathbf{M}_{a}^{-}.$ since\n\\begin{equation*}\n\\alpha _{a}^{-}<\\alpha _{a}^{\\infty ,-}<\\frac{p-2}{2p}D(p)\\left( \\frac\n2S_{p}^{p}}{f_{\\infty }(4-p)}\\right) ^{2\/(p-2)},\n\\end{equation*\none has $u_{a}^{-}\\in \\mathbf{M}_{a}^{(1)}$. Similarly, we obtain \n|u_{a}^{-}|\\in \\mathbf{M}_{a}^{-}$ and $J_{a}(|u_{a}^{-}|)=J_{a}(u_{a}^{-})\n\\alpha _{a}^{-}$. According to Lemma \\ref{g2}, $u_{a}^{-}$ is a positive\nsolution of Eq. $(E_{a})$ when $N\\geq 1.$ Consequently, the proof of Theorem\n\\ref{t2} is complete.\n\nNote that $\\Lambda \\leq \\overline{a}_{\\ast }$ for $N\\geq 5.$ Then it follows\nfrom Theorem \\ref{t0-3}$(ii)$ that for each $0\\sqrt{2}\\left( \\frac{2S_{p}^{p}}\nf_{\\max }(4-p)}\\right) ^{1\/(p-2)}.\n\\end{equation*\nMoreover, by Theorem \\ref{t2} we obtain that for each $0\\frac{p-2}{4p}\\left( \\frac{S_{p}^{p}}{f_{\\max }}\\right)\n^{2\/(p-2)}\\text{ and }\\left\\Vert u_{a}^{-}\\right\\Vert _{H^{1}}<\\left( \\frac\n2S_{p}^{p}}{f_{\\max }(4-p)}\\right) ^{1\/(p-2)}.\n\\end{equation*\nConsequently, we complete the proof of Theorem \\ref{t3}.\n\n\\section{Ground State Solutions}\n\n\\begin{lemma}\n\\label{g14}Suppose that $N=1$ and $f(x)\\in C(\\mathbb{R})$ is weakly\ndifferentiable satisfying\n\\begin{equation*}\n(p-1)(p-2)f\\left( x\\right) +2\\langle \\nabla f(x),x\\rangle \\geq 0.\n\\end{equation*\nLet $u_{0}$ be a nontrivial solution of Eq. $(E_{a}).$ Then $u_{0}\\in\n\\mathbf{M}_{a}^{-}.$\n\\end{lemma}\n\n\\begin{proof}\nSince $u_{0}$ is a nontrivial solution of Eq. $(E_{a})$, we have\n\\begin{equation}\n\\left\\Vert u_{0}\\right\\Vert _{H^{1}}^{2}+a\\left( \\int_{\\mathbb{R}}|\\nabla\nu_{0}|^{2}dx\\right) ^{2}-\\int_{\\mathbb{R}}f(x)|u_{0}|^{p}dx=0. \\label{3-1}\n\\end{equation\nFollowing the argument of \\cite[Lemma 3.1]{DM1}, $u_{0}$ satisfies the\nPohozaev type identity corresponding to Eq. $(E_{a})$ as follows\n\\begin{equation}\n\\frac{1}{2}\\int_{\\mathbb{R}}(-|\\nabla u_{0}|^{2}+u_{0}^{2})dx-\\frac{a}{2\n\\left( \\int_{\\mathbb{R}}|\\nabla u_{0}|^{2}dx\\right) ^{2}=\\frac{1}{p}\\int_\n\\mathbb{R}}(f(x)+\\langle \\nabla f(x),x\\rangle )|u_{0}|^{p}dx. \\label{3-4}\n\\end{equation\nThen it follows from $(\\ref{3-1})-(\\ref{3-4})$ and the assumption of $f(x)$\ntha\n\\begin{eqnarray*}\nh_{a,u_{0}}^{\\prime \\prime }(1) &=&-2\\int_{\\mathbb{R}}|\\nabla u_{0}|^{2}dx\n\\frac{(p+2)}{p}\\int_{\\mathbb{R}}f(x)|u_{0}|^{p}dx+\\frac{2}{p}\\int_{\\mathbb{R\n}\\langle \\nabla f(x),x\\rangle |u_{0}|^{p}dx \\\\\n&&-\\frac{4}{p}\\int_{\\mathbb{R}}f(x)|u_{0}|^{p}dx-\\frac{4}{p}\\int_{\\mathbb{R\n}\\langle \\nabla f(x),x\\rangle |u_{0}|^{p}dx-(p-2)\\int_{\\mathbb{R\n}f(x)|u_{0}|^{p}dx \\\\\n&=&-2\\int_{\\mathbb{R}}|\\nabla u_{0}|^{2}dx-\\frac{1}{p}\\int_{\\mathbb{R}^{N}\n\\left[ (p-1)(p-2)f(x)+2\\langle \\nabla f(x),x\\rangle \\right] |u_{0}|^{p}dx \\\\\n&<&0,\n\\end{eqnarray*\nwhich shows that $u_{0}\\in \\mathbf{M}_{a}^{-}.$ This completes the proof.\n\\end{proof}\n\n\\begin{lemma}\n\\label{g13}Suppose that $N=2$ and $f(x)\\in C(\\mathbb{R}^{2})$ is weakly\ndifferentiable satisfying\n\\begin{equation*}\n(p-2)f(x)+\\langle \\nabla f(x),x\\rangle \\geq 0.\n\\end{equation*}\nLet $u_{0}$ be a nontrivial solution of Eq. $(E_{a}).$ Then $u_{0}\\in\n\\mathbf{M}_{a}^{-}.$\n\\end{lemma}\n\n\\begin{proof}\nSince $u_{0}$ is a nontrivial solution of Eq. $(E_{a})$, there holds\n\\begin{equation}\n\\left\\Vert u_{0}\\right\\Vert _{H^{1}}^{2}+a\\left( \\int_{\\mathbb{R\n^{2}}|\\nabla u_{0}|^{2}dx\\right) ^{2}-\\int_{\\mathbb{R\n^{2}}f(x)|u_{0}|^{p}dx=0. \\label{3-3}\n\\end{equation\nMoreover, $u_{0}$ satisfies the Pohozaev type identity corresponding to Eq. \n(E_{a})$ as follows\n\\begin{equation}\n\\frac{2}{p}\\int_{\\mathbb{R}^{2}}u_{0}^{2}dx-\\frac{1}{2}\\int_{\\mathbb{R\n^{2}}\\langle \\nabla f(x),x\\rangle |u_{0}|^{p}dx=\\int_{\\mathbb{R}^{2}}f\\left(\nx\\right) |u_{0}|^{p}dx. \\notag\n\\end{equation\nUsing the above two equalities give\n\\begin{equation}\n\\frac{2}{p}\\int_{\\mathbb{R}^{2}}u_{0}^{2}dx-\\frac{1}{2}\\int_{\\mathbb{R\n^{2}}\\langle \\nabla f(x),x\\rangle |u_{0}|^{p}dx=\\left\\Vert u_{0}\\right\\Vert\n_{H^{1}}^{2}+a\\left( \\int_{\\mathbb{R}^{2}}|\\nabla u_{0}|^{2}dx\\right) ^{2}.\n\\label{3-6}\n\\end{equation\nThen it follows from $(\\ref{3-3})-(\\ref{3-6})$ and the assumption of $f(x)$\ntha\n\\begin{eqnarray*}\nh_{a,u_{0}}^{\\prime \\prime }(1) &=&2a\\left( \\int_{\\mathbb{R}^{2}}|\\nabla\nu_{0}|^{2}dx\\right) ^{2}-(p-2)\\int_{\\mathbb{R}^{2}}f\\left( x\\right)\n|u_{0}|^{p}dx \\\\\n&=&-2\\int_{\\mathbb{R}^{2}}|\\nabla u_{0}|^{2}dx-\\frac{2(p-2)}{p}\\int_{\\mathbb\nR}^{2}}u_{0}^{2}dx-\\int_{\\mathbb{R}^{2}}\\langle \\nabla f(x),x\\rangle\n|u_{0}|^{p}dx \\\\\n&&-(p-2)\\int_{\\mathbb{R}^{2}}f\\left( x\\right) |u_{0}|^{p}dx \\\\\n&=&-2\\int_{\\mathbb{R}^{2}}|\\nabla u_{0}|^{2}dx-\\frac{2(p-2)}{p}\\int_{\\mathbb\nR}^{2}}u_{0}^{2}dx-\\int_{\\mathbb{R}^{2}}\\left[ (p-2)f(x)+\\langle \\nabla\nf(x),x\\rangle \\right] |u_{0}|^{p}dx \\\\\n&<&0,\n\\end{eqnarray*\nwhich shows that $u_{0}\\in \\mathbf{M}_{a}^{-}.$ This completes the proof.\n\\end{proof}\n\n\\textbf{We are now ready to prove Theorem \\ref{t4}:} Let $u^{-}$ be the\npositive solution of Eq. $(E_{a})$ as described in Theorem \\ref{t1}$(i)$ or\n\\ref{t2}$(i)$. Then there holds $u^{-}\\in \\mathbf{M}_{a}^{-}$ and \nJ_{a}(u^{-})=\\inf_{u\\in \\mathbf{M}_{a}^{-}}J_{a}(u)=\\alpha _{a}^{-}.$ By\nLemma \\ref{g14} or \\ref{g13}, $u^{-}$ is a positive ground state solution of\nEq. $(E_{a}).$\n\n\\begin{lemma}\n\\label{g3}Suppose that $N=3$ and $f(x)\\equiv f_{\\infty }.$ Let $u_{0}$ be a\nnontrivial solution of Eq. $(E_{a}).$ Then for each $a>0$ with $\\sqrt{a^{2}+\n}+\\frac{2}{a}\\geq A_{0},$ there holds $u_{0}\\in \\mathbf{M}_{a}^{-},$ where \nA_{0}>0$ is defined as $(\\ref{1-3}).$\n\\end{lemma}\n\n\\begin{proof}\nLet $u_{0}$ be a nontrivial solution of Eq. $(E_{a})$. Then there holds\n\\begin{equation}\n\\left\\Vert u_{0}\\right\\Vert _{H^{1}}^{2}+a\\left( \\int_{\\mathbb{R\n^{3}}|\\nabla u_{0}|^{2}dx\\right) ^{2}=f_{\\infty }\\int_{\\mathbb{R\n^{3}}|u_{0}|^{p}dx. \\label{6-9}\n\\end{equation\nMoreover, $u_{0}$ satisfies the Pohozaev type identity corresponding to Eq. \n(E_{a})$ as follow\n\\begin{equation}\n\\frac{p}{6}\\int_{\\mathbb{R}^{3}}(|\\nabla u_{0}|^{2}+3u_{0}^{2})dx+\\frac{ap}{\n}\\left( \\int_{\\mathbb{R}^{3}}|\\nabla u_{0}|^{2}dx\\right) ^{2}=f_{\\infty\n}\\int_{\\mathbb{R}^{3}}|u_{0}|^{p}dx. \\label{6-10}\n\\end{equation\nBy Azzollini \\cite[Theorem 1.1]{Az2}, for each $a>0$ there exists a constan\n\\begin{equation*}\nt_{a}=\\frac{-a+\\sqrt{a^{2}+4}}{2}>0\n\\end{equation*\nsuch that $u_{0}(\\cdot )=w_{0}(t_{a}\\cdot ).$ Then it follows from $\\left(\n\\ref{1-8}\\right) ,(\\ref{6-9})-(\\ref{6-10})$ tha\n\\begin{eqnarray*}\nh_{a,u_{0}}^{\\prime \\prime }(1) &=&2a\\left( \\int_{\\mathbb{R}^{3}}|\\nabla\nu_{0}|^{2}dx\\right) ^{2}-\\left( p-2\\right) f_{\\infty }\\int_{\\mathbb{R\n^{3}}|u_{0}|^{p}dx \\\\\n&=&\\frac{\\int_{\\mathbb{R}^{3}}|\\nabla u_{0}|^{2}dx}{6}\\left[\na(-p^{2}+2p+12)\\int_{\\mathbb{R}^{3}}|\\nabla u_{0}|^{2}dx-p(p-2)\\right] \n\\frac{p(p-2)}{2}\\int_{\\mathbb{R}^{3}}u_{0}^{2}dx \\\\\n&=&\\frac{\\int_{\\mathbb{R}^{3}}|\\nabla u_{0}|^{2}dx}{6}\\left[\na(-p^{2}+2p+12)t_{a}^{-2}\\int_{\\mathbb{R}^{3}}|\\nabla w_{0}|^{2}dx-p(p-2\n\\right] -\\frac{p(p-2)}{2}\\int_{\\mathbb{R}^{3}}u_{0}^{2}dx \\\\\n&=&\\frac{\\int_{\\mathbb{R}^{3}}|\\nabla u_{0}|^{2}dx}{6}\\left[ \\frac\n2a(-p^{2}+2p+12)\\int_{\\mathbb{R}^{3}}|\\nabla w_{0}|^{2}dx}{2+a\\sqrt{a^{2}+4}\n-p(p-2)\\right] -\\frac{p(p-2)}{2}\\int_{\\mathbb{R}^{3}}u_{0}^{2}dx. \\\\\n&=&\\frac{\\int_{\\mathbb{R}^{3}}|\\nabla u_{0}|^{2}dx}{6}\\left[ \\frac{3\\left(\np-1\\right) (-p^{2}+2p+12)}{p\\left( \\frac{2}{a}+\\sqrt{a^{2}+4}\\right) }\\left(\n\\frac{S_{p}^{p}}{f_{\\infty }}\\right) ^{\\frac{2}{p-2}}-p(p-2)\\right] -\\frac\np(p-2)}{2}\\int_{\\mathbb{R}^{3}}u_{0}^{2}dx\n\\end{eqnarray*\nThis implies tha\n\\begin{equation*}\nf_{\\infty }\\geq \\left[ \\frac{3\\left( p-1\\right) (-p^{2}+2p+12)}\np^{2}(p-2)\\inf_{a>0}\\left( \\sqrt{a^{2}+4}+\\frac{2}{a}\\right) }\\right]\n^{2\/\\left( p-2\\right) }S_{p}^{p}\n\\end{equation*\n\\begin{equation*}\nh_{a,u_{0}}^{\\prime \\prime }(1)<0\\text{ if }\\sqrt{a^{2}+4}+\\frac{2}{a}\\geq\n\\inf_{a>0}\\sqrt{a^{2}+4}+\\frac{2}{a}\\geq A_{0}=\\frac{3\\left( p-1\\right)\n(-p^{2}+2p+12)}{p^{2}(p-2)}\\left( \\frac{S_{p}^{p}}{f_{\\infty }}\\right) ^\n\\frac{2}{p-2}},\n\\end{equation*\nand so for each $a>0$ with $\\sqrt{a^{2}+4}+\\frac{2}{a}\\geq A_{0},$ there\nholds $u_{0}\\in \\mathbf{M}_{a}^{-}$. This completes the proof.\n\\end{proof}\n\n\\begin{lemma}\n\\label{g5}Suppose that $N=4$ and $f(x)\\equiv f_{\\infty }$. Let $u_{0}$ be a\nnontrivial solution of Eq. $(E_{a}).$ Then for each $00$\nis defined as $(\\ref{1-4}).$\n\\end{lemma}\n\n\\begin{proof}\nLet $u_{0}$ be a nontrivial solution of Eq. $(E_{a})$. Then we have\n\\begin{equation}\n\\left\\Vert u_{0}\\right\\Vert _{H^{1}}^{2}+a\\left( \\int_{\\mathbb{R\n^{4}}|\\nabla u_{0}|^{2}dx\\right) ^{2}=f_{\\infty }\\int_{\\mathbb{R\n^{4}}|u_{0}|^{p}dx. \\label{6-1}\n\\end{equation\nMoreover, $u_{0}$ satisfies the Pohozaev type identity corresponding to Eq. \n(E_{a})$ as follow\n\\begin{equation}\n\\int_{\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx+a\\left( \\int_{\\mathbb{R\n^{4}}|\\nabla u_{0}|^{2}dx\\right) ^{2}+2\\int_{\\mathbb{R}^{4}}u_{0}^{2}dx\n\\frac{4}{p}f_{\\infty }\\int_{\\mathbb{R}^{4}}|u_{0}|^{p}dx \\label{6-2}\n\\end{equation\nBy Azzollini \\cite[Theorem 1.1]{Az2}, for each $00\n\\end{equation*\nsuch that $u_{0}(\\cdot )=w_{0}(t_{a}\\cdot ).$ Then it follows from $(\\re\n{6-1})-(\\ref{6-2})$ tha\n\\begin{eqnarray*}\nh_{a,u_{0}}^{\\prime \\prime }(1) &=&2a\\left( \\int_{\\mathbb{R}^{4}}|\\nabla\nu_{0}|^{2}dx\\right) ^{2}-\\left( p-2\\right) f_{\\infty }\\int_{\\mathbb{R\n^{4}}|u_{0}|^{p}dx \\\\\n&=&\\frac{\\int_{\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx}{4}\\left[\na(4-p)(p+2)\\int_{\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx-p(p-2)\\right] -\\frac\np(p-2)}{4}\\int_{\\mathbb{R}^{4}}u_{0}^{2}dx \\\\\n&=&\\frac{\\int_{\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx}{4}\\left[\na(4-p)(p+2)t_{a}^{-2}\\int_{\\mathbb{R}^{4}}|\\nabla w_{0}|^{2}dx-p(p-2)\\right]\n-\\frac{p(p-2)}{2}\\int_{\\mathbb{R}^{4}}u_{0}^{2}dx \\\\\n&=&\\frac{\\int_{\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx}{4}\\left[ \\frac\na(4-p)(p+2)\\int_{\\mathbb{R}^{4}}|\\nabla w_{0}|^{2}dx}{1-a\\int_{\\mathbb{R\n^{4}}|\\nabla w_{0}|^{2}dx}-p(p-2)\\right] -\\frac{p(p-2)}{2}\\int_{\\mathbb{R\n^{4}}u_{0}^{2}dx \\\\\n&<&0\\text{ for all }0A_{0}^{\\ast },$ there\nholds $u_{0}\\in \\mathbf{M}_{a}^{+},$ where $A_{0}^{\\ast }>0$ is defined as $\n\\ref{1-5}).$\n\\end{lemma}\n\n\\begin{proof}\nLet $u_{0}$ be a nontrivial solution of Eq. $(E_{a}).$ By $(\\ref{6-1})-(\\re\n{6-2}),$ we hav\n\\begin{equation}\n-(p-2)\\left\\Vert u_{0}\\right\\Vert _{H^{1}}^{2}+a(4-p)\\left( \\int_{\\mathbb{R\n^{4}}|\\nabla u_{0}|^{2}dx\\right) ^{2}=p\\int_{\\mathbb{R}^{4}}u_{0}^{2}dx-\n\\left\\Vert u_{0}\\right\\Vert _{H^{1}}^{2}. \\label{D-1}\n\\end{equation\nApplying the Gagliardo-Nirenberg and Young inequalities leads t\n\\begin{eqnarray*}\n\\left\\Vert u_{0}\\right\\Vert _{H^{1}}^{2}+a\\left( \\int_{\\mathbb{R\n^{4}}|\\nabla u_{0}|^{2}dx\\right) ^{2} &\\leq &f_{\\infty }C^{p}\\left( \\int_\n\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx\\right) ^{p-2}\\left( \\int_{\\mathbb{R\n^{4}}u_{0}^{2}dx\\right) ^{\\left( 4-p\\right) \/2} \\\\\n&\\leq &\\frac{p-2}{2}\\left( \\frac{4-p}{p}\\right) ^{\\left( 4-p\\right) \/\\left(\np-2\\right) }\\left( f_{\\infty }C^{p}\\right) ^{\\frac{2}{p-2}}\\left( \\int_\n\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx\\right) ^{2} \\\\\n&&+\\frac{p}{2}\\int_{\\mathbb{R}^{4}}u_{0}^{2}dx,\n\\end{eqnarray*\nwhich implies tha\n\\begin{equation}\np\\int_{\\mathbb{R}^{4}}u_{0}^{2}dx-2\\left\\Vert u_{0}\\right\\Vert\n_{H^{1}}^{2}\\geq 2\\left( a-\\frac{p-2}{2}\\left( \\frac{4-p}{p}\\right) ^{\\left(\n4-p\\right) \/\\left( p-2\\right) }(f_{\\infty }C^{p})^{2\/\\left( p-2\\right)\n}\\right) \\left( \\int_{\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx\\right) ^{2}.\n\\label{D-0}\n\\end{equation\nThus, by $(\\ref{2-2})$ and $(\\ref{D-1})-(\\ref{D-0}),$ for each \na>A_{0}^{\\ast }$ one ha\n\\begin{eqnarray*}\nh_{a,u_{0}}^{\\prime \\prime }(1) &=&-(p-2)\\left\\Vert u_{0}\\right\\Vert\n_{H^{1}}^{2}+a(4-p)\\left( \\int_{\\mathbb{R}^{4}}|\\nabla u_{0}|^{2}dx\\right)\n^{2} \\\\\n&=&p\\int_{\\mathbb{R}^{4}}u_{0}^{2}dx-2\\left\\Vert u_{0}\\right\\Vert\n_{H^{1}}^{2} \\\\\n&\\geq &2\\left( a-\\frac{p-2}{2}\\left( \\frac{4-p}{p}\\right) ^{\\frac{4-p}{p-2\n}(f_{\\infty }C^{p})^{\\frac{2}{p-2}}\\right) \\left( \\int_{\\mathbb{R\n^{4}}|\\nabla u_{0}|^{2}dx\\right) ^{2} \\\\\n&>&0,\n\\end{eqnarray*\nwhich indicates that $u_{0}\\in \\mathbf{M}_{a}^{+}.$ This complete the proof.\n\\end{proof}\n\n\\textbf{We are now ready to prove Theorem \\ref{t5}:} By Lemmas \\ref{g3}--\\re\n{g16}, we can arrive at the conclusions directly.\n\n\\section{Appendix}\n\nIn order to verify $(\\ref{18-4}),$ we use the concentration-compactness\nlemma \\cite{Li1,Li2}. First of all, we discuss the three possibilities on\nthe measures defined by a functional related to $J_{a}^{\\infty }.$\n\nFor $20.\n\\label{7-1}\n\\end{equation\nDefine the functional $\\Phi _{a}^{\\infty }:H^{1}(\\mathbb{R}^{N})\\rightarrow\n\\mathbb{R}$ b\n\\begin{equation*}\n\\Phi _{a}^{\\infty }(u)=\\frac{1}{4}\\left\\Vert u\\right\\Vert _{H^{1}}^{2}-\\frac\n4-p}{4p}\\int_{\\mathbb{R}^{N}}f_{\\infty }|u|^{p}dx.\n\\end{equation*\nBy Lemma \\ref{g1}, for any $u\\in \\mathbf{M}_{a}^{\\infty ,(1)}$ one ha\n\\begin{equation*}\nJ_{a}^{\\infty }(u)=\\Phi _{a}^{\\infty }(u)>0.\n\\end{equation*\nNote that $\\{u_{n}\\}$ is bounded in $H^{1}(\\mathbb{R}^{N}),$ since \n\\{u_{n}\\}\\subset \\mathbf{M}_{a}^{\\infty ,(1)}.$ Then there exist a\nsubsequence $\\{u_{n}\\}$ and $u_{\\infty }\\in H^{1}(\\mathbb{R}^{N})$ such tha\n\\begin{eqnarray*}\nu_{n} &\\rightharpoonup &u_{\\infty }\\text{ weakly in }H^{1}(\\mathbb{R}^{N}),\n\\\\\nu_{n} &\\rightarrow &u_{\\infty }\\text{ strongly in }L_{loc}^{s}(\\mathbb{R\n^{N})\\text{ for }2\\leq s<2^{\\ast }.\n\\end{eqnarray*\nFor any $u_{n}\\in \\mathbf{M}_{a}^{\\infty ,(1)}$, we define the measure \ny_{n}(\\Omega )$ b\n\\begin{equation}\ny_{n}(\\Omega )=\\frac{1}{4}\\int_{\\Omega }(|\\nabla u_{n}|^{2}+u_{n}^{2})dx\n\\frac{4-p}{4p}\\int_{\\Omega }f_{\\infty }|u_{n}|^{p}dx. \\label{7-4}\n\\end{equation\nSince $u_{n}\\in \\mathbf{M}_{a}^{\\infty ,\\left( 1\\right) },$ we have \n\\left\\Vert u_{n}\\right\\Vert _{H^{1}}0,\n\\begin{equation}\n\\lim_{n\\rightarrow \\infty }\\sup_{\\xi \\in \\mathbb{R}^{N}}\\int_{B_{r}(\\xi\n)}dy_{n}=0, \\label{7-5}\n\\end{equation\nwhere $B_{r}(\\xi )=\\{x\\in \\mathbb{R}^{N}:|x-\\xi |0,$ there exists \nr=r(\\delta )>0$ such tha\n\\begin{equation}\n\\int_{B_{r}(\\xi _{n})}dy_{n}\\geq \\alpha _{a}^{\\infty ,-}-\\delta ,\\text{ for\nlarge }n. \\label{7-8}\n\\end{equation}\n\n\\begin{theorem}\n\\label{g10} For $00,$ $\n\\ref{7-5})$ holds. In particular, we deduce that there exists $\\bar{r}>0$\nsuch tha\n\\begin{equation*}\n\\lim_{n\\rightarrow \\infty }\\sup_{\\xi \\in \\mathbb{R}^{N}}\\int_{B_{\\bar{r\n}(\\xi )}u_{n}^{2}dx=0,\n\\end{equation*\nwhich implies that $u_{n}\\rightarrow 0$ strongly in $L^{s}(\\mathbb{R}^{N})$\nfor $20.\n\\end{equation*\nWe now prove that $t_{a,n}^{-}\\leq 1.$ Suppose the contrary. Then \nt_{a,n}^{-}>1.$ Since $t_{a,n}^{-}h_{n}\\in \\mathbf{M}_{a}^{\\infty ,-},$ we\nhav\n\\begin{equation*}\na\\left( \\int_{\\mathbb{R}^{N}}|\\nabla h_{n}|^{2}dx\\right)\n^{2}=-(t_{a,n}^{-})^{-2}\\left\\Vert h_{n}\\right\\Vert\n_{H^{1}}^{2}+(t_{a,n}^{-})^{p-4}\\int_{\\mathbb{R}^{N}}f_{\\infty\n}|h_{n}|^{p}dx.\n\\end{equation*\nUsing the above equality give\n\\begin{eqnarray}\n0 &\\geq &\\left\\langle (J_{a}^{\\infty })^{\\prime }(h_{n}),h_{n}\\right\\rangle\n=\\left\\Vert h_{n}\\right\\Vert _{H^{1}}^{2}+a\\left( \\int_{\\mathbb{R\n^{N}}|\\nabla h_{n}|^{2}dx\\right) ^{2}-\\int_{\\mathbb{R}^{N}}f_{\\infty\n}|h_{n}|^{p}dx \\notag \\\\\n&=&\\left\\Vert h_{n}\\right\\Vert _{H^{1}}^{2}-(t_{a,n}^{-})^{-2}\\left\\Vert\nh_{n}\\right\\Vert _{H^{1}}^{2}+(t_{a,n}^{-})^{p-4}\\int_{\\mathbb{R\n^{N}}f_{\\infty }|h_{n}|^{p}dx \\notag \\\\\n&&-\\int_{\\mathbb{R}^{N}}f_{\\infty }|h_{n}|^{p}dx \\notag \\\\\n&=&[1-(t_{a,n}^{-})^{-2}]\\left\\Vert h_{n}\\right\\Vert\n_{H^{1}}^{2}+[(t_{a,n}^{-})^{p-4}-1]\\int_{\\mathbb{R}^{N}}f_{\\infty\n}|h_{n}|^{p}dx. \\label{7-16}\n\\end{eqnarray\nNote that\n\\begin{equation}\n\\int_{\\mathbb{R}^{N}}f_{\\infty }|h_{n}|^{p}dx<\\frac{2}{4-p\n(t_{a,n}^{-})^{2-p}\\left\\Vert h_{n}\\right\\Vert _{H^{1}}^{2} \\label{7-15}\n\\end{equation\nby $(\\ref{2-2}).$ It follows from $(\\ref{7-16})-(\\ref{7-15})$ tha\n\\begin{eqnarray}\n0 &\\geq &[1-(t_{a,n}^{-})^{-2}]\\left\\Vert h_{n}\\right\\Vert _{H^{1}}^{2}\n\\frac{2}{4-p}[(t_{a,n}^{-})^{-2}-(t_{a,n}^{-})^{2-p}]\\left\\Vert\nh_{n}\\right\\Vert _{H^{1}}^{2} \\notag \\\\\n&=&\\left[ 1+\\frac{p-2}{4-p}(t_{a,n}^{-})^{-2}-\\frac{2}{4-p\n(t_{a,n}^{-})^{2-p}\\right] \\left\\Vert h_{n}\\right\\Vert _{H^{1}}^{2} \\notag\n\\\\\n&=&(t_{a,n}^{-})^{-2}\\left[ (t_{a,n}^{-})^{2}-\\frac{2}{4-p\n(t_{a,n}^{-})^{4-p}+\\frac{p-2}{4-p}\\right] \\left\\Vert h_{n}\\right\\Vert\n_{H^{1}}^{2}, \\label{7-21}\n\\end{eqnarray\nwhich implies tha\n\\begin{equation*}\n(t_{a,n}^{-})^{2}-\\frac{2}{4-p}(t_{a,n}^{-})^{4-p}+\\frac{p-2}{4-p}\\leq 0.\n\\end{equation*\nHowever, we observe that for $20\\text{ for }t>1.\n\\end{equation*\nThis is a contradiction. Thus, $t_{a,n}^{-}\\leq 1.$\n\nNext, let us consider the functional $\\Phi _{a}^{\\infty }(th_{n})$ defined b\n\\begin{equation*}\n\\Phi _{a}^{\\infty }(th_{n})=\\frac{t^{2}}{4}\\left\\Vert h_{n}\\right\\Vert\n_{H^{1}}^{2}-\\frac{(4-p)t^{p}}{4p}\\int_{\\mathbb{R}^{N}}f_{\\infty\n}|h_{n}|^{p}dx\\text{ for }t>0.\n\\end{equation*\nA direct calculation shows that there exists a constan\n\\begin{equation*}\nt_{a}^{\\infty }(h_{n})=\\left( \\frac{2\\left\\Vert h_{n}\\right\\Vert _{H^{1}}^{2\n}{(4-p)f_{\\infty }\\int_{\\mathbb{R}^{N}}|h_{n}|^{p}dx}\\right) ^{1\/(p-2)}>0\n\\end{equation*\nsuch that $\\Phi _{a}^{\\infty }(th_{n})$ is increasing on $(0,t_{a}^{\\infty\n}(h_{n}))$ and is decreasing on $(t_{a}^{\\infty }(h_{n}),\\infty ).$\nFurthermore, by the Sobolev inequality and $(\\ref{7-17})$ we hav\n\\begin{equation*}\nt_{a}^{\\infty }(h_{n})\\geq \\left( \\frac{2}{(4-p)f_{\\infty\n}S_{p}^{-p}\\left\\Vert h_{n}\\right\\Vert _{H^{1}}^{p-2}}\\right) ^{1\/(p-2)}>1.\n\\end{equation*\nThis indicates that $\\Phi _{a}^{\\infty }(t_{a,n}^{-}h_{n})\\leq \\Phi\n_{a}^{\\infty }(h_{n}).$ Hence, for all $n\\geq 1,$ there hold\n\\begin{equation*}\n\\alpha _{a}^{\\infty ,-}\\leq J_{a}^{\\infty }(t_{a,n}^{-}h_{n})=\\Phi\n_{a}^{\\infty }(t_{a,n}^{-}h_{n})\\leq \\Phi _{a}^{\\infty }(h_{n})\\rightarrow\n\\alpha <\\alpha _{a}^{\\infty ,-},\n\\end{equation*\nwhich is a contradiction.\\newline\nCase $(ii):$ Up to a subsequence, $\\left\\langle (J_{a}^{\\infty })^{\\prime\n}(h_{n}),h_{n}\\right\\rangle >0$ and $\\left\\langle (J_{a}^{\\infty })^{\\prime\n}(w_{n}),w_{n}\\right\\rangle >0.$\n\nBy $(\\ref{7-18})$ one has $\\left\\langle (J_{a}^{\\infty })^{\\prime\n}(h_{n}),h_{n}\\right\\rangle =o_{n}(1)$ and $\\left\\langle (J_{a}^{\\infty\n})^{\\prime }(w_{n}),w_{n}\\right\\rangle =o_{n}(1).$ If $t_{a,n}^{-}\\leq\n1+o_{n}(1),$ then we can repeat the argument of Case $(i)$ and arrive at the\ncontradiction. Suppose that\n\\begin{equation*}\n\\lim_{n\\rightarrow \\infty }t_{a,n}^{-}=t_{a,\\infty }^{-}>1.\n\\end{equation*\nSimilar to the argument of $(\\ref{7-16})$, we hav\n\\begin{eqnarray*}\no_{n}(1) &=&\\left\\langle (J_{a}^{\\infty })^{\\prime\n}(h_{n}),h_{n}\\right\\rangle \\\\\n&=&[1-(t_{a,n}^{-})^{-2}]\\left\\Vert h_{n}\\right\\Vert\n_{H^{1}}^{2}+[(t_{a,n}^{-})^{p-4}-1]\\int_{\\mathbb{R}^{N}}f_{\\infty\n}|h_{n}|^{p}dx.\n\\end{eqnarray*\nSimilar to the argument of $(\\ref{7-21}),$ one ge\n\\begin{equation}\no_{n}(1)\\geq (t_{a,n}^{-})^{-2}\\left[ (t_{a,n}^{-})^{2}-\\frac{2}{4-p\n(t_{a,n}^{-})^{4-p}+\\frac{p-2}{4-p}\\right] \\left\\Vert h_{n}\\right\\Vert\n_{H^{1}}^{2}, \\notag\n\\end{equation\nwhich shows that\n\\begin{equation*}\n\\left\\Vert h_{n}\\right\\Vert _{H^{1}}^{2}\\rightarrow 0\\text{ as }n\\rightarrow\n\\infty ,\n\\end{equation*\nwhere we have used the fact of $(t_{a,n}^{-})^{2}-\\frac{2}{4-p\n(t_{a,n}^{-})^{4-p}+\\frac{p-2}{4-p}>0.$ The\n\\begin{equation*}\n\\int_{\\mathbb{R}^{N}}|h_{n}|^{p}dx\\rightarrow 0\\text{ as }n\\rightarrow\n\\infty .\n\\end{equation*\nHence, $\\Phi _{a}^{\\infty }(h_{n})\\rightarrow 0$ as $n\\rightarrow \\infty ,$\nwhich contradicts $(\\ref{7-24})$. Therefore, the dichotomy cannot occur.\n\\end{proof}\n\n\\section*{Acknowledgments}\n\nJ. Sun is supported by the National Natural Science Foundation of China\n(Grant No. 11671236). T.F. Wu is supported in part by the Ministry of\nScience and Technology, Taiwan (Grant 108-2115-M-390-007-MY2), the\nMathematics Research Promotion Center, Taiwan and the National Center for\nTheoretical Sciences, Taiwan.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\\label{sec:intro}\nA wireless relay network is one in which a set of\nrelay nodes assist a source node transmit information to a\ndestination node. Practically the wireless nodes can only support\nhalf-duplex communication \\cite{Laneman04}, i.e., no nodes can\nreceive and transmit information simultaneously on the same\nfrequency band. Different cooperative transmission schemes for\nsystems with half-duplex nodes have been proposed in the literature.\nFundamentally, these schemes consist of two basic steps. First, the\nsource transmits to the destination, and the relay listens and\n``captures'' \\cite{Host05} the transmission from the source at the\nsame time. Next, the relays send processed source information to the\ndestination while the source may still transmit to the destination\ndirectly. Variants of these techniques have been proposed and have\nbeen shown to yield good performance under different circumstances\n\\cite{Avestimehr06,Azarian05,Laneman04}.\n\nAssuming channel state information (CSI) at the nodes, an\nopportunistic decode-and-forward (DF) protocol for half-duplex relay\nchannels is proposed in \\cite{Gunduz07}.\nIn \\cite{Ong07}, the authors present\nrouting algorithms to optimize the rate from a source to a\ndestination, based on the DF technique that uses regular block\nMarkov encoding and windowed decoding \\cite{Xie04,Kramer03}, for the\nGaussian full-duplex multiple-relay channel. The achievable rate of\n\\cite{Xie04} for the Gaussian physically degraded full-duplex\nmulti-relay channel has been established as the capacity of this\nchannel in \\cite{Reznik04}. In \\cite{Ong06}, it is shown that the\ncut-set bound on the capacity of the Gaussian single source-multiple\nrelay-single destination mesh network can be achieved using the\ncompress-and-forward (CF) method, as the relay powers go to\ninfinity.\n\nSome simpler cooperative diversity methods based on network path\nselection have been recently reported \\cite{Bletsas06,Beres06}.\nThese selection methods include: (i) the max-min selection method\n\\cite{Bletsas06}, wherein the relay node with the maximum of the\nminimum of the source-relay and relay-destination channel gains is\nselected; (ii) the harmonic mean selection method \\cite{Bletsas06},\nwherein the relay node with the highest harmonic mean of the\nsource-relay and relay-destination channel gains is selected; and\n(iii) the selection scheme of \\cite{Beres06}, in which the relay\nthat can correctly decode the information from the source and has\nthe best relay-destination channel is selected. These methods\nachieve a DMT of $d(r)=(N-1)(1-2r)$ for\nan $N$ node relay network and multiplexing gain $00.5$).\n\nWe have proposed a cooperative diversity design based on a flow\noptimization approach for a three-node network in \\cite{Wong06}. In\nthis design, the source node broadcasts two distinct flows to the\ndestination and the relay node respectively during the relay's\nlisten period. Then the relay forwards this information using the\nDF approach while the source may also send another flow of\ninformation to the destination during the relay's transmit period. \nThis scheme is shown to achieve the\noptimal diversity order for the three-node relay channel and yield\nperformance very close to optimal full-duplex relaying in both\nlow- and high-rate situations.\n\nHere, we apply this cooperative transmission design to a general\nrelay network, wherein wireless links are present between each\npair of nodes in the network.\nAs in \\cite{Wong06}, assuming CSI is available at all nodes we use\nbroadcasting (BC), multiple access (MA) and time sharing (TS)\ntechniques to formulate a flow theoretic convex optimization problem\nbased on the channel conditions. Instead of considering a total\npower constraint for all the transmitting nodes as in~\\cite{Wong06},\nwe subject each node to a maximum transmit power constraint. This\nyields a more reasonable system model for a general wireless\nrelay network, especially when the number of nodes in the relay network is\nlarge. The resulting relaying protocol will be referred to as the\nflow-optimized (FO) protocol. To obtain a more practical cooperative\ndesign we develop a generalized-link selection (GLS) protocol, in\nwhich we select the best relay node out of the available ones to\nform an equivalent three-node relay network to transmit the information\nfrom the source to the destination. The benefit of this, over other\nnetwork path selection strategies, becomes evident when the rate\nrequirement is high. It is shown that the simple GLS protocol is\noptimal in terms of the DMT\n\\cite{Zheng03} and yields acceptable performance even when the rate\nrequirement is high.\n\nRecently, in~\\cite{Yuksel07}, the authors have shown that compress-and-forward (CF) relaying\nachieves the optimal DMT for the three-node, half-duplex network, and that DF relaying\ncan achieve the optimal DMT of the four-node full-duplex network.\nIn this work, we show that the optimal DMT can be achieved for a general $N$-node ($N\\geq3$)\nhalf-duplex network using the FO or GLS protocols. Here, it should be clarified\nthat we consider that the wireless\nlinks between each node-pair experience\nindependent Rayleigh fading, and this corresponds to the definition\nof \\emph{non-clustered} networks in~\\cite{Yuksel07}.\nThe performances of the FO and GLS protocols are\nevaluated numerically in terms of their outage probabilities for\nfour- and five-node relay networks for uniform and non-uniform\naverage power gains. The numerical results motivate the use of the\nGLS protocol for situations where computation complexity is an issue\nand show a remarkable improvement over the max-min selection method\nof \\cite{Bletsas06}. The proposed designs, based on BC and MA alone,\nare sub-optimal in general. For a fair appraisal of the proposed\nprotocols, we compare the proposed protocols to an upper bound on\nthe maximum rate, derived using the max-flow-min-cut theorem\n\\cite[Thm.~14.10.1]{Cover91}.\n\n\\section{A General Design Using A Flow-theoretic Approach}\n\\label{sec:gdesgn}\n\nWe consider an $N$-node wireless relay network with a link joining each\npair of nodes. Each such wireless link is described by a bandpass\nGaussian channel with bandwidth $W$ and one-sided noise spectral\ndensity $N_0$. We denote the power gain of the link from node $i$ to\nnode $j$ as $Z_{ij}$. The link power gains are assumed to be\nindependent and identically distributed (i.i.d.) exponential random\nvariables with unit mean. This corresponds to the case of\nindependent Rayleigh fading channels with unit average power gains.\nMoreover, we assume that each node has a maximum power limit of $P$\nand can only support half-duplex transmission. Note that this model\ncan be easily generalized to the case where channels may have\nnon-uniform average power gains (for which numerical examples are\npresented in Section~\\ref{sec:num_ex}), and where different nodes\nmay have different maximum power constraints. More specifically, the\nlatter case can be converted into the uniform maximum power\nconstraints case by absorbing the non-uniformity in the transmit\npowers into the average power gains of the corresponding links. In\nthe sequel, we characterize the system in terms of the transmit\nsignal-to-noise ratio (SNR), $S = \\frac{P}{N_0W}$, at the input of\nthe links. Time is divided into unit intervals, and BC and MA are\napplied with a TS strategy that is optimized to maximize the\nspectral efficiency (which we call ``rate'' hereafter for\nconvenience). To avoid interference between concurrent\ntransmissions, a time interval is divided into slots:\n\n\\begin{itemize}\n\\item During the first slot, the source may BC to all the other\n nodes in the network\n\n\\item During the subsequent slots, a relay may BC to all other nodes\n (except the source node), or it may receive flows from all other\n nodes (except from the destination) through MA.\n\n\\item During the very last slot, the source and the relays may send\n information flows to the destination using MA.\n\n\n\\end{itemize}\nNote that the forwarding of information by the relays is based on the DF approach.\nFor practicality consideration, it is assumed that the phases of the\nsimultaneously transmitted signals from different nodes are not\nsynchronized. In general, for the above transmission protocol, there would be a maximum of $2(N-2)+ 2 = 2N-2$ time slots of lengths $t_1, t_2, \\cdots, t_{2N-2}$ respectively.\n\nNext, we describe the optimization problem using a graph-theoretic\nformulation. Define a graph $G=(V,E)$, where $V$ is the set of\nnodes, $E$ is the set of all links joining the nodes in the graph,\nand associate the vector $\\underbar{r}$ to represent the flow rates\nassociated with each link in $E$. Thus, the number of elements in\n$\\underbar{r}$ equals the cardinality of $E$. For convenience, we\nwrite $G=(V,E,\\underbar{r})$. Now denote the source by\n$\\mathcal{S}$, the destination by $\\mathcal{D}$, and the relay nodes\nby $\\mathcal{R}_1, \\ldots, \\mathcal{R}_{N-2}$. The slotting of a\nunit time interval, as described above, yields simpler graphs for\neach time slot, that we call \\emph{basic graphs}. A basic graph is\neither one in which a particular node may BC to several nodes, or in\nwhich several nodes transmit via MA to a particular node. Thus for a\nbasic graph, we need to include only the links between the nodes\nthat may participate during the concerned time slot. For example,\nassume that the relay $\\mathcal{R}_1$ broadcasts to all nodes other than the\nsource, during the $i$-th time slot. The basic graph is given by\n$G_i=(V,E_i,\\underbar{r}_i)$ where $V=\\{\\mathcal{S}, \\mathcal{R}_1,\n\\cdots, \\mathcal{R}_{N-2}, \\mathcal{D}\\},~\nE_i=\\{\\mathcal{R}_1\\mathcal{R}_2, \\cdots,\n\\mathcal{R}_1\\mathcal{R}_{N-2}, \\mathcal{R}_1\\mathcal{D}\\},~\n\\underbar{r}_i =\n\\left(\\frac{x_{\\mathcal{R}_1\\mathcal{R}_2}^{i}}{t_i}~ \\cdots\n~\\frac{x_{\\mathcal{R}_1\\mathcal{R}_{N-2}}^{i}}{t_i}\n~\\frac{x_{\\mathcal{R}_1\\mathcal{D}}^{i}}{t_i}\\right)^T$, where\n$x_{AB}^{i}$ is the flow from node $A$ to node $B$ during the $i$-th\ntime slot.\n\nIn general, the proposed design involves TS between the basic graphs\nto yield the following equivalent graph $G$ corresponding to a unit\ninterval (see \\cite{Wu05} for a similar idea):\n\\begin{equation}\nG = \\left( V, ~\\bigcup_i E_i, ~\\sum_i\nt_i\\underbar{r}_i\\right)\n= t_1 G_1 + t_2 G_2 + \\ldots + t_{2N-2} G_{2N-2}.\n\\label{eqn:G_equiv}\n\\end{equation}\n\\noindent where the number of elements in each vector\n$\\underbar{r}_i$ is extended to $\\vert\\bigcup_i E_i\\vert$ by\ninserting zeros appropriately. The second equality in\n(\\ref{eqn:G_equiv}) implies that $G$ can be viewed as a linear\ncombination of the basic graphs $G_i$s, with the equivalent set of\nedges given by the union of the sets $E_i$, and the equivalent flow\nrate vector given by the linear combination of the individual flow\nrate vectors $\\underbar{r}_i$.\n\nTo maximize the data rate from the source to the destination through\nthe relay network, we need to consider each cut that partitions $V$\ninto sets $V^s$ and $V^d$ with $\\mathcal{S}\\in V^s$ and\n$\\mathcal{D}\\in V^d$.\nClearly, there can be $2^{N-2}$ such possible cuts for the $N$-node\nrelay network. Let these cuts and the corresponding cut sets be denoted by\n$\\mathcal{C}_k$, $V^s_k$, and $V^d_k$, respectively, for\n$k=1,2,\\cdots,2^{N-2}$. Further, for the graph $G$, for any two\nnodes $A\\in V^s_k$ and $B\\in V^d_k$, there exists a \\emph{cut edge}\n$AB$ that crosses the cut. Denote the total flow through cut edge $AB$ in a\nunit time interval by $x_{AB}=\\sum_{i=1}^{2N-2}x_{AB}^{i}$. Now\nrecall from network flow theory~\\cite{Ahuja93} that the maximum flow\nrate from the source to the destination is specified by the minimal\ncut of the equivalent graph (\\ref{eqn:G_equiv}). Consequently, we\narrive at the following convex flow optimization problem that can be\nsolved using standard optimization techniques:\n\\begin{equation}\n \\max\\min \\left(\\sum_{A\\in V^s_1,B\\in V^d_1}x_{AB}, \\sum_{A\\in\n V^s_2,B\\in V^d_2}x_{AB}, \\cdots, \\sum_{A\\in V^s_{2^{N-2}},B\\in\n V^d_{2^{N-2}}}x_{AB} \\right)\\label{eqn:main_opt_prob1}\n\\end{equation}\n\\noindent over all flow allocations $x_{AB}^{i}$ and all time slot\nlengths $t_i$, subject to\n\\begin{itemize}\n\\item the \\emph{non-negativity constraints}: $x_{AB}^{i}$, $t_i \\geq\n 0$ for all cut edges $AB$ and $i=1,2,\\cdots,2N-2$,\n\\item the \\emph{total-time constraint}: $t_1+\\ldots+t_{2N-2} =1$,\n\\item the \\emph{power (capacity) constraints}:\n\\begin{itemize}\n\\item for a BC slot the flow rates should lie in the capacity region\n of the BC channel with the transmitting node having a power\n constraint of $P$,\n\\item for an MA slot the flow rates should lie in the capacity region\n of the MA channel with a maximum power constraint $P$ for each\n transmitting node,\n\\end{itemize}\n\\item the \\emph{flow constraints}: considering steady state operation,\n the total information flow out of a relay should equal the flow into\n the relay in each unit time interval.\n\\end{itemize}\n\n\\noindent Note that the dependence of the objective function on the channel gains and the\ntime slot lengths is implicitly expressed through the capacity constraints.\nDenote the cut separating $\\mathcal{S}$ from all the other nodes and\nthe cut separating $\\mathcal{D}$ from all nodes as\n$\\mathcal{C}_{\\mathcal{S}}$ and $\\mathcal{C}_{\\mathcal{D}}$,\nrespectively. Then we observe that the cost function in\n(\\ref{eqn:main_opt_prob1}) above can be further simplified\nto $\\max~\\min\\left\\{x(\\mathcal{C}_{\\mathcal{S}}),~x(\\mathcal{C}_{\\mathcal{D}})\\right\\}$,\nwhere\n\\begin{equation}\n\\displaystyle x(\\mathcal{C}_{\\mathcal{S}}) = x_{\\mathcal{SD}} +\n\\sum_{j=1}^{N-2}x_{\\mathcal{S}\\mathcal{R}_j}\\mathrm{~~~ and ~~~}\nx(\\mathcal{C}_{\\mathcal{D}}) = x_{\\mathcal{SD}} +\n\\sum_{i=1}^{N-2}x_{\\mathcal{R}_i\\mathcal{D}}\\label{eqn:BC_cut}\n\\end{equation}\nare the total flows across the above-mentioned cuts\n$\\mathcal{C}_{\\mathcal{S}}$ and $\\mathcal{C}_{\\mathcal{D}}$,\nrespectively. To see this, consider the cut $\\mathcal{C}$ with\n$V^s=\\{\\mathcal{S},\\mathcal{R}_1,\\cdots,\\mathcal{R}_l\\}$, and\n$V^d=\\{\\mathcal{R}_{l+1},\\cdots,\\mathcal{R}_{N-2},\\mathcal{D}\\}$ for\nsome $l\\in\\{1,2,\\cdots,N-2\\}$. The total flow across this cut is\ngiven by\n\\begin{equation}\n\\displaystyle x(\\mathcal{C}) = x_{\\mathcal{SD}} +\n\\sum_{j=l+1}^{N-2}x_{\\mathcal{S}\\mathcal{R}_j} +\n\\sum_{i=1}^l\\left(x_{\\mathcal{R}_i\\mathcal{D}} +\n\\sum_{j=l+1}^{N-2}x_{\\mathcal{R}_i\\mathcal{R}_j}\\right).\\label{eqn:flow_thru_C}\n\\end{equation}\n\n\\noindent Now, consider node $i$ for $i\\in\\{1,2,\\cdots,l\\}$.\nAccording to the flow constraint for node $i$,\n\\begin{equation}\n\\displaystyle x_{\\mathcal{R}_i\\mathcal{D}} +\n\\sum_{j=l+1}^{N-2}x_{\\mathcal{R}_i\\mathcal{R}_j} + \\sum_{k=1,k\\neq\ni}^{l}x_{\\mathcal{R}_i\\mathcal{R}_k} = x_{\\mathcal{S}\\mathcal{R}_i}\n+ \\sum_{j=l+1}^{N-2}x_{\\mathcal{R}_j\\mathcal{R}_i} + \\sum_{k=1,k\\neq\ni}^{l}x_{\\mathcal{R}_k\\mathcal{R}_i}.\\label{eqn:node_i_flow}\n\\end{equation}\nSumming (\\ref{eqn:node_i_flow}) over all $i\\in\\{1,2,\\cdots,l\\}$ we get\n\\begin{equation}\n\\sum_{i=1}^l\\left(x_{\\mathcal{R}_i\\mathcal{D}}\n+ \\sum_{j=l+1}^{N-2}x_{\\mathcal{R}_i\\mathcal{R}_j}\\right)=\n\\sum_{i=1}^l\\left(x_{\\mathcal{S}\\mathcal{R}_i} +\n\\sum_{j=l+1}^{N-2}x_{\\mathcal{R}_j\\mathcal{R}_i}\\right).\\label{eqn:nodes_i2l_flow}\n\\end{equation}\nSince\n$\\sum_{i=1}^l\\sum_{j=l+1}^{N-2}x_{\\mathcal{R}_j\\mathcal{R}_i}\\geq 0$,\ncombining (\\ref{eqn:BC_cut}), (\\ref{eqn:flow_thru_C}) and\n(\\ref{eqn:nodes_i2l_flow}) gives $x(\\mathcal{C}) \\geq\nx(\\mathcal{C}_{\\mathcal{S}})$. Similarly, we have\n$x(\\mathcal{C}) \\geq x(\\mathcal{C}_{\\mathcal{D}})$. Thus the cost\nfunction in (\\ref{eqn:main_opt_prob1}) reduces to the above-mentioned form.\n\n\\section{Generalized-link Selection and Its Optimality}\\label{sec:optimality}\n\nIn this section, we present the GLS protocol and establish the\noptimality of the FO and GLS protocols in terms of the\nDMT. This is accomplished in three\nsteps. First, we apply the FO protocol to the three-node relay network.\nNext, we propose the GLS protocol based on a selection strategy that\nis sub-optimal to the FO protocol of Section~\\ref{sec:gdesgn}.\nFinally, the optimality of the GLS protocol, and thereby, that of the FO\nprotocol, is established.\n\n\\subsection{The Three-node Relay Network} \\label{subsec:3node}\n\nThe three-node relay network consists of a source ($\\mathcal{S}$), a relay\n($\\mathcal{R}$), and a destination ($\\mathcal{D}$). We specialize\nthe general design described in the previous section to this\nthree-node relay network. A unit time interval is divided into two time\nslots of lengths $t_1$ and $t_2$ with $t_1 + t_2 = 1$.\nDuring the first time slot, $\\mathcal{S}$ sends (via BC) two flows\nof rates $x_{\\mathcal{SD}}^{1}\/t_1=x_1\/t_1$ and\n$x_{\\mathcal{SR}}^{1}\/t_1=x_2\/t_1$ to $\\mathcal{D}$ and\n$\\mathcal{R}$, respectively, resulting in the basic graph $G_1$.\nDuring the second time slot, $\\mathcal{R}$ and $\\mathcal{S}$ send\n(via MA) two flows of rates $x_{\\mathcal{RD}}^{2}\/t_2=x_4\/t_2$ and\n$x_{\\mathcal{SD}}^{2}\/t_2=x_3\/t_2$ to $\\mathcal{D}$, respectively,\nresulting in the basic graph $G_2$.\nCombining the two basic graphs yields the\nequivalent graph as $G = t_1G_1 + t_2G_2$. Note that the information\nflow of rate $x_4\/t_2$ sent by $\\mathcal{R}$ during the MA time slot\nis from the flow of rate $x_2\/t_1$ it received during the BC time\nslot.\nThus, we have the flow constraint $x_4=x_2$. The rate for this\nnetwork is specified by the min-cut which is clearly $\\min\\{(x_1 +\nx_2 + x_3),(x_1+x_4+x_3)\\}$. Hence, the flow optimization problem is\ngiven by:\n\\begin{equation}\n\\max~\\min~\\{(x_1 + x_2 + x_3),(x_1+x_4+x_3)\\} \\label{eqn:3node_opt}\n\\end{equation}\nover flow allocations $x_1,~x_2,~x_3,~x_4$, and time slot lengths\n$t_1,~t_2$, subject to\n\\begin{itemize}\n\\item {\\it non-negativity constraints: } $x_1, x_2, x_3, x_4 \\geq\n 0,\\quad t_1, t_2 \\geq 0$,\n\\item {\\it total-time constraint: } $t_1+t_2=1$,\n\\item {\\it power constraints: } $S_{BC}\\leq S,~x_1\\leq t_1C(Z_{\\mathcal{SD}}S),~x_2\\leq t_1C(Z_{\\mathcal{SR}}S)$ for the BC slot,\\\\\n \n $\\qquad x_3\\leq t_2C(Z_{\\mathcal{SD}}S),~ x_4\\leq\n t_2C(Z_{\\mathcal{RD}}S),~ x_3 + x_4\\leq\n t_2C(Z_{\\mathcal{SD}}S+Z_{\\mathcal{RD}}S)$ for the MA slot,\n\\item {\\it flow constraint: } $x_2 = x_4$,\n\\end{itemize}\nwhere $C(x)=\\log(1+x)$, and $S_{BC}$, the minimum SNR required for\nthe source to broadcast at rates $x_1\/t_1$ and $x_2\/t_1$ to the\ndestination and the relay, respectively, in the first time slot with\n$0 Z_{\\mathcal{S}\\mathcal{D}}, \\\\\n\\frac{1}{Z_{\\mathcal{S}\\mathcal{R}}} ( e^{x_2\/t_1} -1) +\n\\frac{1}{Z_{\\mathcal{S}\\mathcal{D}}} e^{x_2\/t_1} ( e^ {x_1\/t_1} -1)\n\\mathrm{~ for ~} Z_{\\mathcal{S}\\mathcal{R}} \\leq\nZ_{\\mathcal{S}\\mathcal{D}}.\n\\end{array}\n\\right .\n\\]\nFor $t_1=0$, $S_{BC}=0$. Note that for the BC slot, the last two constraints are redundant when $t_1>0$, and complements the first constraint when $t_1=0$.\n\nThe solution of this flow optimization problem is given in\nAppendix~\\ref{app:opt_soln}. As mentioned in Section~\\ref{sec:intro},\nthe above optimization problem formulation is different from that\nin~\\cite{Wong06} wherein the sum of the source and relay powers,\nrequired to achieve a certain data rate, is minimized. More\nspecifically, when considering individual power constraints for each\nnode, we cannot use part $2$ of~\\cite[Lemma 3.1]{Wong06} to describe\nthe power constraints for the MA slot. This is because doing so would\nrestrict the flows $x_2$ and $x_3$ such that the sum of powers\nexpended at $\\mathcal{S}$ and $\\mathcal{R}$ is minimized. On the\nother hand, in the present problem, the power constraints only dictate\nthat the flow-rates should lie in the MA capacity region specified by\nthe maximum power available at each transmitting node, for the\nparticular fading state. With this modification in the constraint for\nthe MA slot, the solution approach to the above problem needs to be\nmarkedly different from that in~\\cite{Wong06} as shown in Appendix~\\ref{app:opt_soln}.\nThe maximum information rate from the source $\\mathcal{S}$ to the\ndestination $\\mathcal{D}$ for different cases is summarized below:\n\n\\noindent a) $Z_{\\mathcal{SD}}\\geq Z_{\\mathcal{SR}}$: The maximum\nrate is $X(S) = C(Z_{\\mathcal{SD}}S)$ with direct transmission\nfrom $\\mathcal{S}$ to $\\mathcal{D}$.\n\n\\noindent b) $Z_{\\mathcal{SD}}Z_{\\mathcal{SD}}$: Let the set of all\n such node indices be $K$ and for all $i\\in I\\setminus K$,\n $Z_{\\mathcal{SD}}\\geq Z_{\\mathcal{S}\\mathcal{R}_i}$. For this case,\n choose the node $\\mathcal{R}_k'$ as the relay such that\n $k'=\\arg\\max_{k\\in K} X_k(S)$, where $X_k(S)$ is the maximum rate\n for the three-node relay network with the source $\\mathcal{S}$, the relay\n $\\mathcal{R}_k$ and destination $\\mathcal{D}$.\n\\end{enumerate}\nIn terms of the worst-case computational complexities for the FO and GLS\nprotocols, it can be seen that, for an $N$-node relay network with $N>3$, the FO protocol\ninvolves a max-min optimization over $2(N^2-2N+2)$ variables (all possible flows and time slot\nlengths), subject to $N-1$ non-linear and $2(N^2-N+1)$ linear constraints, whereas\nthe GLS protocol involves a maximum of $N-2$ maximizations of a non-linear concave function over two variables,\nsubject to two linear constraints, followed by finding the maximum of $N-2$ real numbers with\na worst-case complexity of $O(N-2)$. Moreover, for $N>3$, for the FO protocol, the BC slots\npotentially involve $(N-1)$- and $(N-2)$-level superposition coding (SPC) or dirty paper coding (DPC)\nimplementations for $\\mathcal{S}$ and the relays respectively, while the MA slots at\nthe relays and $\\mathcal{D}$ may involve a maximum of $(N-3)$ and $(N-2)$\ninterference cancelation (IC) operations respectively. On the other hand, the GLS protocol\ninvolves a maximum of $2$-level SPC\/DPC and one IC operation for the BC and MA slots respectively,\nfor any $N>3$.\n\n\\subsection{Diversity-multiplexing tradeoff} \\label{subsec:dm_tradeoff}\n\nAs in \\cite{Zheng03}, the multiplexing gain $r = \\lim_{S \\rightarrow\n \\infty} \\frac{R(S)}{\\log S}$ where $S$ is the SNR and $R(S)$ is the\nrate at an SNR level of $S$.\nFollowing \\cite{Zheng03}, we parameterize the system, in terms of the\nSNR $S$ and the multiplexing gain, $0