diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzziuji" "b/data_all_eng_slimpj/shuffled/split2/finalzziuji" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzziuji" @@ -0,0 +1,5 @@ +{"text":"\\section{Introduction} \n\n\nA cornerstone of fractal fluids is \nthe non-integer dimension \\cite{Fractal1,Fractal2,CM}. \nThe mass of fractal fluid \nsatisfies a power law relation $M \\sim R^{D}$, \nwhere $M$ is the mass of the ball region with radius $R$,\nand $D$ is the mass dimension \\cite{TarasovSpringer}. \nFractal fluid can be described by four different approaches:\n(1) Using the methods of \"Analysis on fractals\" \n\\cite{Kugami,Strichartz-1,Strichartz-2,Harrison,Kumagai,DGV} \nit is possible to describe fractal media;\n(2) An application of fractional-differential continuum models \nsuggested in \\cite{CCC2001,CCC2002}, and then developed in \n\\cite{CCC2004a,CCC2004c,CCSPZ2009,CCC2009,Yang2013a,Yang2013b},\nwhere so-called local fractional derivatives \\cite{Yang2013c} \nare used; \n(3) Applying fractional-integral continuum models suggested in \n\\cite{PLA2005-1,AP2005-2,IJMPB2005-2,MPLB2005-1,TarasovSpringer} \n(see also \\cite{MOS-3}-\\cite{MOS-5} and \\cite{Bal-1}-\\cite{Bal-4}), where\nintegrations of non-integer orders and \na notion of density of states \\cite{TarasovSpringer} are used; \n(4) Fractal media can be described by using\nthe theory of integration and differentiation for\na non-integer dimensional space \\cite{Collins,Stillinger,PS2004}\n(see also \\cite{CNSNS2015,JMP2014}.\n\n\nLet us note that main difference of \nthe continuum models with non-integer dimensional spaces\nform the fractional continuum models suggested in \n\\cite{PLA2005-1,AP2005-2,IJMPB2005-2,MPLB2005-1,TarasovSpringer}\nmay be reduced to the following.\n(a) Arbitrariness in the choice of the numerical factor\nin the density of states is fixed by the equation\nof the volume of non-integer dimensional ball region.\n(b) In the fractional continuum models suggested \nin \\cite{PLA2005-1,AP2005-2,TarasovSpringer}, \nthe differentiations are integer orders whereas\nthe integrations are non-integer orders.\nIn the continuum models with non-integer dimensional spaces\nthe integrations and differentiations are defined \nfor the spaces with non-integer dimensions.\n\n\nIn this paper, we consider approach based on\nthe non-integer dimensional space.\nThe power law $M \\sim R^{D}$ can be naturally \nderived by using the integrations \nin non-integer dimensional space \\cite{Collins},\nwhere the mass dimension of fractal fluid\nis connected with the dimension of this space.\nA vector calculus for non-integer dimensional space \nproposed in this paper\nallows us to use continuum models with\nnon-integer dimensional spaces to describe\nfor fractal fluids.\nThis is due to the fact that\nalthough the non-integer dimension \ndoes not reflect completely the geometric\nproperties of the fractal media, it nevertheless permits\na number of important conclusions about the behavior\nof fractal structures.\nTherefore continuum models with non-integer dimensional spaces\ncan be successfully used to describe fractal fluids.\n\nIntegration over non-integer dimensional spaces \nare actively used in the theory of critical phenomena and \nphase transitions in statistical physics \n\\cite{WilsonFisher,WK1974}, and \nin the dimensional regularization of\nultraviolet divergences \nin quantum field theory \\cite{HV1972,Leibbrandt,Collins}. \nThe axioms for integrations in non-integer dimensional space are proposed in \\cite{Wilson,Stillinger} and this type of integration is considered in the book by Collins \\cite{Collins}\nfor rotationally covariant functions. \nIn the paper \\cite{Stillinger} a mathematical basis of integration on non-integer dimensional space is given. \nStillinger \\cite{Stillinger} suggested a generalization of \nthe Laplace operator for non-integer dimensional spaces also. \nUsing a product measure approach, \nthe Stillinger's methods \\cite{Stillinger}\nhas been generalized by Palmer and Stavrinou \\cite{PS2004} \nfor multiple variables case\nwith different degrees of confinement in orthogonal directions. \nThe scalar Laplace operators suggested \nby Stillinger in \\cite{Stillinger} \nand Palmer, Stavrinou in \\cite{PS2004}\nfor non-integer dimensional spaces, \nhave successfully been used for \neffective descriptions in physics and mechanics. \nThe Stillinger's form of Laplacian \nfor the Schr\\\"odinger equation in non-integer dimensional space \nis used by He \\cite{XFHe1,XFHe2,XFHe3} to describe\na measure of the anisotropy and \nconfinement by the effective non-integer dimensions. \nQuantum mechanical models with non-integer (fractional) \ndimensional space have been discussed in \n\\cite{Stillinger,PS2004,Thilagam1997b,MA2001a,MA2001b,QM1,QM5}\nand \\cite{Muslih2010,MA2012,QM8,QM9}.\nRecent progress in non-integer dimensional space approach\nalso includes description of\nthe fractional diffusion processes in\nnon-integer dimensional space in \\cite{LSTLRL}, and\nthe electromagnetic fields in non-integer dimensional space\nin \\cite{MB2007,BGG2010,MSBR2010}\nand \\cite{ZMN2010,ZMN2011a,ZMN2011b,ZMN2011c,ZMN}. \n\n\nUnfortunately, \\cite{Stillinger,PS2004} proposed \nonly the second order differential operators \nfor scalar fields in the form of the scalar Laplacian \nin the non-integer dimensional space. \nA generalization of the vector Laplacian \\cite{VLap}\nfor the non-integer dimensional space is not suggested\nin \\cite{Stillinger,PS2004}. \nThe first order operators such as gradient, \ndivergence, curl operators \nare not considered in \\cite{Stillinger,PS2004} also.\nIn the work \\cite{ZMN} \nthe gradient, divergence, and curl operators\nare suggested only as approximations of the square of \nthe Laplace operator.\nConsideration only the scalar Laplacian \nin the non-integer dimensional space approach\ngreatly restricts us in application of continuum models\nwith non-integer dimensional spaces for fractal fluids and material.\nFor example, we cannot use the Stillinger's form of Laplacian\nfor vector field ${\\bf v}({\\bf r},t)$ in hydrodynamics of fractal fluids, in fractal theory of elasticity and thermoelasticity, \nin electromagnetic theory of fractal media\nto describe processes\nin the framework non-integer dimensional space approach.\n\nIn this paper, we propose to use a vector calculus\nfor non-integer dimensional space, and we define\nthe first and second orders differential vector\noperations such as gradient, divergence, \nthe scalar and vector Laplace operators \nfor non-integer dimensional space.\nIn order to derive the vector differential operators \nin non-integer dimensional space\nwe use the method of analytic continuation in dimension.\nFor simplification we consider rotationally covariant \nscalar and vector functions that are independent of angles.\nIt allows us to reduce differential equations \nin non-integer dimensional space to\nordinary differential equations with respect to $r$. \nThe proposed operators allow us to describe fractal media\nto describe processes in the framework of continuum models\nwith non-integer dimensional spaces. \nIn this paper we describe a Poiseuille flow of \nan incompressible viscous fractal fluid in the pipe. \nA generalization of the Navier-Stokes equation \nfor non-integer dimensional space to describe \nfor fractal fluid are suggested. \nA solution of this equation for steady flow of \nfractal fluid in a pipe and corresponding \nfractal fluid discharge are derived.\n\n\n\n\\section{Fractal fluids} \n\n\nA basic characteristic of fractal fluids\nis the non-integer dimensions such as mass \nor \"particle\" dimensions \\cite{TarasovSpringer}. \nFor fractal fluids \nthe number of particles $N_D(W)$ or mass $M_D(W)$ \nin any region $W \\subset \\mathbb{R}^3$ of this fluid \nincrease more slowly \nthan the 3-dimensional volume $V_3(W)$ of this region.\nFor the ball region $W$ with radius $R$ in \nan isotropic fractal fluid, \nthis property can be described by the relation\nbetween the number of particles $N_D(W)$\nin the region $W$ of fractal fluid, and \nthe radius $R$ in the form\n\\be \\label{NDW} \nN_D(W) = N_0 (R\/ R_0)^D , \\quad R\/R_0 \\gg 1 , \n\\ee\nwhere $R_0$ is the characteristic size of fractal fluid\nsuch as a minimal scale of self-similarity \nof a considered fractal fluid. \nThe number $D$ is called the \"particle\" dimension. \nIt is a measure of how the fluid particles\nfill the space.\nThe parameter $D$ does not depend on the shape \nof the region $W$.\nTherefore fractal fluids can be considered \nas fluid with non-integer \"particle\" or mass dimension. \n\nIf the fractal fluid consists of particles \nwith identical masses $m_0$, then relation (\\ref{NDW}) gives\n\\be \\label{MDW} \nM_D(W) = M_0 (R\/ R_0)^D , \\quad R\/R_0 \\gg 1 , \n\\ee\nwhere $M_0=m_0 \\, N_0$. \nIn this case, the mass dimension coincides\nwith the \"particle\" dimension.\n\nAs the basic mathematical tool for\ncontinuum models of fractal fluids,\nwe propose to use the integration and differentiation\nin non-integer dimensional spaces.\nIn Section 7, we will show that \nthe power-law relation (\\ref{MDW}) for \nan isotropic fractal fluid \ncan be naturally derived by using the integration over \nnon-integer dimensional space, where\nthe space dimension is equal to\nthe mass dimension of fractal fluid. \n\n\nIn order to describe fractal fluid by \ncontinuum models with non-integer dimensional spaces,\nwe use the concepts of \ndensity of states $c_3(D,{\\bf r})$\nthat describes how closely packed permitted \nplaces (states) in the space $\\mathbb{R}^3$, \nwhere the fractal fluid is distributed.\nThe expression $dV_D({\\bf r})=c_3(D,{\\bf r})dV_3$ \nis equal to the number of permitted places (states) \nbetween $V_3$ and $V_3 +dV_3$ in $\\mathbb{R}^3$. \nThe notation $d^D {\\bf r}$ \nalso will be used instead of $ d V_D ({\\bf r})$.\nNote that density of states and \ndistribution function are different concepts, \nand it is impossible to describe \nall properties of fractal fluids\nby the distribution function only.\n\nFor fractal fluids, we can use the equation\n\\be \ndN_D(W)= n({\\bf r}) \\, d V_D ({\\bf r}) , \n\\ee\nwhere $n({\\bf r})$ is a concentration of particles\nthat describes a distribution of number of particles\non a set of permitted places (possible states).\nThe density of states is chosen such that\n$d V_D ({\\bf r}) = c_3(D,{\\bf r}) \\, dV_3$\ndescribes the number of permitted states in $dV_3$.\n\nThe form of the function $c_3(D,{\\bf r})$ \nis defined by symmetries of \nconsidered problem and properties of \nthe described fractal fluid.\nA general property of density of states\nfor fractal fluids is a power-law type \nof these functions that reflects a scaling property \n(fractality) of the fractal fluid. \nTo simplify our consideration in this paper \nwe will consider only isotropic fractal fluids \nwith density of states that is independent of angles.\nIn this case, the form of density of states is\ndefined such that $dV_D$ is an elementary volume\nof the non-integer dimensional space.\n\nIn the continuum models of fractal fluids,\nwe should work with the dimensionless variables\n$x\/R_0 \\to x$, $y\/R_0 \\to x$, $z\/R_0 \\to x$, \n${\\bf r}\/R_0 \\to {\\bf r}$,\nin order to physical quantities of \nfractal fluids have correct physical dimensions. \n\n\n\n\n\n\n\\section{Vector differential operators \nin non-integer dimensional space}\n\nTo derive equations for vector differential operators \nin non-integer dimensional space,\nwe use equations for the differential operators \nin the spherical (and cylindrical) coordinates in $\\mathbb{R}^n$ \nfor arbitrary $n$ to highlight the explicit \nrelations with dimension $n$.\nThen the vector differential operators \nfor non-integer dimension $D$ \ncan be defined by continuation in \ndimension from integer $n$ to non-integer $D$. \nTo simplify we will consider only scalar \nfields $\\varphi$ and vector fields ${\\bf v}$ \nthat are independent of angles \n\\[ \\varphi({\\bf r}) =\\varphi(r) ,\n\\quad {\\bf v}({\\bf r})={\\bf v} (r) = v_r\\, {\\bf e}_r , \\]\nwhere $r=|{\\bf r}|$ is the radial distance,\n${\\bf e}_r={\\bf r} \/r$ is the local orthogonal unit \nvector in the directions of increasing $r$, \nand $v_r=v_r(r)$ is the radial component of ${\\bf v}$.\nWe will work with rotationally covariant functions only. \nThis simplification is analogous to the simplification\nfor definition of integration over non-integer dimensional space\ndescribed in Section 4 of the book \\cite{Collins}.\n\n\n\\subsection{Vector differential operators \nfor spherical and cylindrical cases}\n\n\n\nUsing the continuation from integer $n$ \nto arbitrary non-integer $D$, \nwe can get explicit definitions of differential operators\nfor non-integer dimensional space in the following forms.\nNote that the same expressions can be obtained by using \nthe integration in non-integer dimensional space\nand the correspondent Gauss's theorem \\cite{CNSNS2015}\n\nLet us define the differential vector operations\nsuch as gradient, divergence, the scalar and \nvector Laplacian for non-integer dimensional space. \nFor simplifications, we assume that\nthe vector field ${\\bf v}={\\bf v}({\\bf r})$ \nbe radially directed and the scalar and vector \nfields $\\varphi({\\bf r})$, ${\\bf v}({\\bf r})$ \nare not dependent on the angles.\n\nThe divergence in non-integer dimensional space \nfor the vector field ${\\bf v}={\\bf v}(r)$ is\n\\be \\label{Div-D}\n\\operatorname{Div}^{D}_{r} {\\bf v} = \n\\frac{\\partial v_r}{\\partial r} + \\frac{D-1}{r} \\, v_r.\n\\ee\n\nThe gradient in non-integer dimensional space \nfor the scalar field $\\varphi=\\varphi (r)$ is\n\\be \\label{Grad-D}\n\\operatorname{Grad}^{D}_r \\varphi = \\frac{\\partial \\varphi}{\\partial r} \\, {\\bf e}_r .\n\\ee\n\nThe scalar Laplacian in non-integer dimensional space \nfor the scalar field $\\varphi=\\varphi (r)$ is\n\\be \\label{S-Delta-D}\n^S\\Delta^{D}_r \\varphi= \n\\operatorname{Div}^{D}_r \\operatorname{Grad}^{D}_{r} \\varphi =\n\\frac{\\partial^2 \\varphi}{\\partial r^2} + \\frac{D-1}{r} \\, \n\\frac{\\partial \\varphi}{\\partial r} .\n\\ee\n\nThe vector Laplacian in non-integer dimensional space \nfor the vector field ${\\bf v}=v(r) \\, {\\bf e}_r$ is\n\\be \\label{V-Delta-D}\n^V\\Delta^{D}_r {\\bf v} = \n\\operatorname{Grad}^{D}_r \\operatorname{Div}^{D}_{r} {\\bf v} =\n\\Bigl(\n\\frac{\\partial^2 v_r}{\\partial r^2} + \\frac{D-1}{r} \\, \n\\frac{\\partial v_r}{\\partial r} - \\frac{D-1}{r^2} \\, v_r\n\\Bigr) \\, {\\bf e}_r.\n\\ee\n\nIf $D=n$, equations (\\ref{Div-D}-\\ref{V-Delta-D})\ngive the well-known formulas for\ninteger dimensional space $\\mathbb{R}^n$.\n\n\nLet us consider a case of axial symmetry \nof the fluid, where the fields \n$\\varphi(r)$ and ${\\bf v}(r)=v_r(r) \\, {\\bf e}_r$\nare also axially symmetric.\nWe will direct the $Z$-axis along the axis of symmetry. \nTherefore we use a cylindrical coordinate system.\n\nThe divergence in non-integer dimensional space \nfor the vector field ${\\bf v}={\\bf v}(r)$ is\n\\be \\label{Div-DC}\n\\operatorname{Div}^{D}_{r} {\\bf v} = \n\\frac{\\partial v_r}{\\partial r} + \\frac{D-2}{r} \\, v_r.\n\\ee\n\nThe gradient in non-integer dimensional space \nfor the scalar field $\\varphi=\\varphi (r)$ is\n\\be \\label{Grad-DC}\n\\operatorname{Grad}^{D}_r \\varphi = \n\\frac{\\partial \\varphi}{\\partial r} \\, {\\bf e}_r .\n\\ee\n\nThe scalar Laplacian in non-integer dimensional space \nfor the scalar field $\\varphi=\\varphi (r)$ is\n\\be \\label{S-Delta-DC}\n^S\\Delta^{D}_r \\varphi = \n\\frac{\\partial^2 \\varphi}{\\partial r^2} + \\frac{D-2}{r} \\, \n\\frac{\\partial \\varphi}{\\partial r} .\n\\ee\n\nThe vector Laplacian in non-integer dimensional space \nfor the vector field ${\\bf v}=v(r) \\, {\\bf e}_r$ is\n\\be \\label{V-Delta-DC}\n^V\\Delta^{D}_r {\\bf v} = \n\\Bigl(\n\\frac{\\partial^2 v_r}{\\partial r^2} + \\frac{D-2}{r} \\, \n\\frac{\\partial v_r}{\\partial r} - \\frac{D-2}{r^2} \\, v_r\n\\Bigr) \\, {\\bf e}_r .\n\\ee\n\n\n\nEquations (\\ref{Div-DC}-\\ref{V-Delta-DC}) can be\neasy generalized for the case $\\varphi=\\varphi(r,z)$ and \n${\\bf v}(r,z)=v_r(r,z) \\, {\\bf e}_r+ v_r(r,z) \\, {\\bf e}_z$.\nIn this case the curl operator for ${\\bf v}(r,z)$\nis different from zero, and\n\\be \\label{Curl-DC}\n\\operatorname{Curl}^{D}_r {\\bf v} = \n\\left( \\frac{\\partial v_r}{\\partial z} -\n\\frac{\\partial v_z}{\\partial r} \\right) \\, {\\bf e}_{\\theta} .\n\\ee\n\n\nFor $D=3$ equations (\\ref{Div-D}) - (\\ref{V-Delta-DC}) and \n(\\ref{Curl-DC}) give the well-known\nexpressions for the gradient, divergence, curl operator,\nscalar and vector Laplacian operators\n\nThe suggested operators for $01$ can be used \nto describe fractal turbulent flow in pipe.\nThis assumption is based on the fact\nthat trajectories of the fluid particles \nare fractal curve, then $\\alpha_x>1$\n(for example, the Koch curve with \n$\\alpha_x=\\ln(4) \/ \\ln(3) \\approx 1.262$).\n\n\n\n\n\\section{Fractal fluid discharge}\n\n\nIn general, fractal fluids cannot be considered \nas a fluid on fractal.\nReal fractal fluids have a characteristic smallest length scale\nsuch as the radius, $R_0$, of a particle \n(for example, an atom or molecule).\nIn real fluids the fractal structure cannot be observed on all\nscales but only those for which $R >R_0$, \nwhere $R_0$ is the characteristic scale of the particles.\nThe concept of non-integer mass dimension of fractal fluid\nis based on the idea of how the mass \nof a fluid region scales with the region size,\nif we assume unchanged density.\nFor many cases, we can write the asymptotic form for the relation between\nthe mass $M_D(W)$ of a ball region $W$ of fluid,\nand the radius $R$ containing this mass as follows:\n\\be \\label{1-1-MR} \nM_D(W)=M_0 \\left( \\frac{R}{R_0} \\right)^D \\ee\nfor $R\/R_0 \\gg 1$.\nThe constant $M_0$ depends on how the spheres of radius $R_0$ \nare packed.\nThe parameter $D$, which is interpreted as a dimension, \ndoes not depend on the shape of the region $W$,\nor on whether the packing of spheres of radius $R_0$ is \nclose packing,\na random packing or a porous packing with a uniform distribution of holes.\nThe non-integer mass dimension $D$ of fractal fluid \nis a measure of how the fluid fills \nthe integer $n$-dimensional Euclidean space it occupies.\nNote that the fact that a fluid is random or contains cavities\ndoes not necessarily imply that the fluid is fractal.\n\n\nUsing the non-integer dimensional space approach,\nwe can calculate the mass of fractal homogeneous fluids. \nScaling law (\\ref{1-1-MR}) is obtained naturally \nin the framework of this approach.\nWe can use the integration in a non-integer dimensional space \n\\cite{Stillinger} that is described by the equation\n\\be \\label{NI-2}\n\\int_{R^D} d^D {\\bf r} \\, \\varphi ({\\bf r}) =\n\\frac{2 \\pi^{(D-1)\/2}}{\\Gamma((D-1)\/2)}\n\\int^{\\infty}_0 dr \\, r^{D-1} \\,\n\\int^{\\pi}_0 d \\theta \\, \\sin^{D-2}\\theta \\, \\varphi (r,\\theta) ,\n\\ee\nwhere $d^D {\\bf r}$ represent the volume element\nin the non-integer dimensional space.\nUsing (\\ref{NI-2}) with $\\varphi (r,\\theta)=1$, and\n\\be \\int^{\\pi}_0 d \\theta \\, sin^{D-2}\\theta = \n\\frac{\\pi^{1\/2}\\, \\Gamma (D\/2-1)}{\\Gamma(D\/2)} , \\ee\nwe get the volume of $D$-dimensional ball with radius $R$ in the form\n\\be\nV_D = \\frac{\\pi^{D\/2}}{\\Gamma(D\/2+1)} \\, R^{D} .\n\\ee\nThe mass of fluid in $W$ is described by the integral \n\\be \\label{1-MW3} \nM_D(W) = \\int_{W} \\rho({\\bf r}) \\, d^D {\\bf r} ,\n\\ee\nwhere ${\\bf r}$ is dimensionless vector variable. \nFor a ball with radius $R$ and constant density \n$\\rho({\\bf r})=\\rho =\\operatorname{const}$, we get \n\\be \\label{M-D}\nM_D(W) = \\rho \\, V_D =\n\\frac{\\pi^{D\/2} \\, \\rho}{\\Gamma(D\/2+1)} \\, R^{D} .\n\\ee\nThis equation define the mass of the \nfractal homogeneous ball region of fluid with volume $V_D$. \nFor $D=3$, equation (\\ref{M-D}) gives the well-known \nequation for mass of non-fractal ball region\n$M_3 =(4 \\rho \\pi \/3) \\, R^3$ because \n$\\Gamma(3\/2) =\\sqrt{\\pi}\/2$ and $\\Gamma(z+1)=z \\, \\Gamma(z)$. \n\n\nLet us determine the mass $Q$ of fluid passing per unit time \nthrough any cross-section of the pipe (called the discharge). \nNot all pipe volume is occupied by fractal fluid.\nThere are areas unoccupied by particles of fractal fluid.\nIn continuum model of fractal fluids, \nwe take into account this fact by using the integration \nin space with non-integer dimension\n\\be \\label{Q-Def}\nQ= \\rho \\, \\frac{2 \\, \\pi^{d\/2}}{\\Gamma(d\/2)} \n\\int^R_0 v_x(r) \\, r^{d-1} \\, dr ,\n\\ee\nwhere $d=D-1$ is non-integer dimension of \nthe cross-section, $\\rho$ is a constant density, \nand $v_x(r)$ is defined by equation (\\ref{PE-FF}).\nSubstitution of (\\ref{PE-FF}) into (\\ref{Q-Def}) gives\n\\be \\label{QR}\nQ= - \\frac{\\rho \\, \\pi^{(D-1)\/2}}{2 \\, (D+1) \\, \\Gamma((D-1)\/2)\\, \\mu} \n\\frac{dp}{dx} \\, R^{D+1} \\quad (0\\mu$.\nIf $02$. Note that, teleportation with $|\\mbox{EPR}\\rangle$ through\nnoisy channels is depicted in Fig.~\\ref{fig1} and it is discussed in\nRef.~\\cite{0h02}. Also, teleportation of 3GHZ state through various\nnoisy channels has been previously studied in Ref.~\\cite{jung08-2}.\nHere, we are interested to investigate the teleportation process for\n$n$GHZ state through noisy channels for $n\\in\\{4,5,6\\}$. For this\npurpose, we need to solve the master equation with Lindblad form\n\\cite{lind76}\n\\begin{equation}\\label{Lindblad}\n\\frac{\\partial \\rho}{\\partial t} = -\\frac{i}{\\hbar} [H_S, \\rho] +\n\\sum_{i, \\alpha} \\left(L_{i,\\alpha} \\rho L_{i,\\alpha}^{\\dagger} -\n\\frac{1}{2} \\left\\{ L_{i,\\alpha}^{\\dagger} L_{i,\\alpha}, \\rho\n\\right\\} \\right),\n\\end{equation}\nin which $L_{i,\\alpha} = \\sqrt{\\kappa_{i,\\alpha}}\n\\sigma^{(i)}_{\\alpha}$ denote Lindblad operators that describe\ndecoherence and act on the $i$th qubit. Also,\n$\\sigma^{(i)}_{\\alpha}$ are the Pauli spin matrices of the $i$th\nqubit with $\\alpha =\\{ x,y,z\\}$, $\\kappa_{i,\\alpha}$ is the\ndecoherence rate, and $H_S$ is the Hamiltonian of the system.\n\n\\begin{figure}\n\\input{fig1.tex}\n\\caption{\\label{fig2}\n A circuit for quantum teleportation through noisy channels with 4GHZ state.\n The four top lines belong to Alice and the bottom line to Bob.\n $M$ denotes measurement and the dotted box represents noisy channel.\n The Lindblad operator is turned on inside the dotted box.}\n\\end{figure}\n\nThe unknown state to be teleportated can be written as a Bloch vector on a\nBloch sphere\n\\begin{equation}\n\\label{unknown} |\\psi_{\\text{in}}\\rangle = \\cos\n\\left(\\frac{\\theta}{2}\\right) e^{i \\phi \/ 2} |0\\rangle + \\sin\n\\left(\\frac{\\theta}{2}\\right) e^{-i \\phi \/ 2} |1\\rangle,\n\\end{equation}\nwhere $\\theta$ and $\\phi$ denote the polar and azimuthal angles,\nrespectively. Fig.~\\ref{fig2} shows a quantum teleportation circuit\nthrough noisy channels with 4GHZ state in which the input state\ninvolves five qubits as the product state of\n$|\\psi_{\\text{in}}\\rangle $ and $|\\mbox{4GHZ}\\rangle$. The four top\nlines (qubits) belong to Alice and bottom one belongs to Bob. The\ndifference of this circuit with the teleportation circuit for EPR\nstate (Fig.~\\ref{fig1}) is the presence of two more controlled-NOT\n$\\left(\\mbox{CNOT} \\right)$ gates between $\\psi_{\\text{in}}$ and\n$\\mbox{4GHZ}$ states. After measurement of the top four qubits, Bob\ngets the teleported state $|\\psi_{\\text{out}}\\rangle$. It is\nconvenient to describe the teleportation in terms of the density\noperator\n\\begin{equation}\n\\label{out1} \\rho_{\\text{out}} = \\mbox{Tr}_{1,2,3,4} \\left[\nU_{\\mbox{\\tiny tel}} \\rho_{in} \\otimes \\varepsilon\n(\\rho_{4\\mbox{\\tiny GHZ}}) U_{\\mbox{\\tiny tel}}^{\\dagger} \\right],\n\\end{equation}\nwhere $\\rho_{\\text{in}} = |\\psi_{\\text{in}}\\rangle \\langle\n\\psi_{\\text{in}}|$ is density matrix of the unknown initial state\nand $\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}})$ is the density matrix\nafter transmission through noisy channel which is given by the\nLindblad equation. In fact, $\\varepsilon$ is a quantum operation\nthat maps $\\rho_{4\\mbox{\\tiny GHZ}}$ to $\\varepsilon\n(\\rho_{4\\mbox{\\tiny GHZ}})$ because of noisy channel and\n$\\rho_{4\\mbox{\\tiny GHZ}}=|\\mbox{4GHZ}\\rangle\\langle\\mbox{4GHZ}|$.\nMoreover, $U_{\\mbox{\\tiny tel}}$ is the unitary operator\ncorresponding to the quantum circuit and $\\mbox{Tr}_{1,2,3,4}$ is\npartial trace over first four qubits which belong to Alice.\n\nFidelity can be used as a tool to measure how much information is\nlost or preserved through noisy quantum channels in quantum\nteleportation process. It can be written as the overlap between the\ninput state $|\\psi_{in}\\rangle$ and the density operator for the\nteleported state $|\\rho_{\\text{out}} \\rangle$,\n\\begin{equation}\n\\label{fidelity} F = \\langle \\psi_{\\text{in}} | \\rho_{\\text{out}} |\n\\psi_{\\text{in}}\\rangle,\n\\end{equation}\nthat depends on an input state and the type of noise. For the\nperfect teleportation the fidelity is equal to unity. Also, $1-F$\nindicates how much information is lost through the teleportation\nprocess. For all possible unknown input states,\nthe average fidelity is given by\n\\begin{equation}\n\\overline{}\\label{average} \\overline{F} = \\frac{1}{4 \\pi}\n\\int_{0}^{2 \\pi} \\mathrm{d} \\phi \\int_{0}^{\\pi} \\mathrm{d} \\theta\n\\sin \\theta F(\\theta, \\phi).\n\\end{equation}\nSimilarly, we find the unitary operator, fidelity and average\nfidelity for 5GHZ and 6GHZ states in the following sections.\n\n\\section{Four-qubit GHZ state with noisy channels}\\label{sec3}\nIn this section, we analytically solve the Lindblad equation,\nEq.~(\\ref{Lindblad}), for 4GHZ state through various noisy channels.\nFirst, consider $(L_{2,x},L_{3,x},L_{4,x},L_{5,x})$ noise channel\nwith $\\kappa_{2,x}=\\kappa_{3,x}=\\kappa_{4,x}=\\kappa_{5,x}=\\kappa$\nthat acts on 4GHZ state. Also, here and throughout the paper we\nassume $H_S=0$.\n\nFor this case, the Lindblad equation involves 16 diagonal and 120\noff-diagonal coupled linear differential equations which make this\nequation difficult to be solved analytically. To overcome this problem,\nwe find the time evolution of the density matrix for infinitesimal\ntime interval $\\delta t$ using the Lindblad equation as\n\\begin{equation}\\label{del}\n\\rho(\\delta t)= \\rho(0) + \\left[ \\sum_{i, \\alpha} \\left(L_{i,\\alpha}\n\\rho(0) L_{i,\\alpha}^{\\dagger} \\right) - \\frac{1}{2} \\left\\{\nL_{i,\\alpha}^{\\dagger} L_{i,\\alpha}, \\rho(0) \\right\\} \\right]\\delta\nt,\n\\end{equation}\nwhere\n\\begin{equation}\\label{ro0}\n\\rho(0)=|\\mbox{4GHZ}\\rangle\\langle\\mbox{4GHZ}|=\\frac{1}{2}\\left[|0\\rangle^{\\otimes\n4} \\langle 0 |^{\\otimes 4} + |0\\rangle^{\\otimes 4}\\langle\n1|^{\\otimes 4}+|1\\rangle^{\\otimes 4} \\langle 0 |^{\\otimes 4} +\n|1\\rangle ^{\\otimes 4}\\langle 1|^{\\otimes 4}\\right].\n\\end{equation}\nSubstituting $\\rho(0)$ in Eq.~(\\ref{del}) results in\n\\begin{eqnarray}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta t} =\n\\frac{1}{2}{\\left(\n\\begin{smallmatrix}\n1 -4 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 &0 & 0 & 0 & 0 & 0 & 0 & 0 & 1-4 \\kappa \\delta t \\\\\n0 & \\kappa \\delta t & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n0 & 0 & \\kappa \\delta t & 0& 0 & 0& 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 \\\\\n0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0& 0 & 0 & 0& {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & \\kappa \\delta t & 0& 0 & 0 & 0 & \\kappa \\delta t & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & \\kappa \\delta t & \\kappa \\delta t & 0 & 0 & 0& 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & \\kappa \\delta t & \\kappa \\delta t & 0 & 0 & 0& 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0& 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & \\kappa \\delta t & 0& 0 & 0 & 0 & \\kappa \\delta t & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0& 0 & 0 & 0& {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n0 & 0 & \\kappa \\delta t & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & \\kappa \\delta t & 0 & 0 \\\\\n0 & \\kappa \\delta t & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n1 -4 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0 & 0 &\n0 & 0 & 1 -4 \\kappa \\delta t \\end{smallmatrix}\\right),}\n\\end{eqnarray}\nwhere \\textcircled{\\emph{n}} denotes $n$ diagonal zeros. Now, because of the form of the density matrix at $t=\\delta t$, we\nuse the following ansatz for the density matrix for all times\n\\begin{eqnarray}\\label{mat1}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}}) = { \\left(\n\\begin{smallmatrix}\na & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & a \\\\\n0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\n0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 \\\\\n0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & b & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & b & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 \\\\\n0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 \\\\\n0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\na & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & a\n\\end{smallmatrix} \\right).}\n\\end{eqnarray}\nInserting this matrix in the Lindblad equation, Eq.~({\\ref{Lindblad}}),\ngives us a set of three coupled differential equations\n\\begin{eqnarray}\\label{diff1}\n\\left\\{\n\\begin{array}{l}\n\\dot{a}(t) = 4k\\Big(b(t)-a(t)\\Big),\\\\\n\\dot{b}(t) = k\\Big(a(t)-4b(t)+3c(t)\\Big),\\\\\n\\dot{c}(t)=4k\\Big(b(t)-c(t)\\Big),\n\\end{array}\\right.\n\\end{eqnarray}\nsubject to the initial conditions $a(0)=1\/2$ and $b(0)=c(0)=0$ (see\nEq.~(\\ref{ro0})). The solutions are readily given by\n\\begin{eqnarray}\n\\label{elements} \\left\\{\n\\begin{array}{l}\na(t) =\\frac{1}{16}\\left( 1 + 6 e^{-4 \\kappa t} + e^{-8 \\kappa t}\\right),\\\\\nb(t) =\\frac{1}{16}\\left( 1 - e^{-8 \\kappa t}\\right),\\\\\nc(t) =\\frac{1}{16}\\left( 1 - 2 e^{-4 \\kappa t} + e^{-8 \\kappa\nt}\\right).\n\\end{array}\\right.\n\\end{eqnarray}\nIn fact, the infinitesimal temporal behavior of the density matrix\nhelped us to properly suggest the solution and consequently reduced\n136 coupled differential equations to three coupled differential\nequations which are readily solved. It is now easy to check that\n$\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}})$, Eq.~(\\ref{mat1}), exactly\nsatisfies the Lindblad equation, Eq.~({\\ref{Lindblad}}), and the validity\nof the ansatz is verfied.\n\nHaving $\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}})$ and $U_{\\mbox{\\tiny\ntel}}$ which can be read off from Fig.~\\ref{fig2}, it is\nstraightforward to compute $\\rho_{out}$. Thus, the fidelity reads\n\\begin{eqnarray}\n\\label{xfidelity} & &F(\\theta, \\phi) = \\frac{1}{2} \\left[(1 + \\sin^2\n\\theta \\cos^2 \\phi) + e^{-4 \\kappa t} (\\cos^2 \\theta + \\sin^2 \\theta\n\\sin^2 \\phi) \\right],\n\\end{eqnarray}\nand the average fidelity is given by\n\\begin{eqnarray}\n\\label{xfbar} \\overline{F} = \\frac{2}{3} + \\frac{1}{3} e^{-4 \\kappa\nt}.\n\\end{eqnarray}\n\nNow consider $(L_{2,y},L_{3,y},L_{4,y},L_{5,y})$ and assume\n$\\kappa_{2,y}=\\kappa_{3,y}=\\kappa_{4,y}=\\kappa_{5,y}=\\kappa$.\nSimilar to the previous case, using the infinitesimal time evolution\nof the density matrix\n\\begin{eqnarray}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta t} =\n\\frac{1}{2}{\\left(\n\\begin{smallmatrix}\n1 -4 \\kappa \\delta t & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 1-4 \\kappa \\delta t \\\\\n0 & \\kappa \\delta t & 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & -\\kappa \\delta t & 0 \\\\\n0 & 0 & \\kappa \\delta t & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & -\\kappa \\delta t & 0 & 0 \\\\\n0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0& 0 & 0 & 0& {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & \\kappa \\delta t & 0& 0 & 0 & 0& -\\kappa \\delta t & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}}& 0& 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & \\kappa \\delta t & -\\kappa \\delta t & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & -\\kappa \\delta t & \\kappa \\delta t & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & -\\kappa \\delta t & 0& 0 & 0 & 0& \\kappa \\delta t & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}}& 0 & 0 & 0& 0 & 0 & 0& {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n0 & 0 & -\\kappa \\delta t & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & \\kappa \\delta t & 0 & 0 \\\\\n0 & -\\kappa \\delta t & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n1 -4 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0 & 0\n&0 & 0 & 1 -4 \\kappa \\delta t\n\\end{smallmatrix}\\right),}\n\\end{eqnarray}\nwe take the following ansatz\n\\begin{eqnarray}\n \\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}}) =\n \\left(\n\\begin{smallmatrix}\na & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & a \\\\\n0 & b& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & -b & 0 \\\\\n0 & 0 & b & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & -b & 0 & 0 \\\\\n0 & 0 & 0 & c & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& c & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & b & 0& 0 & 0 & 0& 0 & 0 & -b & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & c & 0& 0 & 0 & 0 & c & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0&0 & c & 0& 0 & c & 0& 0& 0 & 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & 0& b & -b & 0 & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & 0& -b & b & 0 & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0&0 & c & 0& 0 & c & 0& 0& 0 & 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & c & 0& 0 & 0 & 0 & c & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &- b & 0& 0 & 0 & 0& 0 & 0 & b & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0 & c & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& c & 0 & 0 & 0 \\\\\n0 & 0 &- b & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & b & 0 & 0 \\\\\n0 & -b & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\na & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & a\n\\end{smallmatrix}\n\\right).\n\\end{eqnarray}\nInserting this matrix in the Lindblad equation, Eq.({\\ref{Lindblad}}),\ngives the previous set of coupled differential equations, Eq.~(\\ref{diff1}), and consequently the solutions agree with\nEq.~(\\ref{elements}). For this case the fidelity becomes\n\\begin{eqnarray}\n\\label{yfidelity} & &F(\\theta, \\phi) = \\frac{1}{2} \\left[1 +( \\sin^2\n\\theta \\sin^2 \\phi + \\cos^2 \\theta) e^{-4 \\kappa t} + \\sin^2 \\theta\n\\cos^2 \\phi e^{-8 \\kappa t} \\right],\n\\\\ \\nonumber\n\\end{eqnarray}\nand the average fidelity reads\n\\begin{eqnarray}\n\\label{yfbar} \\overline{F} = \\frac{1}{2} +\\frac{1}{3} e^{-4 \\kappa\nt} +\\frac{1}{6} e^{-8 \\kappa t}.\n\\end{eqnarray}\n\n\nFor the third case consider $(L_{2,z},L_{3,z},L_{4,z},L_{5,z})$ and\nassume $\\kappa_{2,z}=\\kappa_{3,z}=\\kappa_{4,z}=\\kappa_{5,z}=\\kappa$.\nThe infinitesimal time evolution of the density matrix gives\n\\begin{eqnarray}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta t} =\n\\frac{1}{2} \\left( |0\\rangle^{\\otimes 4} \\langle 0 |^{\\otimes 4} +\n|1\\rangle ^{\\otimes 4}\\langle 1| ^{\\otimes 4}\\right) +\n\\frac{1-8\\kappa \\delta t}{2} \\left(|0\\rangle ^{\\otimes 4} \\langle 1|\n^{\\otimes 4}+ |1\\rangle ^{\\otimes 4}\\langle 0 |^{\\otimes 4} \\right).\n\\end{eqnarray}\nSo the ansatz is\n\\begin{eqnarray}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}}) = a \\left( |0\\rangle^{\\otimes\n4} \\langle 0 |^{\\otimes 4} + |1\\rangle ^{\\otimes 4}\\langle 1|\n^{\\otimes 4}\\right) + b\\left(|0\\rangle ^{\\otimes 4} \\langle 1|\n^{\\otimes 4}+ |1\\rangle ^{\\otimes 4}\\langle 0 |^{\\otimes 4} \\right).\n\\end{eqnarray}\nInserting this matrix in the Lindblad equation, Eq.~({\\ref{Lindblad}}),\nresults in\n\\begin{eqnarray}\\label{diff2}\n\\left\\{\n\\begin{array}{l}\n\\dot{a}(t) = 0,\\\\\n\\dot{b}(t) =-8k\\, b(t) ,\n\\end{array}\\right.\n\\end{eqnarray}\nsubject to the initial condition $a(0)=b(0)=1\/2$. The solution is\n\\begin{equation}\n\\label{zmatrix} \\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}}) = \\frac{1}{2}\n\\left( |0\\rangle^{\\otimes 4} \\langle 0 |^{\\otimes 4} + |1\\rangle\n^{\\otimes 4}\\langle 1| ^{\\otimes 4}\\right) + \\frac{1}{2} e^{-8\n\\kappa t} \\left(|0\\rangle ^{\\otimes 4} \\langle 1| ^{\\otimes 4}+\n|1\\rangle ^{\\otimes 4}\\langle 0 |^{\\otimes 4} \\right).\n\\end{equation}\nAlso, the fidelity and its average read\n\\begin{eqnarray}\n\\label{z,f,fbar}\n\\begin{array}{l}\\displaystyle\nF(\\theta, \\phi) = 1 - \\frac{1}{2} \\left(1 - e^{-8 \\kappa t} \\right)\n\\sin^2 \\theta, \\\\\\\\ \\displaystyle\\overline{F} = \\frac{2}{3} +\n\\frac{1}{3} e^{-8 \\kappa t}.\n\\end{array}\n\\end{eqnarray}\n\nThe next noisy channel is the isotropic noisy channel. For this case, the\nmaster equation involves twelve Lindblad operators\n$(L_{2,\\alpha},L_{3,\\alpha},L_{4,\\alpha},L_{5,\\alpha})$ with\n$\\alpha\\in\\{x,y,z\\}$. At $t=\\delta t$ we have\n\\begin{eqnarray}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta t} =\n\\frac{1}{2}{\\left(\n\\begin{smallmatrix}\n1 -8 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 1-16 \\kappa \\delta t \\\\\n0 & 2\\kappa \\delta t & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 2\\kappa \\delta t & 0& 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0& 0 & 0 & 0& {\\footnotesize \\textcircled{\\tiny 1}} & 0& 0 & 0 \\\\\n0 & 0 & 0& 0 & 2\\kappa \\delta t & 0& 0 & 0 & 0& 0 & 0 & 0 & 0& 0 \\\\\n0 & 0 & 0& 0& 0 &{\\footnotesize \\textcircled{\\tiny 2}} & 0& 0 & {\\footnotesize \\textcircled{\\tiny 2}} &0 & 0 & 0& 0& 0 \\\\\n0 & 0 & 0& 0& 0 &0 & 2\\kappa \\delta t & 0 & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & 0 & 2\\kappa \\delta t & 0 & 0 & 0& 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& 2\\kappa \\delta t & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0& 0 & 0 & 0& 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & 2\\kappa \\delta t & 0 & 0 \\\\\n0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 2\\kappa \\delta t & 0 \\\\\n1 -16 \\kappa \\delta t & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0\n&0 & 0 & 1 -8 \\kappa \\delta t\n\\end{smallmatrix}\\right).}\n\\end{eqnarray}\nSo we take the ansatz\n\\begin{eqnarray}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}}) =\n \\left(\n\\begin{smallmatrix}\na & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & d \\\\\n0 & b& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & b & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & c & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & b & 0& 0 & 0 & 0& 0 & 0 & 0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & c & 0& 0 & 0 & 0 & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0&0 & c & 0& 0 & 0 & 0& 0& 0 & 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & 0& b & 0 & 0 & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & 0& 0 & b & 0 & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0&0 & 0 & 0& 0 & c & 0& 0& 0 & 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & 0 & 0& 0 & 0 & 0 & c & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & b & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& c & 0 & 0 & 0 \\\\\n0 & 0 &0 & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & b & 0 & 0 \\\\\n0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\nd & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & a\n\\end{smallmatrix}\n\\right).\n\\end{eqnarray}\nInserting this solution in the Lindblad equation, Eq.~({\\ref{Lindblad}}),\nwe find\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\n\\dot{a}(t) = 8k\\Big(b(t)-a(t)\\Big),\\\\\n\\dot{b}(t) = 2k\\Big(a(t)-4b(t)+3c(t)\\Big),\\\\\n\\dot{c}(t)=8k\\Big(b(t)-c(t)\\Big),\\\\\n\\dot{d}(t)=-16k\\,d(t),\\\\\n\\end{array}\\right.\n\\end{eqnarray}\nsubject to the initial conditions $a(0)=d(0)=1\/2$ and $b(0)=c(0)=0$.\nThe solutions are\n\\begin{eqnarray}\\left\\{\n\\begin{array}{l}\na(t) =\\frac{1}{16}\\Big( 1 + 6 e^{-8 \\kappa t} + e^{-16 \\kappa t}\\Big),\\\\\nb(t) =\\frac{1}{16}\\Big( 1 - e^{-16 \\kappa t}\\Big),\\\\\nc(t) = \\frac{1}{16}\\Big(1 - 2 e^{-8 \\kappa t} + e^{-16 \\kappa t}\\Big),\\\\\nd(t) = \\frac{1}{2} e^{-16 \\kappa t}.\n\\end{array}\\right.\n\\end{eqnarray}\nAlso the fidelity is\n\\begin{eqnarray}\n& &F(\\theta, \\phi) = \\frac{1}{2} \\left[ 1 + e^{-8 \\kappa t} \\cos^2\n\\theta + e^{-16 \\kappa t} \\sin^2 \\theta \\right],\n\\end{eqnarray}\nand\n\\begin{eqnarray}\n\\label{dfbar} \\overline{F} = \\frac{1}{6} \\left( 3 + e^{-8 \\kappa t}\n+ 2 e^{-16 \\kappa t} \\right).\n\\end{eqnarray}\n\n\nTo this end, we only considered the noisy channels with the same\naxis. Now, as a different-axis noisy channel, consider\n$(L_{2,x},L_{3,y},L_{4,z},L_{5,x})$ noise with\n$\\kappa_{2,x}=\\kappa_{3,y}=\\kappa_{4,z}=\\kappa_{5,x}=\\kappa$ that\nexhibits the effects of noises in different directions. After an\ninfinitesimal time interval and using the Lindblad equation, the\ndensity matrix can be written as\n\\begin{eqnarray}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta t} =\n\\frac{1}{2}{\\left(\n\\begin{smallmatrix}\n1 -6 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 1-10 \\kappa \\delta t \\\\\n0 & \\kappa \\delta t & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0& 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0 & 0 \\\\\n0 & 0 & 0 & \\kappa \\delta t & 0 & 0 &0 & 0 & -\\kappa \\delta t & 0& 0 & 0 \\\\\n0 & 0 & 0 & 0& {\\footnotesize \\textcircled{\\tiny 2}}& 0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}}& 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & 0& \\kappa \\delta t & \\kappa \\delta t & 0 & 0 & 0& 0 & 0 \\\\\n0 & 0 & 0& 0 & 0& \\kappa \\delta t & \\kappa \\delta t & 0 & 0 & 0& 0 & 0 \\\\\n0 & 0 & 0 & 0& {\\footnotesize \\textcircled{\\tiny 2}}& 0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & -\\kappa \\delta t & 0 & 0& 0& 0 & \\kappa \\delta t & 0& 0 & 0 \\\\\n0 & 0 &{\\footnotesize \\textcircled{\\tiny 2}}& 0 & 0 & 0 & 0& 0 & 0 & {\\footnotesize \\textcircled{\\tiny 2}} & 0 & 0 \\\\\n0 & \\kappa \\delta t & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n1 -10 \\kappa \\delta t & 0 & 0 & 0 & 0& 0 & 0 & 0 & 0\n&0 & 0 & 1 -6 \\kappa \\delta t\n\\end{smallmatrix}\\right).}\n\\end{eqnarray}\nSo, the elements of the density matrix for all time can be read off\nas\n\\begin{eqnarray}\n\\varepsilon(\\rho_{4\\mbox{\\tiny GHZ}}) =\n \\left(\n\\begin{smallmatrix}\na & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & g \\\\\n0 & b& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & h & 0 \\\\\n0 & 0 & c & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & m & 0 & 0 \\\\\n0 & 0 & 0 & d & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& k & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & b & 0& 0 & 0 & 0& 0 & 0 & n & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & d & 0& 0 & 0 & 0 & k & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0&0 & d & 0& 0 & f & 0& 0& 0 & 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & 0& b & h & 0 & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0 &0 & 0& h & b & 0 & 0 & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0& 0&0 & f & 0& 0 & d & 0& 0& 0 & 0& 0 & 0 \\\\\n0 & 0 & 0& 0& 0 & k & 0& 0 & 0 & 0 & d & 0& 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & n & 0& 0 & 0 & 0& 0 & 0 & b & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0 & k & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& d & 0 & 0 & 0 \\\\\n0 & 0 & m & 0& 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & c & 0 & 0 \\\\\n0 & h & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\ng & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & a\n\\end{smallmatrix}\n\\right),\n\\end{eqnarray}\nwhich leads to two sets of four and six coupled differential\nequations, namely\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\n\\dot{a}(t) = 3\\kappa\\Big(b(t)-a(t)\\Big),\\\\\n\\dot{b}(t) = \\kappa\\Big(a(t)-3b(t)+2d(t)\\Big),\\\\\n\\dot{c}(t) = 3\\kappa\\Big(d(t)-c(t)\\Big),\\\\\n\\dot{d}(t) = \\kappa\\Big(2b(t)-3d(t)+c(t)\\Big),\n\\end{array}\\right.\n\\end{eqnarray}\nand\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\n\\dot{f}(t) = \\kappa\\Big(-5f(t)+2h(t)-m(t)\\Big),\\\\\n\\dot{g}(t) = \\kappa\\Big(-5g(t)+2h(t)-n(t)\\Big),\\\\\n\\dot{h}(t) = \\kappa\\Big(f(t)+g(t)-5h(t)-k(t)\\Big),\\\\\n\\dot{k}(t) = \\kappa\\Big(-h(t)-5k(t)+m(t)+n(t)\\Big),\\\\\n\\dot{m}(t)=\\kappa\\Big(-f(t)+2k(t)-5m(t)\\Big),\\\\\n\\dot{n}(t)=\\kappa\\Big(-g(t)+2k(t)-5n(t)\\Big),\n\\end{array}\\right.\n\\end{eqnarray}\nsubject to $a(0)=g(0)=1\/2$ and\n$b(0)=c(0)=d(0)=f(0)=h(0)=k(0)=m(0)=n(0)=0$. The solutions are\nreadily found\n\\begin{eqnarray}\\left\\{\n\\begin{array}{l}\na(t) =e^{2 \\kappa t}g(t)=\\frac{1}{16}\\Big( 1 + 3 e^{-2 \\kappa t} +3 e^{-4 \\kappa t} + e^{-6 \\kappa t}\\Big),\\\\\nb(t) =e^{2 \\kappa t}h(t)=-e^{2 \\kappa t}n(t)=\\frac{1}{16}\\Big( 1 + e^{-2 \\kappa t} - e^{-4 \\kappa t} - e^{-6 \\kappa t}\\Big),\\\\\nc(t) =-e^{2 \\kappa t}m(t)=\\frac{1}{16}\\Big( 1 - 3 e^{-2 \\kappa t} +3 e^{-4 \\kappa t} - e^{-6 \\kappa t}\\Big),\\\\\nd(t) =e^{2 \\kappa t}f(t)=-e^{2 \\kappa t}k(t)=\\frac{1}{16}\\Big( 1 -\ne^{-2 \\kappa t} - e^{-4 \\kappa t} + e^{-6 \\kappa t}\\Big).\n\\end{array}\\right.\n\\end{eqnarray}\nThus, the fidelity, $F(\\theta, \\phi)$, and its average, $\\overline{F}$, are given by\n\\begin{eqnarray}\n& &F(\\theta, \\phi) = \\frac{1}{2} \\left[ 1 + e^{-2 \\kappa t} \\cos^2\n\\theta + e^{-4 \\kappa t} \\sin^2\n\\theta \\cos^2 \\phi + e^{-6 \\kappa t} \\sin^2 \\theta \\sin^2 \\phi \\right],\n\\end{eqnarray}\nand\n\\begin{eqnarray}\n\\label{mixfbar} \\overline{F} = \\frac{1}{6} \\left( 3 + e^{-2 \\kappa t}+ e^{-4 \\kappa t}\n+ e^{-6 \\kappa t} \\right).\n\\end{eqnarray}\n\n\\begin{table}\n\\caption{\\label{tab1}Summary of $F(\\theta, \\phi)$ and $\\overline{F}$\nthrough various noisy channels.}\n\\begin{ruledtabular}\n\\begin{tabular}{c|ccc}\n{} & Noise & 3GHZ & 4GHZ \\\\ \\hline\n{} & Pauli-X & $\\frac{1}{2} \\bigg[(1 + \\sin^2 \\theta \\cos^2 \\phi$) & $\\frac{1}{2} \\bigg[ (1 + \\sin^2 \\theta \\cos^2 \\phi$) \\\\\n{} & {} & $+ e^{-4 \\kappa t} (\\cos^2 \\theta + \\sin^2 \\theta \\sin^2 \\phi) \\bigg]$ & $+ e^{-4 \\kappa t} (\\cos^2 \\theta + \\sin^2 \\theta \\sin^2 \\phi) \\bigg]$ \\\\\n$F(\\theta,\\phi)$ & Pauli-Y &$\\frac{1}{2} \\bigg[1 + \\sin^2 \\theta \\sin^2 \\phi e^{-2 \\kappa t} +\\cos^2 \\theta e^{-4 \\kappa t}$ & $\\frac{1}{2} \\bigg[1 + (\\sin^2 \\theta \\sin^2 \\phi +\\cos^2 \\theta) e^{-4 \\kappa t}$ \\\\\n{} & {} & $+ \\sin^2 \\theta \\cos^2 \\phi e^{-6 \\kappa t} \\bigg]$ & $+\\sin^2 \\theta \\cos^2 \\phi e^{-8 \\kappa t} \\bigg]$\\\\\n{} & Pauli-Z & $1 - \\frac{1}{2} (1 -e^{-6 \\kappa t}) \\sin^2 \\theta$ & $1 - \\frac{1}{2} (1 - e^{-8\\kappa t})\\sin^2 \\theta$\\\\\n{} & isotropic & $\\frac{1}{2} (1 + \\cos^2 \\theta e^{-8 \\kappa t} +\\sin^2 \\theta e^{-12 \\kappa t} )$ & $\\frac{1}{2} (1 + \\cos^2 \\theta e^{-8 \\kappa t} + \\sin^2 \\theta e^{-16 \\kappa t} )$\\\\\\hline\n{} & Pauli-X & $\\frac{2}{3} +\\frac{1}{3} e^{-4 \\kappa t}$ & $\\frac{2}{3} + \\frac{1}{3} e^{-4\\kappa t}$\\\\\n$\\overline{F}$ & Pauli-Y &$\\frac{1}{6} (3 + e^{-2 \\kappa t} + e^{-4 \\kappa t} + e^{-6 \\kappa t})$ & $\\frac{1}{6} (3 + 2 e^{-4 \\kappa t} + e^{-8 \\kappa t})$\\\\\n{} & Pauli-Z & $\\frac{2}{3} +\\frac{1}{3} e^{-6 \\kappa t}$ & $\\frac{2}{3} + \\frac{1}{3} e^{-8\\kappa t}$\\\\\n{} & isotropic & $\\frac{1}{6} (3 + e^{-8 \\kappa t} + 2 e^{-12 \\kappa t} )$ & $\\frac{1}{6} (3 + e^{-8 \\kappa t} + 2 e^{-16 \\kappa t})$\n\\end{tabular}\n\\end{ruledtabular}\n\\end{table}\n\nIn Table \\ref{tab1}, a summary of fidelity and average fidelity for\n3GHZ \\cite{jung08-2} and 4GHZ states is reported and compared. Also,\ntheir average fidelity versus time is depicted in Fig.~\\ref{fig3}\nfor various noisy channels. Comparing 3GHZ and 4GHZ states shows\nthat for $(L_{2,x},L_{3,x},L_{4,x},L_{5,x})$ noise both states have\nthe same fidelity. This result also agrees with Bell state\n$|\\beta_{00}\\rangle=\\frac{1}{\\sqrt{2}}(|00\\rangle+|11\\rangle)$\n\\cite{0h02}. However, for other cases 3GHZ state is more robust,\ni.e., loses less quantum information in the quantum teleportation\nprocess with respect to 4GHZ state. Note that, for the isotropic\ncase, the fidelities are approximately equal. These results and\nthose obtained in Refs.~\\cite{jung08-2,0h02} show that increasing\nthe number of qubits can enhance the rate of information lost in\nquantum teleportation process. Moreover, using a proper ansatz for\nthe density matrix, we reduced the number of coupled differential\nequations from 136 to at most four coupled equations.\nFig.~\\ref{fig7} shows average fidelity for 4GHZ state through\nvarious noises. As it can be seen from the figure,\n$(L_{2,x},L_{3,x},L_{4,x},L_{5,x})$ noise does lose less quantum\ninformation with respect to others. The next noise with small\ninformation lost is $(L_{2,x},L_{3,y},L_{4,z},L_{5,x})$ for $\\kappa\nt<0.2$. However, for $\\kappa t>0.2$,\n$(L_{2,z},L_{3,z},L_{4,z},L_{5,z})$ noise represents a better\nbehavior. Moreover, the isotropic noise and the noise in $y$\ndirection always result in low fidelity quantum teleportation. In\nthe following sections, we exactly solve the Lindblad equation for\n5GHZ and 6GHZ states through two types of noisy channels.\n\n\\begin{figure}\n\\centering\n\\includegraphics[width=8cm]{fig4GHZ.eps}\n\\caption{\\label{fig7} The plot of time dependence of average\nfidelity through noisy channels for\n4GHZ state.}\n\\end{figure}\n\n\n\\begin{figure}\n\\centerline{\\begin{tabular}{ccc}\n\\includegraphics[width=8cm]{figF1.eps}\n &\\hspace{2.cm}&\n\\includegraphics[width=8cm]{figF2.eps}\\\\\n\\includegraphics[width=8cm]{figF3.eps}\n &\\hspace{2.cm}&\n\\includegraphics[width=8cm]{figF4.eps}\n\\end{tabular}}\n\\caption{\\label{fig3} The plot of time dependence of average\nfidelity for Pauli-X (left up),\nPauli-Y (right up),\nPauli-Z (left down), and isotropic\n(right down) noisy channels.}\n\\end{figure}\n\n\\section{Five-qubit GHZ state with noisy channels}\\label{sec4}\nIn this section, we teleport 5GHZ state through noisy channels as\ndepicted in Fig.~\\ref{fig4}. For this case the solution of the\nLindblad equation is a $32\\times32$ matrix that results in a set of\n32 diagonal and 496 off-diagonal coupled differential equations.\nHowever, we show that the number of required equations can be\nconsiderably reduced by choosing appropriate ansatz for the density\nmatrix.\n\n\\begin{figure}[t]\n\\input{fig3.tex}\n\\caption{\\label{fig4}\n A circuit for quantum teleportation through noisy channels with 5GHZ state.\n The five top lines belong to Alice and the bottom line to Bob.\n $M$ denotes measurement and the dotted box represents noisy channel.\n The Lindblad operator is turned on inside the dotted box.}\n\\end{figure}\n\nFirst, consider ($L_{2,x}$,$L_{3,x}$,$L_{4,x}$,$L_{5,x}$,$L_{6,x}$)\nnoise and assume\n$\\kappa_{2,x}=\\kappa_{3,x}=\\kappa_{4,x}=\\kappa_{5,x}=\\kappa_{6,x}=\\kappa$.\nThe infinitesimal time evolution of the density matrix now reads\n\n\\begin{eqnarray}\n\\hspace{-1cm} \\varepsilon(\\rho_{5\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta\nt} = \\frac{1}{2}{ \\left(\n\\begin{smallmatrix}\n1- 5 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1- 5 \\kappa \\delta t \\\\\n0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 \\\\\n0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & \\kappa \\delta t & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 & {\\footnotesize \\textcircled{\\tiny 3}} & 0 & 0& 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 3}} & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 &0 & 0& {\\footnotesize \\textcircled{\\tiny 6}} & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 6}} & 0 & 0 & 0 & 0& 0 & 0 & 0 \\\\\n0& 0 & 0 & 0& 0& 0 & 0 & 0 & \\kappa \\delta t & \\kappa \\delta t & 0 & 0 & 0& 0& 0 & 0 & 0 & 0 \\\\\n0& 0 & 0 & 0& 0& 0 & 0 & 0 & \\kappa \\delta t & \\kappa \\delta t & 0 & 0 & 0& 0& 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 &0 & 0& {\\footnotesize \\textcircled{\\tiny 6}} & 0& 0 & {\\footnotesize \\textcircled{\\tiny 6}} & 0 & 0 & 0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0& 0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 & {\\footnotesize \\textcircled{\\tiny 3}} & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 3}} &0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 &0 & 0 & \\kappa \\delta t & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0& 0& 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 \\\\\n0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n1- 5 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 &\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0\n& 1- 5 \\kappa \\delta t\n\\end{smallmatrix} \\right).}\n\\hspace{0.5cm}\n\\end{eqnarray}\nSo we take the ansatz as\n\\begin{eqnarray}\n\\hspace{-1cm} \\varepsilon(\\rho_{5\\mbox{\\tiny GHZ}}) = \\left(\n\\begin{smallmatrix}\na & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & a \\\\\n0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\n0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 \\\\\n0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}}_{1} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}}_{1} & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & b & 0 & 0& 0 & 0 & 0 & 0 &0 & 0 & b & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 & \\tiny\\textcircled{\\emph{c}}_{3} & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}}_{3} &0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 & 0 & b & 0 & 0 & 0 & 0 & b & 0 & 0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 &0 & 0& \\tiny\\textcircled{\\emph{c}}_{6} & 0 & 0 & \\tiny\\textcircled{\\emph{c}}_{6} &0 & 0 & 0 & 0& 0 & 0 & 0 \\\\\n0& 0 & 0 & 0& 0& 0 & 0 & 0& b & b & 0 & 0 & 0& 0& 0 & 0 & 0 & 0 \\\\\n0& 0 & 0 & 0& 0& 0 & 0 & 0 & b & b & 0 & 0 & 0& 0& 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 &0 & 0& \\tiny\\textcircled{\\emph{c}}_{6} & 0 & 0 & \\tiny\\textcircled{\\emph{c}}_{6} &0 & 0 & 0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 & 0 & b & 0 & 0 & 0 & 0 & b & 0& 0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 &0 & \\tiny\\textcircled{\\emph{c}}_{3} & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}}_{3} &0 & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0& 0 & b &0 & 0& 0 & 0 & 0 & 0 &0 & 0 & b & 0& 0 & 0 & 0 \\\\\n0 & 0 & 0 &\\tiny\\textcircled{\\emph{c}}_{1} & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}}_{1} & 0 & 0 & 0 \\\\\n0 & 0 & b & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & b & 0 & 0 \\\\\n0 & b& 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\na & 0 & 0 & 0 & 0 & 0 & 0& 0 & 0 & 0& 0 & 0 & 0 & 0 & 0 & 0 & 0 & a\n\\end{smallmatrix} \\right).\n\\hspace{0.5cm}\n\\end{eqnarray}\nHere $\\textcircled{\\emph{c}}_{n}$ denotes\n$n$ diagonal $c$.\n\nNow inserting this matrix in Lindblad equation, Eq.~(\\ref{Lindblad}), four\ncoupled differential equations are obtained as follows\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\n\\dot{a}(t) = 5k\\Big(b(t)-a(t)\\Big),\\\\\n\\dot{b}(t) = k\\Big(a(t)-5b(t)+4c(t)\\Big),\\\\\n\\dot{c}(t)=2k\\Big(b(t)-c(t)\\Big).\\\\\n\\end{array}\\right.\n\\end{eqnarray}\nSolving this set of equations with the initial conditions\n$a(0)=1\/2$, $b(0)=c(0)=0$, leads to the following solution\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\na(t) =\\frac{1}{32}\\Big( 1 + 10 e^{-4 \\kappa t} + 5e^{-8 \\kappa t}\\Big),\\\\\nb(t) =\\frac{1}{32}\\Big( 1 + 2 e^{-4 \\kappa t}- 3e^{-8 \\kappa t}\\Big),\\\\\nc(t) = \\frac{1}{32}\\Big(1 - 2 e^{-4 \\kappa t} + e^{-8 \\kappa\nt}\\Big).\n\\end{array}\\right.\n\\end{eqnarray}\nSubstituting $\\varepsilon(\\rho_{5\\mbox{\\tiny GHZ}})$ in\nEq.~(\\ref{out1}) and using Eqs.~(\\ref{fidelity}) and (\\ref{average})\nfidelity and its average are given by\n\\begin{eqnarray}\nF(\\theta, \\phi) &=& \\frac{1}{2} \\left[ 1 + \\sin^2 \\theta \\cos^2\n\\phi+ e^{-4 \\kappa t} (\\cos^2 \\theta +\n \\sin^2 \\theta \\sin^2 \\phi )\\right],\\\\\n\\overline{F} &=& \\frac{1}{3} \\left( 2 + e^ {-4 \\kappa t} \\right).\n\\end{eqnarray}\n\nFor ($L_{2,z}$,$L_{3,z}$,$L_{4,z}$,$L_{5,z}$,$L_{6,z}$) noise with\n$\\kappa_{2,z}=\\kappa_{3,z}=\\kappa_{4,z}=\\kappa_{5,z}=\\kappa_{6,z}=\\kappa$,\nthe infinitesimal evolution matrix is\n\\begin{eqnarray}\n\\varepsilon(\\rho_{5\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta t} =\n\\frac{1}{2} \\left( |0\\rangle^{\\otimes 5} \\langle 0 |^{\\otimes 5} +\n|1\\rangle ^{\\otimes 5}\\langle 1| ^{\\otimes 5}\\right) +\n\\frac{1-10\\kappa \\delta t}{2} \\left(|0\\rangle ^{\\otimes 5} \\langle\n1| ^{\\otimes 5}+ |1\\rangle ^{\\otimes 5}\\langle 0 |^{\\otimes 5}\n\\right).\n\\end{eqnarray}\nUsing the ansatz\n\\begin{eqnarray}\n \\varepsilon(\\rho_{5\\mbox{\\tiny GHZ}}) =\na \\left( |0\\rangle^{\\otimes 5} \\langle 0 |^{\\otimes 5} + |1\\rangle\n^{\\otimes 5}\\langle 1| ^{\\otimes 5}\\right) + b\\left(|0\\rangle\n^{\\otimes 5} \\langle 1| ^{\\otimes 5}+ |1\\rangle ^{\\otimes 5}\\langle\n0 |^{\\otimes 5} \\right),\n\\end{eqnarray}\nwe obtain two coupled equations\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\n\\dot{a}(t) =0,\\\\\n\\dot{b}(t) = -10kb(t),\\\\\n\\end{array}\\right.\n\\end{eqnarray}\nsubject to $a(0)=1\/2$, $b(0)=0$. Therefore, the density matrix reads\n\\begin{equation}\n\\label{zmatrix5} \\varepsilon(\\rho_{5\\mbox{\\tiny GHZ}}) = \\frac{1}{2}\n\\left( |0\\rangle^{\\otimes 5} \\langle 0 |^{\\otimes 5} + |1\\rangle\n^{\\otimes 5}\\langle 1| ^{\\otimes 5}\\right) + \\frac{1}{2} e^{-10\n\\kappa t} \\left(|0\\rangle ^{\\otimes 5} \\langle 1| ^{\\otimes 5}+\n|1\\rangle ^{\\otimes 5}\\langle 0 |^{\\otimes 5} \\right),\n\\end{equation}\nand the fidelity and its average are given by\n\\begin{eqnarray}\nF(\\theta, \\phi) &=& 1 - \\frac{1}{2} \\left( 1 - e^{-10 \\kappa t}\\right) \\sin^2 \\theta ,\\\\\n\\overline{F} &=& \\frac{1}{3} \\left( 2 + e^ {-10 \\kappa t} \\right).\n\\end{eqnarray}\n\n\\section{Six-qubit GHZ state with noisy channels}\\label{sec5}\nA quantum circuit for teleportation through noisy channels with\n6GHZ state is depicted in Fig.~\\ref{fig5}. In the dotted box the\nLindblad operators act on the $64\\times64$ density matrix that\ninvolves five Alice's qubits and one Bob's qubits. The Lindblad\nequation, Eq.~(\\ref{Lindblad}), leads to 64 diagonal and 2016 off-diagonal\nlinear coupled differential equations. However, similar to previous\nsections, we first study infinitesimal temporal behavior of the\ndensity matrix and use a proper ansatz to considerably reduce the\nnumber of required equations.\n\n\n\\begin{figure}[t]\n\\input{fig4.tex}\n\\caption{\\label{fig5}\n A circuit for quantum teleportation through noisy channels with 6GHZ state.\n The six top lines belong to Alice and the bottom line to Bob.\n $M$ denotes measurement and the dotted box represents noisy channel.\n The Lindblad operator is turned on inside the dotted box.}\n\\end{figure}\n\n\nFor ($L_{2,x}$,$L_{3,x}$,$L_{4,x}$,$L_{5,x}$,$L_{6,x},L_{7,x}$)\nnoise and\n$\\kappa_{2,x}=\\kappa_{3,x}=\\kappa_{4,x}=\\kappa_{5,x}=\\kappa_{6,x}=\\kappa_{7,x}=\\kappa$,\nthe Lindblad operators after an infinitesimal time transform the\ninput density matrix\n$\\rho(0)=|6\\mbox{GHZ}\\rangle\\langle6\\mbox{GHZ}|$ to\n\\begin{eqnarray}\n\\hspace{-1cm}\\varepsilon(\\rho_{6\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta\nt}= \\frac {1}{2}\\left(\\mbox{\\footnotesize$\n\\begin{smallmatrix}\n 1 - 6 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 - 6 \\kappa \\delta t \\\\\n 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n 0 & 0 & \\kappa \\delta t& 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 \\\\\n 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 3}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 3}} & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 7}} & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 7}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 14}} & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 14}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 14}} & 0 & 0 &{\\footnotesize \\textcircled{\\tiny 14}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 7}} & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 7}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 3}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 3}} & 0 & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 \\\\\n 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & {\\footnotesize \\textcircled{\\tiny 1}} & 0 & 0 & 0 \\\\\n 0 & 0 & \\kappa \\delta t& 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 & 0 \\\\\n 0 & \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\kappa \\delta t & 0 \\\\\n 1 - 6 \\kappa \\delta t & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 - 6 \\kappa \\delta t\n\\end{smallmatrix}$}\\right).\n\\end{eqnarray}\nSo consider\nthe ansatz\n\\begin{eqnarray}\n\\hspace{-2.3cm}\\varepsilon(\\rho_{6\\mbox{\\tiny GHZ}})=\n\\left(\\mbox{\\footnotesize$\n\\begin{smallmatrix}\na & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & a \\\\\n0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\n0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 \\\\\n0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{d}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & d & 0 & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & \\tiny\\textcircled{\\emph{c}} & 0 & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 & 0 & 0 \\\\\n0 & 0 & 0 & c & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & c & 0 & 0 & 0 \\\\\n0 & 0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 & 0 \\\\\n0 & b & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & b & 0 \\\\\na & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & a \\\\\n\\end{smallmatrix}$}\\right),\n\\end{eqnarray}\nwhere $\\small\\textcircled{\\emph{c}}$ and $\\small\\textcircled{\\emph{d}}$ denote two diagonal $c$ and $d$, respectively. Substituting this matrix\ninto the Lindblad equation leads to four coupled equations\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\n\\dot{a}(t) = 6k\\Big(b(t)-a(t)\\Big),\\\\\n\\dot{b}(t) = k\\Big(a(t)-6b(t)+5c(t)\\Big),\\\\\n\\dot{c}(t)=2k\\Big(b(t)-3c(t)+2d(t)\\Big),\\\\\n\\dot{d}(t)=-6k\\Big(c(t)-d(t)\\Big),\\\\\n\\end{array}\\right.\n\\end{eqnarray}\nsubject to $a(0)=1\/2$ and $b(0)=c(0)=d(0)=0$. Thus, the solutions\nread\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\na(t) =\\frac{1}{64}\\left( 1 + 15 e^{-4 \\kappa t} + 15 e^{-8 \\kappa t} + e^{-12 \\kappa t}\\right),\\\\\nb(t) =\\frac{1}{64}\\left( 1 +5 e^{-4 \\kappa t} - 5 e^{-8 \\kappa t} - e^{-12 \\kappa t}\\right),\\\\\nc(t) =\\frac{1}{64}\\left( 1 - e^{-4 \\kappa t} - e^{-8 \\kappa t} +\ne^{-12 \\kappa\nt}\\right),\\\\\nd(t) =\\frac{1}{64}\\left( 1 - 3 e^{-4 \\kappa t} + 3 e^{-8 \\kappa t}-\ne^{-12 \\kappa t}\\right),\n\\end{array}\\right.\n\\end{eqnarray}\nand finally\n\\begin{eqnarray}\nF(\\theta, \\phi) &=& \\frac{1}{2} \\left[ 1 + \\sin^2 \\theta \\cos^2\n\\phi+ e^{-4 \\kappa t} (\\cos^2 \\theta +\n \\sin^2 \\theta \\sin^2 \\phi )\\right],\n \\\\ \\overline{F} &=& \\frac{1}{3} \\left( 2 + e^ {-4 \\kappa t}\n\\right).\n\\end{eqnarray}\n\n\nFor the last case, we study\n($L_{2,z}$,$L_{3,z}$,$L_{4,z}$,$L_{5,z}$,$L_{6,z},L_{7,z}$) noise\nwith\n$\\kappa_{2,z}=\\kappa_{3,z}=\\kappa_{4,z}=\\kappa_{5,z}=\\kappa_{6,z}=\\kappa_{7,z}=\\kappa$.\nFor this case, the temporal evolution matrix is\n\\begin{eqnarray}\n\\varepsilon(\\rho_{6\\mbox{\\tiny GHZ}})\\Big|_{t=\\delta t} =\\frac{1}{2}\n\\left( |0\\rangle^{\\otimes 6} \\langle 0 |^{\\otimes 6} + |1\\rangle\n^{\\otimes 6}\\langle 1| ^{\\otimes 6}\\right) + \\frac{1-12\\kappa\\delta\nt}{2} \\left(|0\\rangle ^{\\otimes 6} \\langle 1| ^{\\otimes 6}+\n|1\\rangle ^{\\otimes 6}\\langle 0 |^{\\otimes 6} \\right),\n\\end{eqnarray}\nTherefore, using the ansatz\n\\begin{eqnarray}\n\\varepsilon(\\rho_{6\\mbox{\\tiny GHZ}}) = a \\left( |0\\rangle^{\\otimes\n6} \\langle 0 |^{\\otimes 6} + |1\\rangle ^{\\otimes 6}\\langle 1|\n^{\\otimes 6}\\right) + b\\left(|0\\rangle ^{\\otimes 6} \\langle 1|\n^{\\otimes 6}+ |1\\rangle ^{\\otimes 6}\\langle 0 |^{\\otimes 6} \\right),\n\\end{eqnarray}\nwe obtain two simple differential equations\n\\begin{eqnarray}\n\\left\\{\n\\begin{array}{l}\n\\dot{a}(t) =0,\\\\\n\\dot{b}(t) = -12kb(t),\\\\\n\\end{array}\\right.\n\\end{eqnarray}\nsubject to $a(0)=b(0)=1\/2$. So the solution is given by\n\\begin{equation}\n\\label{zmatrix6} \\varepsilon(\\rho_{6\\mbox{\\tiny GHZ}}) = \\frac{1}{2}\n\\left( |0\\rangle^{\\otimes 6} \\langle 0 |^{\\otimes 6} + |1\\rangle\n^{\\otimes 6}\\langle 1| ^{\\otimes 6}\\right) + \\frac{1}{2} e^{-12\n\\kappa t} \\left(|0\\rangle ^{\\otimes 6} \\langle 1| ^{\\otimes 6}+\n|1\\rangle ^{\\otimes 6}\\langle 0 |^{\\otimes 6} \\right),\n\\end{equation}\nand the fidelity and its average read\n\\begin{eqnarray}\nF(\\theta, \\phi) &=& 1 - \\frac{1}{2} \\left( 1 - e^{-12 \\kappa t}\\right) \\sin^2 \\theta ,\\\\\n\\overline{F} &=& \\frac{1}{3} \\left( 2 + e^ {-12 \\kappa t} \\right).\n\\end{eqnarray}\n\n\\section{Conclusions}\\label{sec6}\nIn this paper, we studied quantum teleportation through noisy\nchannels for $n$GHZ states, $n\\in\\{4,5,6\\}$), so that the noisy\nchannels lead to the quantum channels to be mixed states. We\nexactly solved the Lindblad equation and obtained corresponding\ndensity matrices after the transmission process. The Lindblad\noperators are responsible for the decoherence of quantum states and\nare defined to be proportional to the Pauli matrices. Solving the\nLindblad equation for $n>2$ is not a trivial task in general. For\ninstance, we need to solve 2080 coupled differential equations to\nfind the density matrix for 6GHZ state. We overcame this problem by\nstudying the temporal evolution of the input state and using a\nproper ansatz for the density matrix. Therefore, we reduced 2080\ncoupled equations to at most four coupled equations which are\nreadily solved. We found the fidelity and the average fidelity for\nvarious cases and showed that for the Lindblad operators\ncorresponding to $x$ direction the fidelity is the same for EPR and\n$n$GHZ states where $n\\in\\{3,4,5,6\\}$. However, 3GHZ state does lose\nless quantum information for other types of noisy channel. Note\nthat, In Ref.~\\cite{jung08-2} the authors only studied the same-axis\nnoisy channels and conjectured that ``average fidelity with\nsame-axis noisy channels are in general larger than average fidelity\nwith different-axis noisy channels''. However, we showed the failure\nof this conjecture for 4GHZ state which is apparent in\nFig.~\\ref{fig7}. In the appendix we showed this conjecture also\nfails for 3GHZ state (see Fig.~\\ref{fig6}). In fact, for\ndifferent-axes noises, the analytical solutions can be obtained in\nthe same way, but the number of coupled differential equations\nusually increases with respect to the same-axes noises.\n\n\\acknowledgments\nWe would like to thank Robabeh Rahimi for fruitful discussions and\nsuggestions and for a critical reading of the paper.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\nSince the seminal work of \\citet{Coo95}, \\textit{statistical shape models} became an emerging tool to capture natural shape variability within a given class of objects.\nAs a result, a large number of shape models were developed during the last decades.\nThe probably most well-known models were built for the human face using textured 3D face scans. \nIntroduced by \\citet{Bla99}, this class of statistical shape models is commonly known as \\textit{3D Morphable Models} (3DMMs) and includes models such as the \\textit{Basel Face Model} (BFM) and \\textit{Large Scale Facial Model} (LSFM) presented by \\citet{Pay09} and \\citet{Boo16}, respectively.\nWell-studied applications of 3DMMs include face recognition, expression transfer between individuals, face animation, and 3D face reconstruction from a single 2D photograph \\citep{Egg20}.\nEspecially in the last decade, shape analysis also gained popularity in the field of computational anatomy, where statistical shape models are successfully used to model variations of anatomical objects such as bones and organs. \nLater, these models are utilized for a variety of medical applications including (but not limited to) image segmentation, surgical simulation, therapy planning, and motion analysis \\citep{Amb19}.\nDespite the popularity of statistical shape models in the aforementioned areas, to date and to the best of our knowledge, no publicly available 3D statistical shape model of the female breast exists.\n\nWith breast cancer being the most common malignant neoplasm among women \\citep{Sun21}, successful \\textit{breast reconstruction surgery} (BRS) is crucial for patients undergoing mastectomy.\nIn order to give patients a first impression about what their breast might look like after BRS, surgical outcomes are more and more often simulated using patient-specific 3D breast scans, acquired in a standing position (see Section \\ref{sec:related_work} for an overview). \nTypically, simulations are performed using physically motivated deformable models of the breast.\nWhile these models take into account material properties and physical effects such as gravity, they may \\textit{not always} produce realistic-looking shapes as no prior knowledge in form of example shapes is included \\citep{Roo06,Su11}.\nHence, simulated outcomes might be physically plausible, but definitely lack \\textit{statistical plausibility} in the sense that generated breast shapes are somehow \\textit{likely} or similar to those typically observed within a target population.\nAs the ultimate goal of BRS is an outcome looking as \\textit{natural} as possible, we believe that simulation of surgical outcomes must not only rely on physically based deformable models but also should take into account statistical effects.\nIndeed, the phenomenon that humans tend to compare themselves with others clearly underlines the importance of the fact that simulated breast shapes should look similar to the breasts within the target population.\nAs a first step towards combining physical \\textit{and} statistical plausibility of simulated breast surgery outcomes, we propose to use 3D statistical shape models built from natural-looking female breasts.\nIn addition, by introducing such models into the breast shape domain, many of the aforementioned applications from other domains could be transferred to the breast as we will exemplary show later in this article.\n\nTo this end, this paper introduces the first publicly available 3D statistical shape model of the female breast built from 110 breast scans.\nTogether with the model, we present a fully automated, pairwise registration pipeline especially tailored for 3D breast scans and its application in the context of statistical shape modeling.\nOur method is computationally efficient and requires only four landmarks to guide the registration process.\n\n\\subsection{Challenges}\nCompared to shape modeling of most parts of the human body, building a statistical shape model for the female breast imposes some new challenges as discussed in the following.\n\n\\textbf{Data acquisition.}\nFirstly, acquiring a sufficient amount of high-quality training data is challenging.\nUsually, 3D scanning and \\textit{manual} landmark detection is an uncomfortable situation for the participants, in which their upper body is required to be naked.\nMoreover, landmarks can be identified only through palpation and by using a regular tape measure, both requiring a physical examination in a clinical environment.\nIn addition, during the \\textit{whole} examination, a specified posture needs to be held fixed ensuring a similar pose across all subjects.\nThis can be very exhausting, especially if 3D breast scans are taken in a standing position in which both arms should ideally be held away from the body in order to capture the breast as isolated as possible.\nAs a result, 3D scanning protocols used in clinical practice are often designed to be carried out relatively fast, thus lacking necessary precision for pose standardization, see Figure \\ref{fig:database_samples}.\nNote that the problem of only \\textit{quasi-similar} postures was also recently observed by \\citet{Maz21}.\nAll in all, the aforementioned factors definitely hinder the implementation of large-scale, high-quality data surveys.\n\n\\begin{figure}[b]\n\\includegraphics{.\/images\/1_database_samples}\n\\caption{Three typical 3D breast scans sampled from our database.\nAlthough a common pose was declared during data acquisition, a lot of pose variations are still present (indicated through a skeleton drawn in red). \nThese mainly emerge from the arms and shoulders.}\n\\label{fig:database_samples}\n\\end{figure}\n\n\\begin{figure}[t]\n\\includegraphics[width=88.5mm]{.\/images\/2_anatomy}\n\\caption{A brief overview of common landmarks and key anatomical structures of the female breast and thorax. \nThe illustration was adapted from \\citet{Hwa15}.}\n\\label{fig:anatomy}\n\\end{figure}\n\n\\textbf{Correspondence estimation.}\nSecondly, establishing correspondence among 3D breast scans by means of surface registration (rigid and non-rigid) is difficult due to the lack of reliable landmarks.\nIn essence, only four valid landmarks can be used for non-rigid registration.\nThese include both nipples as well as both lower breast poles (throughout this work defined as \\textit{the lowest (most caudal) point of the breast}, see Figure \\ref{fig:anatomy}). \nAnatomical landmark points outside the breast region (such as the sternal notch or processus coracoideus) cannot be used for non-rigid registration purposes as they would recover undesired shape variations during statistical shape modeling.\nThis way, the processus coracoideus disqualifies as its position depends strongly on the position of the arms and shoulders.\nOn the other hand, the xiphoid, located at the center of the thorax, \\textit{could} technically be used for registration. \nHowever, it cannot be reliably and consistently located across a wide range of differently-shaped breasts because of its dorsal tilt, effectively hindering the identification of a unique point.\nAnother fundamental problem is the complete lack of robust landmarks on the lateral part of the breast. \nAlthough \\citet{Har20} defined a lateral breast pole as the orthogonal intersection between the anterior axillary line and a line passing through the nipple, this point cannot be used for non-rigid registration as its position is also affected by the pose of the arms and shoulders.\nFinally, already the initial, rigid alignment of 3D breast scans is challenging due to the lack of reliable landmarks that do not undergo large soft tissue deformations.\n\n\\textbf{Non-separability of breast and thorax.} \nLast but most importantly, the region of interest, i.e. the breast, obviously cannot be well separated from the rest of the thorax when considering 3D breast scans.\nThis is primarily because the chest wall separating the breast from the thorax cannot be captured using 3D surface scanning devices, but also due to the reason that no commonly accepted and \\textit{exact} definition of the breast contour exists, see e.g. \\citet{Lot21}. \nA statistical shape model built from 3D breast scans will therefore \\textit{necessarily} capture also shape variations of the thorax, even after pose standardization.\nIn particular, these include morphological shape variations of the underlying chest wall and upper abdomen, but also those emerging from the arms, armpits, and shoulders due to improper pose standardization (also seen from Figure \\ref{fig:database_samples}).\nIf not reduced to a minimum, this will cause the following unwanted effect.\nSince breast and thorax shapes are tightly correlated in the subspace spanned by the model, the range of representable breast shapes is limited.\nThe reason is that a considerably large part of the subspace accounts for the unwanted shape variations of the surrounding regions.\nHence, the breast area should be decoupled from the thorax as much as possible in order to build an expressive and well-performing statistical shape model.\n\n\\subsection{Contributions}\nAs key contribution, this work presents the \\textit{Regensburg Breast Shape Model} (RBSM) -- a 3D statistical shape model of the female breast.\nIn order to weaken the strong coupling between the breast and surrounding regions, we propose to minimize the \\textit{variance} outside the breast region as much as possible.\nTo achieve this goal, a novel concept called \\textit{breast probability masks} (BPMs) is introduced. \nA BPM assigns probabilities to each point of a 3D breast scan, telling how \\textit{likely} it is that a particular point belongs to the breast region. \nLater, during pairwise registration, we use the BPMs to align the template to the target as accurately as possible \\textit{inside} the breast region and only roughly outside. \nThis way, only the most prominent and global shape variations outside the breast region will be recovered, effectively reducing the unwanted variance in these areas to a minimum. Figure \\ref{fig:overview_approach} illustrates this idea.\n\nTo summarize, the contributions of this paper are three-fold:\n\\begin{itemize}\n\\item We introduce the \\textit{Regensburg Breast Shape Model} (RBSM) -- an open-access 3D statistical shape model of the female breast built from 110 breast scans. It is available at \\url{https:\/\/www.rbsm.re-mic.de\/}.\n\\item We propose a fully automated, pairwise registration pipeline used to establish correspondence among our 3D breast scans. It uses BPMs to decouple the breast from the surrounding regions as much as possible and requires only four landmarks to guide the registration process.\n\\item We present two exemplary applications demonstrating how the RBSM can be used for surgical outcome simulation and the prediction of a missing breast from the remaining one.\n\\end{itemize}\nThe remainder of this paper is organized as follows: Section \\ref{sec:related_work} briefly reviews some related work. \nSection \\ref{sec:methodology} describes the entire model building pipeline used to construct the RBSM. \nIn particular, it formally introduces the notion of a BPM, followed by a detailed description of the proposed registration pipeline.\nSection \\ref{sec:evaluation} presents an extensive evaluation of the RBSM in terms of the common metrics compactness, generalization, and specificity. \nUsing the RBSM, two exemplary applications are showcased in Section \\ref{sec:applications}.\nFinally, Section \\ref{sec:discussion} discusses the results whereas Section \\ref{sec:conclusion} concludes this article.\n\n\n\\section{Related work}\n\\label{sec:related_work}\nIn this section, we briefly summarize some related work concerning breast surgery simulation, statistical shape models of the female breast, and popular techniques for pairwise surface registration used within the context of statistical shape modeling in general. \nNote that we have limited our review to the 3D case. \n\n\\textbf{Breast surgery simulation.}\nMost of the existing methods for pre-operative breast surgery simulation are designed to simulate alloplastic, implant-based breast augmentation procedures, either for aesthetic reasons or after mastectomy as part of BRS.\nTypically, those methods first generate a patient-specific, geometric representation of the breast (using tetrahedral meshes, for instance). \nAfterwards, the soft tissue deformation caused by implant insertion is simulated using a geometric and biomechanical model of the implant and breast, respectively.\n\n\\begin{figure}[t]\n\\includegraphics[]{.\/images\/3_overview_approach}\n\\caption{The proposed concept of BPMs (top row) allows to minimize unwanted shape variations of the thorax by registering a template surface as precisely as possible \\textit{inside} the breast region and only roughly outside. \nThe breast region is defined by BPMs in a probabilistic manner (top left, red areas correspond to a high probability of belonging to the breast region). \nThis simple yet effective strategy decouples the breast from the surrounding regions by reducing the variance outside the breast area. \nIn the last column, the per-vertex variance over the whole dataset is visualized on the resulting mean shape. \nThe regions showing the highest variance (red) are almost coincident with the breasts in our proposed BPM-based approach.\nContrary, without BPMs (bottom row), a lot of unwanted variance is present in the surrounding regions, especially around the arms, armpits, shoulders, and upper abdomen. \nThis implies a strong coupling between breast and thorax in the final statistical shape model.}\n\\label{fig:overview_approach}\n\\end{figure}\n\nAs such, \\citet{Roo06} used the tensor-mass model introduced by \\citet{Cot00}, a combination of classical finite element and mass-spring models, for implant-based breast augmentation planning.\n\\citet{Cie12} proposed a web-based tool for breast augmentation planning which requires only 2D photographs and anthropometric measurements as input and allows the user to choose from a variety of different implants. \nTheir method automatically reconstructs a 3D breast model and subsequently applies a tissue elastic model closely resembling the finite element model.\n\\citet{Geo14} utilizes patient-specific finite element models generated from 3D breast scans. \nCombined with a novel mechanism called \\textit{displacement template}, geometric implant models are no longer required, thus breaking up the coupling between implant and enclosing breast.\n\n\\begin{figure*}[t!]\n\\centering\n\\includegraphics{.\/images\/4_overview_pipeline}\n\\caption{An overview of the pipeline used to build the RBSM. \nWe first establish dense correspondence by means of pairwise surface registration. \nBased on BPMs, our method starts by rigidly aligning the template to the target. \nSubsequently, a non-rigid alignment is applied in a hierarchical, multi-resolution fashion. \nIn this step, BPMs are used to precisely recover only shape variations of the breast.\nFinally, we perform classical \\textit{Generalized Procrustes Analysis} (GPA) and \\textit{Principal Component Analysis} (PCA) to build the model.}\n\\label{fig:overview_pipeline}\n\\end{figure*}\n\nBesides the simulation of breast augmentation procedures using implants, some methods exist especially addressing the simulation of BRS \\textit{without} implant insertion, for example by means of autologous fat tissue.\nBased on Pascal's principle and volume conservation, \\citet{Cos01} developed a novel approach for real-time physically based simulation of deformable objects, called \\textit{Long Elements Method} (LEM).\nAn extension of LEM, known as the \\textit{Radial Elements Method} \\citep{Bal03}, was later used by \\citet{Bal06} for cosmetic and reconstructive breast surgery simulation.\n\\citet{Wil03} employ a finite element approach incorporating the Mooney-Rivlin hyperelastic material model for the realistic simulation of soft tissue to simulate \\textit{transverse rectus abdominis myocutaneous} flap breast reconstruction.\n\nFor most of the existing methods, however, various authors highlighted the disagreement between simulated and actual outcomes \\citep{Cru15,Roo06} or only partially satisfying results for certain types of breast shapes \\citep{Vor17}.\n\n\\textbf{Statistical shape models of the female breast.}\nLiterature about 3D statistical shape models of the female breast is sparse.\nIn the early work of \\citet{Seo07}, a 3D statistical shape model was built from 28 breast scans with the goal of analyzing breast volume and surface measurements.\nHowever, they assume symmetric breasts obtained by simply mirroring the right breast.\nTo date and to the best of our knowledge, this is the only work primarily addressing the construction of a statistical shape model from 3D breast scans, which is thus closest to our work.\nBesides, \\citet{Rui18} utilized a 3D statistical shape model built from 310 breast scans for the validation of a novel weighted regularized projection method used for 3D reconstruction.\nAs their focus did not lie on the construction of a well-performing statistical shape model of the breast, they did not provide detailed information about the registration method used to establish correspondence, the training data nor a comprehensive evaluation in terms of common metrics.\n\n\\textbf{Pairwise surface registration.}\nDuring the last few decades, countless algorithms were proposed tackling the problem of (pairwise) non-rigid surface registration.\nTo date, however, none of them were used for 3D breast scan registration.\n\nOne of the most widely used methods is the \\textit{Optimal Step Non-rigid Iterative Closest Point} (NICP) framework proposed by \\citet{Amb07} and based on the work of \\citet{All03}. \nNICP is the only method already used in the female breast shape domain to reconstruct a 3D breast model from a sequence of depth images \\citep{Lac19}.\nMoreover, NICP was employed to construct the BFM and LSFM.\nA second class of non-rigid registration methods is based on splines, such as \\textit{Thin Plate Splines} (TPS) or B-splines.\nPioneered by \\citet{Boo89}, TPS were utilized by \\citet{Pau02} to construct a statistical shape model of the human ear canal.\nB-splines are extensively used in the \\textit{free-form deformations} framework and, among others, used for the creation of shape models of the human heart \\citep{Ord07}.\nA third, recently introduced framework is based on \\textit{Gaussian Process Morphable Models} (GPMMs) introduced by \\citet{Lue18}.\nGPMMs are statistical shape models themselves generalizing classical point-based models as proposed by \\citet{Coo95}.\nBy means of GPMMs, expected deformations can be modeled using analytic covariance functions and later used as a \\textit{prior} for non-rigid surface registration, thus effectively reducing the search space. \nThis approach was successfully used by \\citet{Ger17} to build an improved version of the BFM.\nLastly, pairwise non-rigid surface registration algorithms are built upon the \\textit{as-rigid-as-possible} (ARAP) real-time mesh deformation framework proposed by \\citet{Sor07}.\nThis method was successfully transferred to the registration domain, yielding to similar methods constraining deformations to be \\textit{as conformal as possible} \\citep{Yos14} or \\textit{as similar as possible} \\citep{Jia17,Yam13}. \nRecently, a variant of ARAP was utilized by \\citet{Dai20} to built a shape model of the full human head. \n\n\\section{Methodology}\n\\label{sec:methodology}\nThis section describes the entire pipeline used to build the RBSM. \nAs outlined in Figure \\ref{fig:overview_pipeline}, we start by establishing correspondence among our training data.\nBased on a novel concept called \\textit{breast probability masks} (Section \\ref{subsec:bpms}), this is achieved by means of a fully automated, pairwise registration pipeline as proposed in Section \\ref{subsec:registration}.\nFinally, we follow the typical workflow used to build a point-based statistical shape model by applying a \\textit{Generalized Procrustes Analysis} and a \\textit{Principal Component Analysis} to the registered data set (briefly summarized in Section \\ref{subsec:model_building}).\n\nIn what follows, 3D breast scans are represented using triangular surface meshes.\nA triangle mesh $\\mathcal{M}=(V,E,\\mathcal{P})$ is fully specified by a set of $n$ vertices $V\\subset\\mathbb{N}$, edges $E\\subset V\\times V$, and an embedding $\\mathcal{P}=\\{\\mathbf{p}_1,\\mathbf{p}_2\\ldots,\\mathbf{p}_n\\}\\subset\\mathbb{R}^3$. \nSometimes, however, instead of arranging points $\\mathbf{p}_i$ into a set, it is more convenient to use a matrix notation $\\mathbf{P}=(\\mathbf{p}_1,\\mathbf{p}_2,\\ldots,\\mathbf{p}_n)^\\top\\in\\mathbb{R}^{n\\times 3}$.\nHence, we will denote a triangle mesh either as $\\mathcal{M}=(V,E,\\mathcal{P})$ or equivalently as $\\mathcal{M}=(V,E,\\mathbf{P})$.\n\n\\subsection{Breast probability masks}\n\\label{subsec:bpms}\nGiven a 3D breast scan represented as triangle mesh $\\mathcal{M}=(V,E,\\mathcal{P})$, we call\n\\begin{equation}\np_\\mathcal{M}:\\mathcal{P}\\longrightarrow(0,1]\n\\end{equation}\na \\textit{breast probability mask} (BPM). Technically, a BPM is a scalar field defined over $\\mathcal{M}$ assigning each point $\\mathbf{p}_i$ of a 3D breast scan a probability $p_\\mathcal{M}(\\mathbf{p}_i)$ telling how \\textit{likely} it is that $\\mathbf{p}_i$ belongs to the breast region.\n\n\\textbf{Concrete mapping.} As a concrete mapping for $p_\\mathcal{M}$ we propose to use a normalized sum of \\textit{elliptical basis functions} (EBFs), centered at the nipples.\nWe use EBFs instead of ordinary \\textit{radial basis functions} (RBFs) because we found that they better capture the natural teardrop shape of the breast (see Figure \\ref{fig:comparison_bpms} for a comparison between RBFs and EBFs).\nTechnically, EBFs are a generalization of RBFs using the Mahalanobis distance instead of an ordinary vector norm.\nFormally, an EBF $\\phi:[0,\\infty)\\longrightarrow\\mathbb{R}$ centered at a point $\\mathbf{c}\\in\\mathbb{R}^n$ is of the form $\\phi(\\mathbf{x})=\\phi\\left(d_M\\left(\\mathbf{x},\\mathbf{c}\\right)\\right)$. \nHere, $d_M$ is the Mahalanobis distance, defined as\n\\begin{equation}\nd_M(\\mathbf{x},\\mathbf{c})\\coloneqq\\sqrt{\\left(\\mathbf{x}-\\mathbf{c}\\right)^\\top\\mathbf{S}^{-1}\\left(\\mathbf{x}-\\mathbf{c}\\right)}\\,,\n\\end{equation}\nwhere $\\mathbf{S}\\in\\mathbb{R}^{n\\times n}$ is a symmetric positive definite matrix, also called \\textit{covariance matrix}.\nTo stress that the Mahalanobis distance depends on $\\mathbf{S}$, we write $d_M(\\mathbf{x},\\mathbf{c};\\mathbf{S})$ in the following.\n\nNow, in order to define a concrete BPM using EBFs, let $\\mathbf{p}_\\text{N}^\\tau\\in\\mathcal{P}$ denote the position of the left (L) and right (R) nipple, respectively, and $\\tau\\in\\{\\text{L, R}\\}$.\nWe first construct two individual probability masks for the left and the right breast, given as\n\\begin{equation}\np_\\mathcal{M}^\\tau(\\mathbf{p}_i)=\\phi\\left(d_M\\left(\\mathbf{p}_i,\\mathbf{p}_\\text{N}^\\tau;\\mathbf{S}_\\tau\\right)\\right).\n\\end{equation}\nHereby, we define $\\phi:[0,\\infty)\\longrightarrow(0,1]$ as \n\\begin{equation}\n\\phi(x)=\\exp\\left(-x^2\\right).\n\\end{equation}\nFinally, the BPM for a whole 3D breast scan is given as the normalized sum\n\\begin{equation}\np_\\mathcal{M}(\\mathbf{p}_i)=\\frac{1}{4}\\left(p_\\mathcal{M}^\\text{L}(\\mathbf{p}_i)+\\hat{p}_\\mathcal{M}^\\text{L}(\\mathbf{p}_i)+p_\\mathcal{M}^\\text{R}(\\mathbf{p}_i)+\\hat{p}_\\mathcal{M}^\\text{R}(\\mathbf{p}_i)\\right),\n\\end{equation}\nwhere \n\\begin{equation}\n\\hat{p}_\\mathcal{M}^\\tau(\\mathbf{p}_i)=\\phi\\left(d_M\\left(\\mathbf{p}_i,\\mathbf{\\hat{p}}_\\text{N}^\\tau;\\mathbf{\\hat{S}}_\\tau\\right)\\right)\n\\end{equation}\nare shifted BPMs of the left and right breast added to better mimic the teardrop shape, and $\\mathbf{\\hat{p}}_\\text{N}^\\tau=\\mathbf{p}_\\text{N}^\\tau+\\mathbf{t}_\\tau$ with translation vectors $\\mathbf{t}_\\tau\\in\\mathbb{R}^3$.\n\n\\begin{figure}[h!]\n\\includegraphics[]{.\/images\/5_comparison_bpms}\n\\caption{From left to right: comparison between RBF, EBF, and a sum of two EBFs, illustrated as contour plots. \nWhile a simple RBF or EBF is not able to accurately mimic the typical teardrop shape of the breast, a sum of two EBFs comes close.}\n\\label{fig:comparison_bpms}\n\\end{figure}\n\n\\textbf{Parameter selection.} In order to fully define a BPM, appropriate matrices $\\mathbf{S}_\\tau,\\mathbf{\\hat{S}}_\\tau\\in\\mathbb{R}^{3\\times 3}$ and translation vectors $\\mathbf{t}_\\tau\\in\\mathbb{R}^3$ need to be chosen first.\nAs such, a total of 30 values are required to be properly determined (six for each $\\mathbf{S}_\\tau$ and $\\mathbf{\\hat{S}}_\\tau$, and three for each $\\mathbf{t}_\\tau$). \nTo simplify that task, we assume diagonal covariance matrices and utilize previously marked landmarks on the 3D breast scans.\nSpecifically, denote the landmark points shown in Figure \\ref{fig:anatomy} as $\\mathbf{p}_\\text{SN},\\mathbf{p}_\\text{XI}\\in\\mathcal{P}$ for sternal notch and xiphoid, and $\\mathbf{p}_\\text{LaBP}^\\tau,\\mathbf{p}_\\text{LBP}^\\tau\\in\\mathcal{P}$ for left and right lateral and lower breast pole, respectively.\nWe then define\n\\begin{equation}\n\\begin{split}\n\\mathbf{S}_\\tau&=\n\\!\\begin{multlined}[t][60mm]\n\\frac{1}{2}\\diag\\Bigl(d_G\\left(\\mathbf{p}^\\tau_\\text{LaBP},\\mathbf{p}^\\tau_\\text{N}\\right)+d_G\\left(\\mathbf{p}^\\tau_\\text{N},\\mathbf{p}_\\text{XI}\\right),\\\\\nd_G\\left(\\mathbf{p}^\\tau_\\text{N},\\mathbf{p}^\\tau_\\text{LBP}\\right), d_G\\left(\\mathbf{p}^\\tau_\\text{LaBP},\\mathbf{p}^\\tau_\\text{N}\\right)\\Bigr)\\,,\n\\end{multlined}\\\\\n\\mathbf{\\hat{S}}_\\tau&=\n\\!\\begin{multlined}[t][60mm]\n\\frac{1}{2}\\diag\\Bigl(d_G\\left(\\mathbf{p}^\\tau_\\text{LaBP},\\mathbf{p}^\\tau_\\text{N}\\right)+d_G\\left(\\mathbf{p}^\\tau_\\text{N},\\mathbf{p}_\\text{XI}\\right),\\\\\nd_G\\left(\\mathbf{p}^\\tau_\\text{N},\\mathbf{p}_\\text{SN}\\right),d_G\\left(\\mathbf{p}^\\tau_\\text{LaBP},\\mathbf{p}^\\tau_\\text{N}\\right)\\Bigr)\\,,\n\\end{multlined}\\\\\n\\mathbf{t}_\\tau&=\\mathbf{p}^\\tau_\\text{N}+\\frac{1}{5}\\left(0,d_G\\left(\\mathbf{p}^\\tau_\\text{N},\\mathbf{p}_\\text{SN}\\right),0\\right),\n\\end{split}\n\\end{equation}\nwhere $d_G$ denotes the Geodesic distance between two points on the surface mesh.\nNote that $\\mathbf{S}_\\tau$ and $\\mathbf{\\hat{S}}_\\tau$ differ only in the second diagonal element.\n\n\\subsection{Registration of 3D breast scans}\n\\label{subsec:registration}\nFollowing Figure \\ref{fig:overview_pipeline}, the proposed pairwise registration pipeline is mainly composed of rigid alignment (Section \\ref{subsubsec:rigid}) and non-rigid alignment (Section \\ref{subsubsec:non-rigid}). \nFollowing common practice, the latter is carried out in a hierarchical, multi-resolution fashion (Section \\ref{subsubsec:multi_res_fitting}).\n\nBoth phases make extensive use of BPMs in order to align a template surface $\\mathcal{S}=(V,E,\\mathbf{P})$ to a target $\\mathcal{T}$ as accurately as possible \\textit{inside} the breast region and only roughly outside, effectively decoupling the breast from the rest of the thorax by reducing the \\textit{variance} outside the breast region to a minimum.\nThis is justified as the covariance $\\cov(x,y)$ becomes smaller if $\\var(x)$ or $\\var(y)$ is lowered, following from the well known fact that $\\left|\\cov(x,y)\\right|\\leq\\sqrt{\\var(x)}\\sqrt{\\var(y)}$ (which holds via the Cauchy\u2013Schwarz inequality).\n\nFinally, note that the target surface $\\mathcal{T}$ can be given in any representation that allows for closest point search.\nWe use a triangular surface mesh but write $\\mathcal{T}\\subset\\mathbb{R}^3$ for the sake of notational simplicity.\n\n\\subsubsection{Rigid alignment}\n\\label{subsubsec:rigid}\nThe overall goal of the rigid alignment is to move the template as close as possible to the \\textit{rigid part} of the target, which we define as \\textit{the thorax without the breast}.\nIn particular, we expect that the thoraxes of two subjects without the breast region can be sufficiently well aligned if we assume the breast to be the only part of the thorax that deforms non-rigidly.\nBased on this assumption, the absence of suitable landmarks, and due to the fact that our initial 3D breast scans are already reasonably well aligned (see Section \\ref{subsec:data}) we propose a modified version of the \\textit{ Iterative Closest Point} (ICP) algorithm, originally introduced by \\citet{Bes92}.\n\nEssentially, compared to the standard version of the ICP algorithm, our modified version differs in the following three aspects: (i) a scaling factor is added to the rigid transformation effectively allowing for Euclidean similarity transformations \\citep{Du07,Zin05}. \nSecondly, (ii) to ensure that only the rigid parts of the 3D breast scans are used for alignment, correspondences, where both points have a high probability belonging to the breast region, are discarded.\nThis is implemented by thresholding the template and target BPMs.\nFinally, (iii) rotations are restricted to the $x$-axis corresponding to the transversal plane.\nRotations around the $y$- and $z$-axis (sagittal and coronal plane), possibly introduced due to severe overweight in conjunction with an uneven distribution of abdominal fat could destroy the initial alignment and lead to misalignment.\nIn any case, asymmetries introduced due to the thorax should \\textit{not} affect the rigid alignment of the template.\n\n\\subsubsection{Non-rigid alignment}\n\\label{subsubsec:non-rigid}\nGiven the rigidly aligned template $\\mathcal{S}=(V,E,\\mathbf{P})$, the goal of the non-rigid alignment is to gradually deform $\\mathcal{S}$ into a new surface $\\mathcal{S}'=(V,E,\\mathbf{P}')$ with identical topology such that $\\mathcal{S}'$ is as close as possible to the target $\\mathcal{T}$ \\textit{inside} the breast region.\nFollowing various authors including \\citet{Jia17} and \\citet{Yam13}, we formulate our non-rigid registration problem using the following non-linear energy functional\n\\begin{equation}\nF\\left(\\mathbf{P}'\\right)=F_D\\left(\\mathbf{P}'\\right)+\\alpha F_R\\left(\\mathbf{P}'\\right)+\\beta F_L\\left(\\mathbf{P}'\\right),\n\\label{eq:non_rigid_cost}\n\\end{equation}\nwhere $F_D$ is a distance term used to penalize the point-to-point distance between the template and target surface, $F_R$ is a regularization term constraining deformations \\textit{as similar as possible}, and $F_L$ constitutes a landmark term ensuring certain points to be matched.\n$\\alpha,\\beta\\geq 0$ are weights controlling the individual contribution of each term to the cost function.\nMinimizing $F$ finally leads to the new points $\\mathbf{P}'$ of the deformed template surface $\\mathcal{S}'$, i.e.\n\\begin{equation}\n\\mathbf{P}'=\\argmin_{\\mathbf{P'}\\in\\mathbb{R}^{n\\times 3}}F(\\mathbf{P'}).\n\\label{eq:non_rigid_opt_problem}\n\\end{equation}\nAdapting the strategy proposed by \\citet{All03}, instead of computing (\\ref{eq:non_rigid_opt_problem}) only once, we minimize $F$ several times but each time lowering the regularization weight $\\alpha$ in (\\ref{eq:non_rigid_cost}).\nAs later demonstrated by \\citet{Amb07}, this strategy is able to recover the whole range of global and local non-rigid deformations efficiently.\nFollowing various authors \\citep{Jia17,Sor07}, the optimization problem in (\\ref{eq:non_rigid_opt_problem}) is solved using an alternating minimization (AM) approach as briefly summarized in \\ref{app:solve_am}.\n\n\\textbf{Distance term.}\nThe distance term $F_D$ is used to attract the template $\\mathcal{S}$ to the target $\\mathcal{T}$.\nAssuming fixed correspondences between both surfaces, i.e. $\\left\\{(\\mathbf{p}_1,\\mathbf{q}_1),(\\mathbf{p}_2,\\mathbf{q}_2),\\ldots,(\\mathbf{p}_n,\\mathbf{q}_n)\\right\\}$ with $\\mathbf{q}_i\\in\\mathcal{T}$ being the closest point to $\\mathbf{p}_i$, the distance term can be written as\n\\begin{equation}\nF_D(\\mathbf{P}')=\\frac{1}{2}\\left\\Vert\\mathbf{C}\\left(\\mathbf{P}'-\\mathbf{Q}\\right)\\right\\Vert^2_F,\n\\end{equation}\nwhere $\\mathbf{C}\\coloneqq\\text{diag}(c_1,c_2,\\ldots,c_n)$, $c_i\\geq 0$ for all $i\\in\\left\\{1,2,\\ldots,n\\right\\}$ are weights used to quantify the reliability of a match, and $\\mathbf{Q}\\coloneqq\\left(\\mathbf{q}_1,\\mathbf{q}_2,\\ldots,\\mathbf{q}_n\\right)^\\top\\in\\mathbb{R}^{n\\times 3}$.\nUsing the BPMs $p_\\mathcal{S}$ and $p_\\mathcal{T}$ of the template and target, we set\n\\begin{equation}\nc_i=\\frac{p_\\mathcal{S}(\\mathbf{p}_i)+p_\\mathcal{T}(\\mathbf{q}_i)}{2}.\n\\end{equation}\nThis way, correspondences $(\\mathbf{p}_i,\\mathbf{q}_i)$ mapping from one breast region to the other have a greater impact on the overall distance term as $c_i\\in(0,1]$ becomes large in this case.\nConversely, the influence tends to zero if $c_i\\rightarrow 0$, i.e. if both points are less likely to belong to the breast region.\nAs such, the deformation of points $\\mathbf{p}_i$ on the template with a small value for $c_i$ is mainly controlled by the regularization term, as previously described by \\citet{All03}.\n\n\\textbf{Regularization term.}\nThe regularization term $F_R$ should prevent the template surface from shearing and distortion while simultaneously ensuring structure preservation and smooth deformations.\nTo do so, we adapt the \\textit{consistent as-similar-as-possible} (CASAP) regularization technique in which deformations are constrained to be \\textit{locally} as similar as possible \\citep{Jia17,Yam13}.\nSpecifically, given a local neighborhood $E_i\\subset E$ around each point $\\mathbf{p}_i$, the template surface is only allowed to move in terms of an Euclidean similarity transformation\n\\begin{equation}\n\\mathbf{p}'_j-\\mathbf{p}'_k=s_i\\mathbf{R}_i\\left(\\mathbf{p}_j-\\mathbf{p}_k\\right)\\qquad\\forall(j,k)\\in E_i,\n\\label{eq:asap}\n\\end{equation}\nwhere $s_i>0$ is a scaling factor and $\\mathbf{R}_i\\in\\SO(3)$ a rotation matrix.\nFollowing \\citet{Cha10}, we define $E_i$ as \\textit{the set containing all (directed) edges of triangles incident to} $\\mathbf{p}_i$, also known as \\textit{spokes-and-rims}.\nFinally, our CASAP regularization term reads\n\\begin{multline}\nF_R(\\mathbf{P}')=\\frac{1}{2}\\sum_{i=1}^n w_i\\left[\\sum_{(j,k)\\in E_i}w_{jk}\\left\\Vert\\left(\\mathbf{p}'_j-\\mathbf{p}'_k\\right)-s_i\\mathbf{R}_i\\left(\\mathbf{p}_j-\\mathbf{p}_k\\right)\\right\\Vert^2_2+\\right.\\\\\n\\left.\\lambda\\sum_{l\\in N_i}w_{il}\\left\\Vert\\mathbf{R}_i-\\mathbf{R}_l\\right\\Vert^2_F\\right]\\,,\n\\label{eq:casap_regularization}\n\\end{multline}\nwhere weights $w_i>0$ are added to individually control the amount of regularization for each particular point.\nAs mentioned above, since the deformation of points $\\mathbf{p}_i$ with a small value for $c_i$ is mainly controlled by the regularization term, we define\n\\begin{equation}\nw_i=\\frac{1}{(h-1)c_i+1}\\qquad\\text{with}\\qquad\\frac{1}{h}\\leq w_i<1\n\\end{equation}\nfor all $i\\in\\{1,2,\\ldots,n\\}$ and $h\\in\\mathbb{N}^+$ (we used $h=2$ throughout this paper).\nAs seen, this strategy keeps points $\\mathbf{p}_i$ of the template relatively stiff if (i) $\\mathbf{p}_i$ has a low probability belonging to the breast region and (ii) if the corresponding point on the target is also not likely to be part of the breast region (because $w_i\\rightarrow 1$ if $c_i\\rightarrow 0$), thus effectively preventing the template from adapting too close to the target outside the breast region.\nLastly, $N_i\\subset V$ in (\\ref{eq:casap_regularization}) denotes the one-ring neighborhood of the $i$-th point and $w_{jk}\\in\\mathbb{R}$ are the popular cotangent weights, see e.g. \\citet{Bot10}.\n$\\lambda\\geq 0$ is usually set to $0.02A$, where $A\\geq 0$ is the total surface area of $\\mathcal{S}$ \\citep{Lev14}.\n\n\\textbf{Landmark term.}\nThe goal of the landmark term $F_L$ is to keep certain positions (i.e. landmarks) fix during the registration process.\nLet $I\\subset\\mathbb{N}$ be an index set containing the indices of the $m$ landmarks specified on the template surface $\\mathcal{S}$.\nDefine a matrix $\\mathbf{D}\\in\\mathbb{R}^{m\\times n}$ as\n\\begin{equation}\n\\mathbf{D}=(d_{ij}):=\n\\begin{cases}\n1, & \\text{if }j\\in I,\\\\\n0, & \\text{otherwise}\n\\end{cases}\n\\end{equation}\nfor $i=1,2,\\ldots,m$ and $j=1,2,\\ldots,n$.\nNext, denote the corresponding landmarks on the target surface by $\\{\\mathbf{q}_1,\\mathbf{q}_2,\\ldots,\\mathbf{q}_m\\}\\subset\\mathcal{T}$. \nThen, the landmark term is defined as\n\\begin{equation}\nF_L(\\mathbf{P}')=\\frac{1}{2}\\left\\Vert\\mathbf{DP}'-\\mathbf{Q}_L\\right\\Vert^2_F,\n\\end{equation}\nwhere $\\mathbf{Q}_L\\coloneqq\\left(\\mathbf{q}_1,\\mathbf{q}_2,\\ldots,\\mathbf{q}_m\\right)^\\top\\in\\mathbb{R}^{m\\times 3}$.\n\n\\subsubsection{Multi-resolution fitting strategy}\n\\label{subsubsec:multi_res_fitting}\nFollowing common practice, instead of applying the previously described non-rigid alignment only once, we employ a hierarchical, multi-resolution fitting strategy composed of initial fitting, coarse fitting, and fine fitting (see also Figure \\ref{fig:overview_pipeline}).\n\n\\textbf{Initial fitting.}\nHaving a low-resolution instance of the rigidly aligned template at hand, the goal of the initial fitting is to roughly adapt the coarse template to the key features (i.e. landmarks) of the target.\nTo do so, we strictly prioritize the landmark constraints and do not use BPMs in this phase.\n\n\\textbf{Coarse fitting.}\nIn this step, the initially fitted low-resolution template is gradually deformed towards the target.\n\n\\textbf{Upsampling.}\nNext, the deformation obtained from the previous step is applied to the original, full-resolution template. \nThis is achieved using a concept called \\textit{Embedded Deformation}, introduced by \\citet{Sum07}.\nIn essence, the deformation of the coarse template obtained from the previous step is transferred to the template by linearly interpolating the transformation at each point.\n\n\\textbf{Fine fitting.}\nLastly, the upsampled template is fitted to the target again to produce the final result. \n\n\\subsection{Model building}\n\\label{subsec:model_building}\nOnce the data set is brought into correspondence, we follow the typical workflow used to build a classical point-based statistical shape model as proposed by \\citet{Coo95}.\nFor notational simplicity, instead of stacking points $\\mathcal{P}$ of a triangular mesh $\\mathcal{M}=(V,E,\\mathcal{P})$ into a matrix $\\mathbf{P}\\in\\mathbb{R}^{n\\times 3}$ as before, we use a vectorized representation, denoted as $\\mathbf{x}=\\vecz(\\mathbf{P})\\in\\mathbb{R}^{3n}$ in the following.\n\nBriefly, given a set of $k$ breast scans $\\{\\mathbf{x}_1,\\mathbf{x}_2,\\ldots,\\mathbf{x}_k\\}\\subset\\mathbb{R}^{3n}$ in correspondence, we first perform a \\textit{Generalized Procrustes Analysis} (GPA) as introduced by \\citet{Gow75}.\nGPA iteratively aligns the objects to the arithmetic mean $\\mathbf{\\bar{x}}\\in\\mathbb{R}^{3n}$ (successively estimated from the data) by using an Euclidean similarity transformation, effectively transforming the objects into the shape space.\nSecondly, a \\textit{Principal Component Analysis} (PCA) is carried out on the Procrustes-aligned shapes.\nLet $\\{\\lambda_1,\\lambda_2,\\ldots,\\lambda_q\\}\\subset\\mathbb{R}^+$ be the $q0$,\nwhere $a$ is a negative parameter as expected by general discussions (given in the last section) and\n$\\Pi_1$ is positive (see the expression (\\ref{def_Pi1})). The momentum-dependence of the $U_0$ is characterized by introducing the nucleon effective k-mass\\,\\cite{LiBA2018PPNP,LiBA15,LiBA16}, i.e.,\n\\begin{equation}\nM_0^{\\ast}(\\rho,|\\v{k}|)\/M=\\left[1+\\frac{M}{|\\v{k}|}\\frac{\\partial\nU_0}{\\partial|\\v{k}|}\\right]^{-1}.\n\\end{equation}\nSpecifically, we obtain for the current EDF the following expression \n\\begin{align}\n\\frac{M_0^{\\ast}(\\rho,|\\v{k}|)}{M}=\\Bigg[&1+C_{\\rm{tot}}M\\left(\\frac{\\rho}{\\rho_0}\\right)\n\\Bigg[\\frac{2a\\Pi_1}{7}\\left(\\frac{k_{\\rm{F}}}{\\Lambda^2}\\right)^{1\/2}|\\v{k}|^{-3\/2}\\notag\\\\\n&+\\frac{9b\\Pi_2}{40}\\left(\\frac{k_{\\rm{F}}}{\\Lambda^2}\\right)^{1\/3}|\\v{k}|^{-5\/3}\\Bigg]\\Bigg]^{-1},\n\\end{align}\nwhich is further evaluated at the Fermi surface $|\\v{k}|=k_{\\rm{F}}$ to give the density dependence of the $M_0^{\\ast}(\\rho)\/M$ as,\n\\begin{align}\\label{FFG-M0-mm}\n\\frac{M_0^{\\ast}(\\rho)}{M}\n=\\Bigg[&1+C_{\\rm{tot}}M\\left(\\frac{\\rho}{\\rho_0}\\right)\n\\Bigg[\\frac{2a\\Pi_1}{7}\\frac{1}{k_{\\rm{F}}\\Lambda}\\notag\\\\\n&+\\frac{9b\\Pi_2}{40}\\left(\\frac{1}{\\Lambda^2k_{\\rm{F}}^4}\\right)^{1\/3}\\Bigg]\\Bigg]^{-1}.\n\\end{align}\nIn the high density limit, we have\n\\begin{equation}\\label{def_11}\n\\lim_{\\rm{large }\\rho}\\frac{M_0^{\\ast}(\\rho)}{M}\n\\approx\\left[1+\\frac{2M\\Pi_1aC_{\\rm{tot}}}{7k_{\\rm{F}}\\Lambda}\\left(\\frac{\\rho}{\\rho_0}\\right)+\\mathcal{O}\\left(k_{\\rm{F}}^{-1\/3}\\right)\\right]^{-1},\n\\end{equation}\nwhich approaches zero in the limit $\\rho\\to\\infty$. In the FFG\nmodel, the density dependence of the effective mass is given in Eq.\\,(\\ref{FFG-M0}), i.e., by taking $\\Pi_1=1$ in (\\ref{def_11}).\n\n\n\\subsection{Symmetry Energy}\\label{sb_Esym}\n\nWe now discuss the symmetry energy, its slope parameter, and the isovector (symmetry) potential $U_{\\rm{sym}}(\\rho,|\\v{k}|)$. As derived in\ndetailed in APPENDIX \\ref{app1}, the symmetry energy for the general HMT model is given by,\n\\begin{widetext}\n\\begin{align}\\label{HMT-Esym-mm}\nE_{\\rm{sym}}(\\rho)=&\\frac{k_{\\rm{F}}^2}{6M}\\left[1+C_0(1+3C_1)\\left(5\\phi_0+\\frac{3}{\\phi_0}-8\\right)\n+3C_0\\phi_1\\left(1+\\frac{5}{3}C_1\\right)\\left(5\\phi_0-\\frac{3}{\\phi_0}\\right)+\\frac{27C_0\\phi_1^2}{5\\phi_0}\\right]\n\\notag\\\\\n&+\\frac{1}{4}C_\\ell\\Delta_0^2\\left(\\frac{\\rho}{\\rho_0}\\right)\\left[\nY_{10}+Y_{11}a\\frac{k_{\\rm{F}}}{\\Lambda}+Y_{12}b\\left(\\frac{k_{\\rm{F}}}{\\Lambda}\\right)^{2\/3}\n\\right]\n+\\frac{1}{4}C_{\\rm{u}}\\Delta_0^2\\left(\\frac{\\rho}{\\rho_0}\\right)\n\\left[\nZ_{10}+Z_{11}a\\frac{k_{\\rm{F}}}{\\Lambda}+Z_{12}b\\left(\\frac{k_{\\rm{F}}}{\\Lambda}\\right)^{2\/3}\n\\right]\\notag\\\\\n&+\\frac{3}{2}C_\\ell\\Delta_0C_0\\left(\\frac{\\rho}{\\rho_0}\\right)\\left[\nY_{20}+Y_{21}a\\frac{k_{\\rm{F}}}{\\Lambda}+Y_{22}b\\left(\\frac{k_{\\rm{F}}}{\\Lambda}\\right)^{2\/3}\n\\right]\n+\\frac{3}{2}C_{\\rm{u}}\\Delta_0C_0\\left(\\frac{\\rho}{\\rho_0}\\right)\\left[\nZ_{20}+Z_{21}a\\frac{k_{\\rm{F}}}{\\Lambda}+Z_{22}b\\left(\\frac{k_{\\rm{F}}}{\\Lambda}\\right)^{2\/3}\n\\right]\\notag\\\\\n&+\\frac{9}{4}C_\\ell C_0^2\\left(\\frac{\\rho}{\\rho_0}\\right)\\left[\nY_{30}+Y_{31}a\\frac{k_{\\rm{F}}}{\\Lambda}+Y_{32}b\\left(\\frac{k_{\\rm{F}}}{\\Lambda}\\right)^{2\/3}\n\\right]+\\frac{9}{4} C_{\\rm{u}} C_0^2\\left(\\frac{\\rho}{\\rho_0}\\right)\\left[\nZ_{30}+Z_{31}a\\frac{k_{\\rm{F}}}{\\Lambda}+Z_{32}b\\left(\\frac{k_{\\rm{F}}}{\\Lambda}\\right)^{2\/3}\n\\right]\\notag\\\\\n&+\\frac{1}{4}A_{\\rm{d}}\\left(\\frac{\\rho}{\\rho_0}\\right)-\\frac{Bx}{\\sigma+1}\\left(\\frac{\\rho}{\\rho_0}\\right)^{\\sigma},\\end{align}\n\\end{widetext}\nwith the expressions for $Y_{ij}$ and $Z_{ij}$ given in Eqs.\\,(\\ref{def_Y1}), (\\ref{def_Y20}), (\\ref{def_Y21}), (\\ref{def_Y22}),\n(\\ref{def_Y30}), (\\ref{def_Y31}), (\\ref{def_Y32}),\n (\\ref{def_Z1}), (\\ref{def_Z20}), (\\ref{def_Z21}), \n(\\ref{def_Z22}), (\\ref{def_Z30}), (\\ref{def_Z31}), and (\\ref{def_Z32}), respectively.\nThe second line of Eq.\\,(\\ref{HMT-Esym-mm}) represents the\ncontribution from the depletion in the $n_{\\v{k}}^J$ characterized by $\\Delta_0^2$, while the fourth line has the HMT contribution indicated by the\n$C_0^2$. The third line of Eq.\\,(\\ref{HMT-Esym-mm}) is the mixing of\nthe depletion and the HMT characterized by $\\Delta_0C_0$. In the FFG\nmodel, only the first two lines of Eq.\\,(\\ref{HMT-Esym-mm}) survive.\nMoreover, $A_{\\rm{d}}$ and $C_{\\rm{d}}$ are the\ndifference between the like and unlike terms,\n\\begin{equation}\\label{ad}\nA_{\\rm{d}}=A_{\\rm{d}}^0+\\frac{4xB}{1+\\sigma},~~C_{\\rm{d}}=C_\\ell-C_{\\rm{u}},\n\\end{equation}\nwith $\nA_{\\rm{d}}^0=A_\\ell^0-A_{\\rm{u}}^0$. The density structure of\nthe symmetry energy is the same as that of the SNM EOS, see Eq.\\,(\\ref{str}), thus it is straightforward to obtain the slope\nparameter $L(\\rho)$, see Eq.\\,(\\ref{HMT-L}).\n\nSimilarly, the isovector (symmetry) potential $U_{\\rm{sym}}(\\rho,|\\v{k}|)$ is defined through the following Lane potential \\,\\cite{Lan62} \n\\begin{equation}\nU_J(\\rho,\\delta,|\\v{k}|)\\approx\nU_0(\\rho,|\\v{k}|)+U_{\\rm{sym}}(\\rho,|\\v{k}|)\\tau_3^J\\delta+\\mathcal{O}(\\delta^2).\n\\end{equation}\nSuch a decomposition of single-nucleon potential in ANM has been verified by various phenomenological model analyses of nucleon-nucleus scattering data and predictions of microscopic nuclear many-body theories, see, e.g., Refs. \\cite{mu04,Li04,Ron06,Dal1,zuo05,Beh11,LiX13,LiX15}. In the HMT model, it is given by\n\\begin{align}\\label{HMT-Usym-mm}\n&U_{\\rm{sym}}(\\rho,|\\v{k}|)=\\frac{1}{2}A_{\\rm{d}}\\left(\\frac{\\rho}{\\rho_0}\\right)-\\frac{2Bx}{\\sigma+1}\\left(\\frac{\\rho}{\\rho_0}\\right)^{\\sigma}\\notag\\\\\n&+\nC_{\\rm{d}}\\left(\\frac{\\rho}{\\rho_0}\\right)\\left[\\gamma_0+\\gamma_1a\\left(\\frac{|\\v{k}|k_{\\rm{F}}}{\\Lambda^2}\\right)^{1\/2}\n+\\gamma_2b\\left(\\frac{|\\v{k}|k_{\\rm{F}}}{\\Lambda^2}\\right)^{1\/3}\\right].\n\\end{align}\nThe expressions for $\\gamma_0,~\\gamma_1$ and $\\gamma_2$ are given in\nEqs.\\,(\\ref{def_gamma0}), (\\ref{def_gamma1}) and (\\ref{def_gamma2}), respectively.\nNotice that in the FFG model, besides $\n\\Delta_0\\to1$, we also have\n\\begin{align}\n\\Delta_0\\Delta_1=&-{3C_0[C_1(\\phi_0-1)+\\phi_1]}\/{\\phi_0}\\to0,\\\\\n\\Delta_0\\Delta_2=&-{3C_0\\phi_1(C_1-\\phi_1)}\/{\\phi_0}\\to0,\n\\end{align}\nsee (\\ref{Def_Delta012}), thus,\n\\begin{align}\n&Y_{10}\\to2,~~Y_{11}\\to{32}\/{21},~~Y_{12}\\to{33}\/{20},\\\\\n&Z_{10}\\to-2,~~Z_{11}\\to-{8}\/{7},~~Z_{12}\\to-{27}\/{20},\n\\end{align}\ntogether with $Y_{2j}\\to0,~Z_{2j}\\to0,~Y_{3j}\\to0,~Z_{3j}\\to0$ with $j=0,1,2$, see\nthe expressions (\\ref{def_Y20}), (\\ref{def_Y21}), (\\ref{def_Y22}),\n(\\ref{def_Z20}), (\\ref{def_Z21}), (\\ref{def_Z22}), (\\ref{def_Y30}),\n(\\ref{def_Y31}), (\\ref{def_Y32}), (\\ref{def_Z30}),\n(\\ref{def_Z31}) and (\\ref{def_Z32}) and the consequent result is the\nsymmetry energy in the FFG model, see Eq.\\,(\\ref{FFG-Esym}).\nMoreover, the expressions for the slope parameter $L(\\rho)$ and the symmetry potential\n$U_{\\rm{sym}}(\\rho,|\\v{k}|)$ in the FFG model are given in Eq.\\,(\\ref{FFG-L}) and Eq.\\,(\\ref{FFG-Usym}), respectively.\n\n\nConsidering the expression for the symmetry energy, the factor characterizing the effects of the HMT, i.e.,\n\\begin{align}\\label{def_Upsilon}\n&\\Upsilon_{\\rm{sym}}=1+C_0(1+3C_1)\\left(5\\phi_0+\\frac{3}{\\phi_0}-8\\right)\\notag\\\\\n&+3C_0\\phi_1\\left(1+\\frac{5}{3}C_1\\right)\\left(5\\phi_0-\\frac{3}{\\phi_0}\\right)\n+\\frac{27C_0\\phi_1^2}{5\\phi_0},\n\\end{align}\nis generally smaller than unity, leading to a reduction of the\nkinetic symmetry energy\\,\\cite{Cai15}. According to the single-nucleon potential decomposition of the symmetry\nenergy\\,\\cite{XuC10,XuC11,CheR12} based on the Hugenholtz-Van Hove (HVH) theorem\\,\\cite{Hug58}, a contribution of\n$2^{-1}U_{\\rm{sym}}(\\rho,|\\v{k}|=k_{\\rm{F}})$ to the\nsymmetry energy is obtained. Thus a reduction of the kinetic symmetry energy\nwill induce an enhancement of the symmetry potential, i.e., the\nsymmetry energy encapsulating the SRC effects will intrinsically\nhave impacts on the symmetry potential.\nIn addition, by using the expression of $A_{\\rm{d}}$ in Eq.\\,({\\ref{ad}) the last two terms in Eq.\\,(\\ref{HMT-Esym-mm}) can be rewritten as\n\\begin{equation}\n\\frac{1}{4}A_{\\rm{d}}^0\\left(\\frac{\\rho}{\\rho_0}\\right)+\\frac{xB}{1+\\sigma}\\left(\\frac{\\rho}{\\rho_0}\\right)\n\\left[1-\\left(\\frac{\\rho}{\\rho_0}\\right)^{\\sigma-1}\\right],\n\\end{equation}\nindicating that the $x$ does not affect the symmetry energy at\n$\\rho_0$\\,\\cite{Che05,XuJ10,XuJ15}. On the other hand, the corresponding contribution to the slope parameter of symmetry energy\n$L(\\rho)$ is $\n3A_{\\rm{d}}^0\/4+3xB({1-\\sigma})\/({1+\\sigma})$ at $\\rho=\\rho_0$. Thus\na larger (positive) $x$ corresponds to a softer symmetry energy if\n$B(1-\\sigma)<0$, once the other phenomenological parameters are\nfixed\\,\\cite{Che05}, see TABLE \\ref{tab_para}.\n\n\n\n\n\n\\setcounter{equation}{0}\n\\section{Numerical Demonstrations}\\label{sec5}\n\n\n\\subsection{Schemes for Determining the Coupling Constants}\\label{sb_parameter}\n\n\nIn this subsection, we discuss the scheme to determine the model couplings, i.e., $A_{\\rm{tot}},~B,~C_{\\rm{tot}},~A_{\\rm{d}},~C_{\\rm{d}},~\\sigma,~a,~b,~x$ and $\\Lambda$.\nAs pointed out in Ref.\\,\\cite{CLC18} that the function $\\Omega(\\v{k},\\v{k}')$ is invariant under the transformations,\n\\begin{equation}\na\\to a\/\\xi^{3\/2},~~b\\to\\xi b,~~\\Lambda\\to\\Lambda\/\\xi^{3\/2},\n\\end{equation}\nwith $\\xi>0$ being any scaling factor, i.e., we have the freedom to first fix one of them without affecting the physical results. We set $b = 2$ in the following and then determine other parameters using known empirical constraints.\n\n\n\\begin{table*}[t!]\n\\caption{Coupling constants used in the two models (right side) and\nsome empirical properties of ANM used to fix\nthem (left side), $b=2$ and $\\Lambda=1.6\\,\\rm{GeV}$ are fixed,\n$K_0\\equiv K_0(\\rho_0),~M_0^{\\ast}\\equiv M_0^{\\ast}(\\rho_0),~L\\equiv\nL(\\rho_0)$.\nSee also Ref.\\,\\cite{CLC18}.\n}\\label{tab_para} {\\normalsize\\begin{tabular}{lr||cccc}\n\\hline\\hline Quantity& Value & Coupling & FFG&HMT-SCGF&HMT-exp\\\\\n\\hline $\\rho_0$ (fm$^{-3}$) & $0.16$ & $A_\\ell^0$ $(\\rm{MeV})$ &\n$-266.2934$&$520.3611$&$146.6085$\\\\\n\\hline $E_0(\\rho_0)$\n$(\\rm{MeV})$&$-16.0$&$A_{\\rm{u}}^0$ $(\\rm{MeV})$&$-86.8331$&$805.3082$&$1216.2500$\\\\\n\\hline $M_0^{\\ast}\/M$ &$0.58$&$B$ $(\\rm{MeV})$&$517.5297$&$-256.9850$&$-64.5669$\\\\\n\\hline $K_0$ ($\\rm{MeV}$) &$230.0$&$C_\\ell$ $(\\rm{MeV})$&$-155.6406$&$-154.7508$&$-37.3249$\\\\\n\\hline $U_0(\\rho_0,0)$ ($\\rm{MeV}$)\n&$-100.0$&$C_{\\rm{u}}$ $(\\rm{MeV})$&$-285.3256$&$-351.0989$&$-679.5379$\\\\\n\\hline $E_{\\rm{sym}}(\\rho_0)$ ($\\rm{MeV}$)&$31.6$&$\\sigma$&$1.0353$&$0.9273$&$0.6694$\\\\\n\\hline $L$ ($\\rm{MeV}$)&$58.9$&$a$&$-5.4511$&$-5.0144$&$-4.1835$\\\\\n\\hline $U_{\\rm{sym}}(\\rho_0,1\\,\\rm{GeV})$ ($\\rm{MeV}$)&$-20.0$&$x$&$0.6144$&$0.3774$&$-0.2355$\\\\\n\\hline\\hline\n\\end{tabular}}\n\\end{table*}\n\n\nIn our scheme, the values of $E_0(\\rho_0),~ \\rho_0, ~K_0\\equiv\nK_0(\\rho_0), ~M_0^{\\ast}(\\rho_0)$ and $U_0(\\rho_0,0\\,\\rm{MeV})$\nfor SNM, and those of $E_{\\rm{sym}}(\\rho_0), ~L\\equiv L(\\rho_0)$ and\n$U_{\\rm{sym}}(\\rho_0,1\\,\\rm{GeV})$ for the symmetry energy, its slope and the\nsymmetry potential are fixed at their currently known most probable empirical values. For example, the value of the incompressibility $K_0=230\\pm20\\,\\rm{MeV}$ was\ndetermined from analyzing nuclear giant resonances\n(GMR)\\,\\cite{You99,Shl06,Pie10,Che12,Col14,Stone2014PRC,Garg2018PPNP,LiBA2021PRC,XuJ2021PRC,ZhangZ2021CPC}, and the nucleon effective mass at $\\rho_0$ is fixed at $0.58M$\\,\\cite{CLC18,LiBA2018PPNP}.\nFor the $E_\\text{sym}(\\rho_0)$ and $L $, all existing constraints extracted\nso far from both terrestrial laboratory measurements and\nastrophysical observations are found to be essentially consistent\nwith the global averages of $E_{\\text{sym}}({\\rho _{0}}) \\approx\n31.6\\pm2.66$ MeV and $L \\approx 58.9 \\pm 16$ MeV\\,\\cite{LiBA13}, see also Ref.\\,\\cite{Chen17,LiBA21}. Consequently, we can fix for the SNM five phenomenological parameters, i.e.,\n$A_{\\rm{tot}},~B,~C_{\\rm{tot}},~\\sigma$ and $a$, while for the symmetry\nenergy and the symmetry potential, the remaining three parameters,\ni.e., $A_{\\rm{d}},~C_{\\rm{d}}$ and $x$ can be fixed. Equivalently,\nthe totally eight parameters, i.e.,\n$A_\\ell^0,~A_{\\rm{u}}^0,~B,~C_\\ell,~C_{\\rm{u}},~\\sigma,~a$ and $x$ can be\ndetermined.\n\nNext, we need to select the cutoff $\\Lambda$.\nAt first glance, the $\\Lambda$ could be determined by adopting another empirical constraint, e.g., the symmetry energy at two times the saturation density $E_{\\rm{sym}}(2\\rho_0)$\\,\\cite{LiBA21}.\nBut it does not work in this manner.\nIn particular, one can demonstrate that once we fix the above eight quantities, i.e., $E_0(\\rho_0),~\\rho_0,~K_0,~M_0^{\\ast}(\\rho_0),~U_0(\\rho_0,0\\,\\rm{MeV}),~E_{\\rm{sym}}(\\rho_0),~L$ and\n$U_{\\rm{sym}}(\\rho_0,1\\,\\rm{GeV})$, these quantities at other densities or other energy scales do not further depend on the cutoff $\\Lambda$, see APPENDIX \\ref{app2} for more discussions.\nThe role of the $\\Lambda$ is different from the other model parameters.\nHowever, the $\\Lambda$ parameter may affect quantities which are not used in the fixing scheme, e.g., $E_{\\rm{sym},4}(\\rho)$.\nIn the following we select the $\\Lambda$ parameter based on the empirical fact that the fourth-order symmetry energy at $\\rho_0$ in the FFG model is smaller than about 3\\,MeV\\,\\cite{Cai12,Sei14,Gon17,PuJ17,CWZC2022}.\n\nWe construct three models of EOS of ANM using\ndifferent HMT input parameters described in section \\ref{sec2},\ni.e., the FFG model as well as the HMT-SCGF and HMT-exp models. Again, we emphasize that the ``FFG'' or the ``HMT'' is only used to label the $n_{\\v{k}}^J(\\rho,\\delta)$ and the potential EOS is included in all of these models.\nSee TABLE \\ref{tab_para} for the values of the model parameters, here the cutoff $\\Lambda$ is set to be 1.6\\,GeV to fulfill the aforementioned constraint (specifically $1.40\\,\\rm{GeV}\\lesssim\\Lambda\\lesssim1.65\\,\\rm{GeV}$).\nIt is necessary to point out that since we fixed the parameters of\nthe HMT by using experimental data and\/or microscopic-theory calculations at the saturation density, the possible density dependence of those\nparameters, i.e., $C_0,~C_1,~\\phi_0$ and $\\phi_1$ is not\nexplored here (also see discussions given in Ref.\\,\\cite{Cai15}). The density dependence of the various terms\nin the kinetic EOS is thus only due to that of the nucleon Fermi momentum $k_{\\rm{F}}$.\nFrom the table, one can clearly find that the parameter $a$ is negative, i.e., the moving nucleons feel weaker interaction compared to the static ones,\nsee the discussions given after the formula (\\ref{MFunc}).\nThe EDF constructed is abbreviated as the abMDI\\,\\cite{CLC18}.\n\n\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[height=2.8cm]{CC-Lambda-a.eps}\\quad\n \\includegraphics[height=2.8cm]{CC-Lambda-Ctot.eps}\\quad\n \\includegraphics[height=2.8cm]{CC-Lambda-RA.eps}\n \\caption{The $\\Lambda$-dependence of the coupling constants $a$ (left) and $C_{\\rm{tot}}$ (middle) and the ratio $C_{\\rm{tot}}\/a^2$ (right) in the three models.}\n \\label{fig_CC-Lambda}\n\\end{figure}\n\nWe show in FIG.\\,\\ref{fig_CC-Lambda} the $\\Lambda$-dependence of the coupling constants $a$ (left panel) and $C_{\\rm{tot}}$ (middle panel) as well as the ratio $C_{\\rm{tot}}^2\/a$ (right panel) in the three models. The parameters $a$ and $C_{\\rm{tot}}$ evolve according to $a=a^0(\\Lambda\/\\Lambda_0)^{1\/3}$ and $C_{\\rm{tot}}=C_{\\rm{tot}}^0(\\Lambda\/\\Lambda_0)^{2\/3}$, respectively, see (\\ref{CC-aa}), and thus $C_{\\rm{tot}}\/a^2=C_{\\rm{tot}}^0\/(a^{0})^2=\\rm{const.}$, with the constant depending on the fitting scheme, i.e., it depends on the values of $\\rho_0,~E_0(\\rho_0),~K_0,~M_0^{\\ast}(\\rho_0)$ and $U_0(\\rho_0,0)$ via the specific model.\nSee the right panel of FIG.\\,\\ref{fig_CC-Lambda}, e.g., the ratio $C_{\\rm{tot}}\/a^2$ in the HMT-exp model is about \n$-40.95\\,\\rm{MeV}$ (see TABLE \\ref{tab_para}).\nOne can also show numerically that as $\\Lambda\\to0$, both $a$ and $C_{\\rm{tot}}$ approach zero, but their ratio keeps a constant.\nThe $\\Lambda$-dependence of other coupling constants ($A_{\\ell}^0,~A_{\\rm{u}}^0,~C_{\\ell},~C_{\\rm{u}}$) could be obtained similarly, and would not be analyzed here further (see APPENDIX B for more discussions).\n\n \n \n\\subsection{SNM EOS $E_0(\\rho)$ and Single-nucleon Potential $U_0(\\rho,|\\v{k}|)$}\n\nIn FIG.\\,\\ref{fig_ab_E0}, we show the SNM EOSs in the three models.\nIt is first interesting to see that the HMT models predict a\nslightly harder SNM EOS at supra-saturation densities than the\nFFG model, while by design they all have the same values of\n$M_0^{\\ast}$, $\\rho_0$, $E_0(\\rho_0)$ and $K_0$. Physically, this is\neasy to understand because of the large contribution to the kinetic EOS by the\nhigh momentum nucleons in the HMT models\\,\\cite{Cai16b}. For\nexample, the skewness of the SNM characterizing the high density\nbehavior of the EOS in the FFG model is found to be\n$J_0(\\rm{FFG})\\approx-381\\,\\rm{MeV}$, while that in the two HMT\nmodels is about $J_0(\\rm{HMT-SCGF})\\approx-376\\,\\rm{MeV}$ and\n$J_0(\\rm{HMT-exp})\\approx-329\\,\\rm{MeV}$, respectively. The small\nchange in the skewness from the FFG model to the HMT-SCGF model can\nbe traced back to the enhancement factor\n$\\Upsilon_0=1+C_0(5\\phi_0+3\/\\phi_0-8)$, which is unity in the FFG\nmodel and becomes about 1.18\/1.83 in the HMT-SCGF\/HMT-exp model. It\nis thus not surprising that the predictions on the pressure in SNM\nis also very similar, and can all pass through the constraints from analyzing \nthe collective flow date from relativistic heavy-ion collisions\\,\\cite{Dan02},\nsee FIG.\\,\\ref{fig_ab_Flow}.\nIt is necessary to point out that since the high momentum nucleon fraction in the HMT-SCGF is relatively lower than that in the HMT-exp model, the former is much closer to the FFG model (see the black dash line and the blue dash-dotted line shown in the inset of FIG.\\,\\ref{fig_ab_E0}).\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[width=6.5cm]{E0-EOS.eps}\n \\caption{EOS of SNM in the FFG model and the HMT-SCGF as well as the HMT-exp models.}\n \\label{fig_ab_E0}\n\\end{figure}\n\n\n\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[width=6.5cm]{P0Flow.eps}\n \\caption{Comparison between the pressure $P_0$ of SNM with the experimental constraints from analyzing nuclear collective flows in heavy ion collisions\\,\\cite{Dan02}.}\n \\label{fig_ab_Flow}\n\\end{figure}\n\n\nIn FIG.\\,\\ref{fig_ab_U0}, we show the single-nucleon potential in SNM\n$U_0$ as a function of kinetic energy $E_{\\rm{kin}}=[|\\v{k}|^2+M^2]^{1\/2}+U_0(\\rho_0,|\\v{k}|)-M$\nin the three models. As discussed earlier, the SRC-induced HMT (slightly) enhances\nthe kinetic EOS of SNM, and in order to maintain the total EOS of SNM\nat the saturation density to be consistent with empirical constraints the\npotential part should be correspondingly reduced, which is reflected\nin FIG.\\,\\ref{fig_ab_U0}. As the difference of the fraction of high\nmomentum nucleons in SNM and in PNM becomes larger, the kinetic EOS\nof SNM is much more enhanced, leading to a much softer potential\npart, i.e., the $U_0$ in the HMT-exp model\nis smaller than that in the HMT-SCGF model. Also shown in FIG.\\,\\ref{fig_ab_U0} are the predictions on the $U_0(\\rho_0,|\\v{k}|)$\nfrom other approaches, including the global relativistic fitting of\nelectron-scattering data up to about 1\\,GeV\\,\\cite{Ham90} (blue ``+''); the\npredictions by the neutron optical model up to about\n200\\,MeV\\,\\cite{LiX15} (dashed black band), and those from the chiral effective field\ntheories\\,\\cite{Hol16} (cyan band). It is obvious that the predictions on $U_0$\nboth in the HMT-SCGF and the HMT-exp models are consistent with these approaches, at kinetic energies $\\lesssim600\\,\\rm{MeV}$.\nIn this sense, the momentum-dependence in the function (\\ref{MFunc})\ngives a reasonable description of the kinetic-energy-dependence of the single-nucleon isoscalar\npotential. As the kinetic energy $E_{\\rm{kin}}$ becomes larger, the\ndeviation between the predictions on the $U_0$ in the three models and\nthe global relativistic fitting also becomes larger, indicating the breakdown of the perturbative construction for $\\Omega(\\v{k},\\v{k}')$.\n\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[width=6.5cm]{U0Hama.eps}\n \\caption{single-nucleon potential in SNM as a function of kinetic energy in the FFG model and\n HMT-SCGF\/HMT-exp model. Constraints from other approaches are shown for comparison.}\n \\label{fig_ab_U0}\n\\end{figure}\n\n\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[width=6.5cm]{Mstar.eps}\n \\caption{Nucleon effective k-mass as a function of density in the FFG and HMT-SCGF\/HMT-exp models.}\n \\label{fig_ab_M0}\n\\end{figure}\n\n\n\nIn FIG.\\,\\ref{fig_ab_M0}, the density dependence of the effective k-mass is\nshown. It is straightforward to understand that the three models give very\nsimilar results. On one hand, three points of the effective mass are\nfixed, i.e., $M_0^{\\ast}(0)\/M=1$, $M_0^{\\ast}(\\rho_0)\/M=0.58$ and\n$M_0^{\\ast}(\\infty)\/M=0$ (see Eq.\\,(\\ref{def_11}) since both $a$ and $C_{\\rm{tot}}$ are negative, leading to $aC_{\\rm{tot}}>0$). Moreover, the\n$M_0^{\\ast}\/M$ monotonically decreases in the whole density range\nand is concave, e.g., for large density $\\rho$, one has from (\\ref{def_11}) that $\nM_0^{\\ast}\/M\\sim\\rho^{-2\/3}$. Meeting all of these common constraints the\n$M_0^{\\ast}(\\rho)\/M$ in the three models behaves thus very similarly (see similar discussions given in Ref.\\,\\cite{Cai16b} for the nonlinear RMF model counterpart).\n\n\n\n\n\\subsection{Reduction of the Kinetic Symmetry Energy $E_{\\textmd{sym}}^{\\textmd{kin}}(\\rho)$ and Enhancement of the Symmetry Potential $U_{\\textmd{sym}}(\\rho,|\\v{k}|)$}\n\n\nWe now discuss effects of the SRC\/HMT on the symmetry energy and isovector (symmetry) potential.\nIn FIG.\\,\\ref{fig_ab_Esym}, we show the density dependence of the nuclear\nsymmetry energy in the three models\\,\\cite{CLC18}.\nAlso shown include the constraints on the $E_{\\rm{sym}}(\\rho)$\naround the saturation density from analyses of heavy-ion collisions\\,\\cite{Tsa12} and isobaric analog state studies\\,\\cite{Dan14}. Although the symmetry energy from the three models can pass through these constraints, they have very different behaviors at super-saturation\ndensities\\,\\cite{CLC18}. Specifically, as discussed in the above sections, the\nreduction of the kinetic symmetry energy should be compensated by\nthe potential part to keep the total symmetry energy around $\\rho_0$ consistent with certain empirical constraints\\,\\cite{LiBA13,Chen17,LiBA21}.\nMore quantitatively, the factor $\\Upsilon_{\\rm{sym}}$ (defined in Eq.\\,(\\ref{def_Upsilon})) in the HMT-SCGF model is found to be about\n$\\Upsilon_{\\rm{sym}}\\approx0.71$, which is not far different from\nthat in the FFG model (which is unity), while the\n$\\Upsilon_{\\rm{sym}}$ factor in the HMT-exp model is about\n$\\Upsilon_{\\rm{sym}}\\approx-1.12$\\,\\cite{Cai15}, totally different from the FFG\nprediction. Thus it is not surprising that the effects of HMT in the\nHMT-exp model is much more apparent. \n\nIn addition, there are two critical densities $\\rho_1$ and $\\rho_2$ corresponding to the point above which the symmetry energy starts decreasing and its zero-point when the symmetry energy vanishes. They are defined as $L(\\rho_1)=0$ and $E_{\\rm{sym}}(\\rho_2)=0$, respectively. The latter case indicates the onset of the so-called isospin separation instability, namely it is energetically more favorable to split SNM into pure neutron matter and proton matter when the symmetry energy becomes negative\\,\\cite{Kut93,Kut94,LiBA02}. Possible appearances and ramifications of such instability in heavy-ion reactions and neutron stars have been studied in several works, see., e.g., Refs.\\,\\cite{Kut93,Kut94,LiBA02,Szm06,wen,Kho96,Bas07,Ban00,Kubis1}.\nWithout considering the SRC-induced HMT, the FFG model predicts $\\rho_1^{\\rm{FFG}}\\approx4.4\\rho_0$ and $\\rho_2^{\\rm{FFG}}\\approx11.2\\rho_0$, while the HMT reduces the $\\rho_1$ to $3.5\\rho_0$ ($1.9\\rho_0$) and the $\\rho_2$ to $8.5\\rho_0$ ($4.1\\rho_0$) in the HMT-SCGF (HMT-exp) model.\nWe see that the experimental HMT constraints largely reduce the zero-point density $\\rho_2$, with an effect of about 63\\%.\nSince the $\\rho_2^{\\rm{HMT}\\mbox{-}\\rm{exp}}\\approx4.1\\rho_0$ is quite low considering neutron star issues, it is expected that the reduction of the symmetry energy at supra-saturation densities in the HMT-exp model may have sizable effects on properties of neutron stars, e.g., the mass-radius relation.\nOn the other hand, the reduction of the $E_{\\rm{sym}}(\\rho)$ at sub-saturation densities (see the inset of FIG.\\,\\ref{fig_ab_Esym}) may also induce changes on, e.g., the core-crust transition densities in neutron stars.\nThese issues are investigated in some details in subsection \\ref{sb_NS}.\n\nFurthermore, we also show in FIG.\\,\\ref{fig_ab_Esym} two constraints on the $E_{\\rm{sym}}(2\\rho_0)$. The pink solid circle at $E_{\\rm{sym}}(2\\rho_0)\\approx45\\pm3\\,\\rm{MeV}$ is from chiral perturbation theories with consistent nucleon-nucleon and three-nucleon interactions up to the fourth-order\\,\\cite{Dri20}. The grey diamond at $E_{\\rm{sym}}(2\\rho_0)\\approx51\\pm13\\,\\rm{MeV}$ is the fiducial value from surveying analyses of both relativistic heavy-ion reactions and neutron star properties since GW170817\\,\\cite{LiBA21}. It is clearly seen that the calculations with the SRC\/HMT-induced reduction of $E_{\\rm{sym}}(2\\rho_0)$ are consistent with these constraints\\,\\cite{LiBA21,Dri20,CL2021}.\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[width=6.5cm]{Esym.eps}\n \\caption{Density dependence of the symmetry energy in the FFG model and HMT-SCGF\/HMT-exp model. Constraints from heavy-ion collisions\\,\\cite{Tsa12} and the isobaric analog state studies\\,\\cite{Dan14} are shown for comparison.\nSee also Ref.\\,\\cite{CLC18} and the text for details on the pink solid circle and the grey diamond.}\n \\label{fig_ab_Esym}\n\\end{figure}\n\n\n\nThe reduction of the\nsymmetry energy found here is qualitatively consistent with the one from the nonlinear RMF\nmodels\\,\\cite{Cai16b}. \nHowever, they give quantitatively different results. \nIn particular, due to the specific structure of the nonlinear RMF model, the symmetry energy should never becomes negative.\nOn the other hand, the EOS in a non-relativistic EDF could be treated as an effective expansion in density. \nIn order to investigate the probable origin of this quantitative difference, it is enough for our purpose to assume that the symmetry energy takes the form $E_{\\rm{sym}}(\\rho)=E_{\\rm{sym}}^{\\rm{kin}}(\\rho_0)(\\rho\/\\rho_0)^{2\/3}+s_1(\\rho\/\\rho_0)^{\\alpha}+s_2(\\rho\/\\rho_0)^{\\beta}$, where $\\beta>\\alpha>2\/3$, and $E_{\\rm{sym}}^{\\rm{kin}}(\\rm{FFG})\\approx12.3\\,\\rm{MeV}$.\nThe coefficients $s_1$ and $s_2$ can be determined via the values of $E_{\\rm{sym}}(\\rho_0)$ and $L\\equiv L(\\rho_0)$. Specifically, the expression for $s_2$ is given by,\n\\begin{equation}\ns_2=\\frac{1}{\\beta-\\alpha}\\left(\\alpha-\\frac{2}{3}\\right)E_{\\rm{sym}}^{\\rm{kin}}(\\rho_0)+\\frac{L-3\\alpha E_{\\rm{sym}}(\\rho_0)}{\\beta-\\alpha}.\n\\end{equation}\nThe second term here is negative since $\\alpha>2\/3$ and thus $L-3\\alpha E_{\\rm{sym}}(\\rho_0)<0$ while $\\beta-\\alpha>0$. \nSimilarly, the coefficient of the first term of $E_{\\rm{sym}}^{\\rm{kin}}(\\rho_0)$ is positive (due to the same consideration), thus:\n\\begin{enumerate}\n\\item[(a)]One has $s_2^{\\rm{FFG}}>s_2^{\\rm{HMT}}$ as $E_{\\rm{sym}}^{\\rm{kin}}(\\rm{FFG})>E_{\\rm{sym}}^{\\rm{kin}}(\\rm{HMT})$. \n\\item[(b)]In addition, $s_2<0$ for the HMT-exp model, as the $E_{\\rm{sym}}^{\\rm{kin}}(\\rho_0)\\approx-13.8\\,\\rm{MeV}$ is negative. \n\\end{enumerate}\nThe point (a) explains why the $E_{\\rm{sym}}(\\rho)$ considering the HMT should be reduced while the point (b) indicates that in the HMT-exp model (with a negative $E_{\\rm{sym}}^{\\rm{kin}}(\\rho_0)$) the $E_{\\rm{sym}}(\\rho)$ must start decreasing above some critical density, and become negative at even higher densities. In our EDF, $\\beta=4\/3$ and one can find actually that the $s_2$'s in all three models are negative, see TABLE \\ref{tab_para1}.\nThis explains why the reduction of symmetry energy using the non-relativistic EDF should be much stronger than the one in the nonlinear RMF models.\n\n\nThe reduction of the symmetry energy at supra-saturation densities also generates corresponding reduction of its curvature coefficient $K_{\\rm{sym}}$. Quantitatively, its value changes from $-109$\\,MeV in the FFG model to about\n$-121\\,\\rm{MeV}$ ($-223$\\,MeV) in the HMT-SCGF (HMT-exp) model.\nWhile the $K_{\\rm{sym}}\\approx-223\\,\\rm{MeV}$ in the HMT-exp model is somewhat smaller than its current fiducial value of about $-107\\pm 88$ MeV (see FIG.\\,2 of Ref.\\,\\cite{LiBA21}), those from the FFG and the HMT-SCGF model are very consistent with the fiducial value. Thus, the general tendency is that the SRC\/HMT reduces the symmetry energy curvature coefficient $K_{\\rm{sym}}$. This may have some other consequences. For example, the pressure of neutron-rich nucleonic matter is given by\n\\begin{align}\nP\/3\\rho\\approx&L(\\rho)\\delta^2+3\\rho\\frac{\\d E_0(\\rho)}{\\d\\rho}\\notag\\\\\n\\approx&\nL\\delta^2\\left(1+3\\chi+\\frac{9}{2}\\chi^2\\right)\n+K_0\\chi(1+3\\chi)+\\frac{1}{2}\\chi^2J_0\\notag\\\\\n&+K_{\\rm{sym}}\\delta^2\\chi(1+3\\chi)\\left[1+\\frac{J_{\\rm{sym}}}{K_{\\rm{sym}}}\n\\frac{\\chi}{2(1+3\\chi)}\\right]\\notag\\\\\n\\approx&L\\delta^2\\left(1+3\\chi+\\frac{9}{2}\\chi^2\\right)\n+K_0\\chi(1+3\\chi)\\notag\\\\\n&+\\frac{1}{2}\\chi^2J_0+K_{\\rm{sym}}\\delta^2\\chi(1+3\\chi),\n\\end{align}\nwhere the skewness coefficient $J_{\\rm{sym}}$ is neglected in the last step.\nSince the $K_0$ and $L$ are fixed in our study, and as discussed earlier the $J_0$ is slightly increased by the SRC\/HMT, the reduction of $K_{\\rm{sym}}$ may reduce the pressure $P$ at densities below about $3\\rho_0$ above which the $J_{\\rm{sym}}$ becomes important (the $\\chi$ is 2\/3 for $\\rho=3\\rho_0$ and thus $\\chi\/[2(1+3\\chi)]=1\/9$). Moreover, the reduction of $K_{\\rm{sym}}$ may also modify the saturation line (on which the pressure is zero) of ANM. More quantitatively, the incompressibility coefficient of ANM along its saturation line is $K(\\delta)=K_0+K_{\\rm{sat},2}\\delta^2$\\,\\cite{LWChen2009} where the $K_{\\rm{sat},2}$ is\n \\begin{equation}\nK_{\\rm{sat,2}}=K_{\\rm{sym}}-6L-\\frac{J_0L}{K_0}.\n\\end{equation}\nIt changes from $-365\\,\\rm{MeV}$ in the FFG model to $-378\\,\\rm{MeV}$ ($-492\\,\\rm{MeV}$) in the HMT-SCGF (HMT-exp) model. \nIt is seen that the maximum reduction is about 127\\,\\rm{MeV}.\n\n\\begin{table*}[t!]\n\\caption{Parameterizations of the EOS of SNM and the symmetry energy as\nwell as the nucleon potential. Momentum $|\\v{k}|$ is measured in MeV and\nthe density $\\rho$ in $\\rm{fm}^{-3}$,\n$E_0(\\rho),E_{\\rm{sym}}(\\rho),U_0(\\rho,|\\v{k}|)$ and\n$U_{\\rm{sym}}(\\rho,|\\v{k}|)$ are in MeV.}\\label{tab_para1}\n{\\normalsize\n\\begin{tabular}{c||c}\n\\hline\\hline Model for $n_{\\v{k}}^J$& Parametrization of the Physical Quantity\\\\\\hline \n\\hline FFG &$x_{\\rm{SNM}}^{\\rm{HMT}}=0\\%,x_{\\rm{PNM}}^{\\rm{HMT}}=0\\%$\n\\\\\n\\hline $E_0(\\rho)$ &\n$E_0(\\rho)\\approx75.00\\rho^{2\/3}+1695.42\\rho^{1.04}-551.76\\rho-1378.02\\rho[1-0.81\\rho^{1\/3}+0.55\\rho^{2\/9}]$\\\\\n\\hline $U_0(\\rho,|\\v{k}|)$ &$U_0(\\rho,|\\v{k}|)\\approx3450.68\\rho^{1.04}-1103.52\\rho-2756.04\\rho[1-0.04|\\v{k}|^{1\/2}\\rho^{1\/6}+0.08|\\v{k}|^{1\/3}\\rho^{1\/9}]$\\\\\n\\hline $E_{\\rm{sym}}(\\rho)$ & $E_{\\rm{sym}}(\\rho)\\approx41.67\\rho^{2\/3}+1101.25\\rho-1041.63\\rho^{1.04}-229.34\\rho^{4\/3}+180.92\\rho^{11\/9}$\\\\\n\\hline $U_{\\rm{sym}}(\\rho,|\\v{k}|)$ &\n$U_{\\rm{sym}}(\\rho,|\\v{k}|)\\approx-2083.27\\rho^{1.04}+\\rho[2202.51-40.52|\\v{k}|^{1\/2}\\rho^{1\/6}+69.80|\\v{k}|^{1\/3}\\rho^{1\/9}]$\\\\\\hline\n\\hline HMT-SCGF &$x_{\\rm{SNM}}^{\\rm{HMT}}=12\\%,x_{\\rm{PNM}}^{\\rm{HMT}}=4\\%$\n\\\\\n\\hline $E_0(\\rho)$ & $E_0(\\rho)\\approx88.28\\rho^{2\/3}-729.40\\rho^{0.93}+2071.36\\rho-1580.78\\rho[1-0.79\\rho^{1\/3}+0.57\\rho^{2\/9}]$\\\\\n\\hline $U_0(\\rho,|\\v{k}|)$ &$U_0(\\rho,|\\v{k}|)\\approx-1405.76\\rho^{0.93}+4142.72\\rho-3161.56\\rho[1-0.04|\\v{k}|^{1\/2}\\rho^{1\/6}+0.08|\\v{k}|^{1\/3}\\rho^{1\/9}]$\\\\\n\\hline $E_{\\rm{sym}}(\\rho)$ & $E_{\\rm{sym}}(\\rho)\\approx29.53\\rho^{2\/3}-146.18\\rho+275.30\\rho^{0.93}-467.07\\rho^{4\/3}+343.07\\rho^{11\/9}$\\\\\n\\hline $U_{\\rm{sym}}(\\rho,|\\v{k}|)$ &\n$U_{\\rm{sym}}(\\rho,|\\v{k}|)\\approx550.60\\rho^{0.93}+\\rho[-292.37-57.07|\\v{k}|^{1\/2}\\rho^{1\/6}+106.42|\\v{k}|^{1\/3}\\rho^{1\/9}]$\\\\\\hline\n\\hline HMT-exp &$x_{\\rm{SNM}}^{\\rm{HMT}}=28\\%,x_{\\rm{PNM}}^{\\rm{HMT}}=1.5\\%$\n\\\\\n\\hline $E_0(\\rho)$ & $E_0(\\rho)\\approx137.31\\rho^{2\/3}-131.89\\rho^{0.67}+2129.47\\rho-2240.19\\rho[1-0.77\\rho^{1\/3}+0.63\\rho^{2\/9}]$\\\\\n\\hline $U_0(\\rho,|\\v{k}|)$ &$U_0(\\rho,|\\v{k}|)\\approx-220.18\\rho^{0.67}+4258.93\\rho-4480.39\\rho[1-0.04|\\v{k}|^{1\/2}\\rho^{1\/6}+0.08|\\v{k}|^{1\/3}\\rho^{1\/9}]$\\\\\n\\hline $E_{\\rm{sym}}(\\rho)$ & $E_{\\rm{sym}}(\\rho)\\approx-46.83\\rho^{2\/3}+392.52\\rho-31.06\\rho^{0.67}-1837.48\\rho^{4\/3}+1421.05\\rho^{11\/9}$\\\\\n\\hline $U_{\\rm{sym}}(\\rho,|\\v{k}|)$ &\n$U_{\\rm{sym}}(\\rho,|\\v{k}|)\\approx-62.11\\rho^{0.67}+\\rho[785.04-156.10|\\v{k}|^{1\/2}\\rho^{1\/6}+348.23|\\v{k}|^{1\/3}\\rho^{1\/9}]$\\\\\n\\hline\\hline\n\\end{tabular}}\n\\end{table*}\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[height=3.8cm]{Usym-1.eps}\n \\includegraphics[height=3.8cm]{Usym-2.eps}\n \\caption{Symmetry potential in the FFG model and the HMT-SCGF\/HMT-exp model at $\\rho=\\rho_0$ (left) and\n at $\\rho=3\\rho_0$ (right).\nConstraints on the symmetry potential at $\\rho=\\rho_0$ from other approaches are shown for\ncomparisons.}\n \\label{fig_ab_Usym}\n\\end{figure}\n\nIn FIG.\\,\\ref{fig_ab_Usym}, we show the momentum dependence of the\nsymmetry potential $U_{\\rm{sym}}(\\rho,|\\v{k}|)$ at $\\rho_0$ (left panel) and $3\\rho_0$ (right panel), respectively. The predictions on\n$U_{\\rm{sym}}(\\rho_0,|\\v{k}|)$ from other microscopic\napproaches\\,\\cite{LiX15,Hol16,Fri05} are also shown for comparisons. The empirical band is parametrized as $U_{\\rm{sym}}(E)\\approx\n(28\\pm6)\\,\\rm{MeV}-(0.15\\pm0.05)E$\\,\\cite{Hol16}. It can be found \nthat the uncertainties on the symmetry potential are larger than\nthose of the isoscalar potential $U_0$. This is mainly due to the poorly-known isospin dependence of nuclear interactions\\,\\cite{LCK08,LiBA2018PPNP}.\nWhile the prediction by the HMT-exp on the\n$U_{\\rm{sym}}(\\rho_0,|\\v{k}|)$ has certain deviation from the microscopic model predictions, the FFG and HMT-SCGF model\npredictions are quite consistent with the optical model fitting\\,\\cite{LiX15} as well as the chiral effective theory prediction\\,\\cite{Hol16}.\nSince there exists no direct constraint on the symmetry potential at, e.g., a large density $\\rho=3\\rho_0$, and it is beyond the current application domain of microscopic nuclear many-body theories, it will be interesting to test the predictions shown in the right window of FIG.\\,\\ref{fig_ab_Usym}, thus the SRC\/HMT effects in dense neutron-rich matter, using nuclear reactions with high-energy radioactive beams. Of course, for this purpose it is first necessary to investigate experimental observables sensitive to the high-density symmetry potential. The results presented here especially the various analytical expressions and parameterizations will facilitate the explorations of this topic using transport models for nuclear reactions with radioactive beams. \nListed in TABLE \\ref{tab_para1} are parametrizations of the SNM EOS $E_0(\\rho)$, nucleon isoscalar potential $U_0(\\rho,|\\v{k}|)$, symmetry energy $E_{\\rm{sym}}(\\rho)$ and isovector potential $U_{\\rm{sym}}(\\rho,|\\v{k}|)$ corresponding to the three models (with their parameters given in TABLE \\ref{tab_para}), for future applications.\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[height=3.3cm]{Esym-xpara-1.eps}\n \\includegraphics[height=3.3cm]{Esym-xpara-2.eps}\n \\includegraphics[height=3.3cm]{Esym-xpara-3.eps}\n \\caption{Symmetry energy with different effective three-nucleon force parameter $x$ in the three models indicated. }\n \\label{fig_ab_Esym_x}\n\\end{figure}\n\nIn our calculations up to now, the effective three-nucleon force parameter $x$ is self-consistently determined by fixing the $L(\\rho_0)$, see TABLE \\ref{tab_para}.\nSince the $x$ parameter only affects the slope parameter of the symmetry energy instead of its magnitude at $\\rho_0$ itself, we can adjust $x$ to investigate how the density dependence of $E_{\\rm{sym}}(\\rho)$ changes when the SRC-induced HMT is taken into account. In FIG.\\,\\ref{fig_ab_Esym_x}, the symmetry energy obtained within the three\nmodels indicated by only changing the $x$ parameter (corresponding to changing only the $L$ coefficient) while fixing at the same time all the other seven parameters listed\nin TABLE \\ref{tab_para}, is shown. For the FFG model, the $x$-dependence of $E_{\\rm{sym}}(\\rho)$ is the same as in the conventional MDI\nmodel\\,\\cite{Das2003,Che05,LCK08,XuJ10,XuJ15}. In the presence of the SRC\/HMT, the large\nuncertainties of $E_{\\rm{sym}}(\\rho)$ at high densities can come from the poorly known physics of either the three-nucleon forces and\/or the SRC\/HMT.\n\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[height=3.7cm]{Esym-frac-1.eps}\\quad\n \\includegraphics[height=3.7cm]{Esym-frac-2.eps}\n \\caption{The symmetry energy at $2\\rho_0$ and $3\\rho_0$ with different fractions of high momentum nucleons in SNM and\/or in PNM, respectively. The black circles are for the default set on $x_{\\rm{SNM}}^{\\rm{HMT}}$ and $x_{\\rm{PNM}}^{\\rm{HMT}}$, i.e., $x_{\\rm{SNM}}^{\\rm{HMT}}=28\\%$ and $x_{\\rm{PNM}}^{\\rm{HMT}}=1.5\\%$.}\n \\label{fig_ab_EsymFrac}\n\\end{figure}\n\nAs discussed in the INTRODUCTION, there are still uncertainties about the fraction of high momentum nucleons in either SNM and\/or PNM and thus its isospin dependence\\,\\cite{Hen14,Rio14}. Thus, it is interesting to study how the relevant quantities change when the HMT fractions in SNM and\/or PNM are modified.\nIn FIG.\\,\\ref{fig_ab_EsymFrac}, we show the symmetry energy at two\nreference densities, i.e., $\\rho_{\\rm{f}}=2\\rho_0$ and\n$\\rho_{\\rm{f}}=3\\rho_0$, by continuously changing the high\nmomentum nucleon fraction $x_{\\rm{SNM}}^{\\rm{HMT}}$ in SNM when fixing the $x_{\\rm{PNM}}^{\\rm{HMT}}$ in PNM at the\ndefault value of the HMT-exp model (left panel) (in PNM when fixing\nthat in SNM at the default value of the HMT-exp model (right panel)), respectively. \nThe values of $C_0$ and $C_{\\rm{n}}^{\\rm{PNM}}=C_0(1+C_1)$ are fixed at 0.161 and 0.12, respectively, while the values of $\\phi_0$ and $\\phi_1$ are readjusted correspondingly according to the given $x_{\\rm{SNM}}^{\\rm{HMT}}$ and $x_{\\rm{PNM}}^{\\rm{HMT}}$ values. The black circles are for the default parameter set of $x_{\\rm{SNM}}^{\\rm{HMT}}$ and $x_{\\rm{PNM}}^{\\rm{HMT}}$, i.e., $x_{\\rm{SNM}}^{\\rm{HMT}}=28\\%$ and $x_{\\rm{PNM}}^{\\rm{HMT}}=1.5\\%$.\nAs the difference between the $x_{\\rm{SNM}}^{\\rm{HMT}}$ and\n$x_{\\rm{PNM}}^{\\rm{HMT}}$, i.e., $\\delta x=x_{\\rm{SNM}}^{\\rm{HMT}}-x_{\\rm{PNM}}^{\\rm{HMT}}$, increases, the reduction of the\n$E_{\\rm{sym}}(\\rho_{\\rm{f}})$ becomes much more apparent. While the\nsymmetry energy at the saturation density is fixed (at $31.6\\,\\rm{MeV}$),\nthese effects on $E_{\\rm{sym}}(2\\rho_0)$ are relatively minor compared with those at $\\rho=3\\rho_0$.\n\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[height=3.6cm]{Ksat2-frac-1.eps}\\quad\n \\includegraphics[height=3.6cm]{Ksat2-frac-2.eps}\n \\caption{The same as FIG.\\,\\ref{fig_ab_EsymFrac} but for the isospin incompressibility coefficient $K_{\\rm{sat},2}=K_{\\rm{sym}}-6L-J_0L\/K_0$. }\n \\label{fig_Ksat2_frac}\n\\end{figure}\n\nSimilarly, we show in FIG.\\,\\ref{fig_Ksat2_frac} the dependence of the isospin incompressibility coefficient $K_{\\rm{sat},2}$ on the high momentum nucleon fraction.\nAs the $L$ and $K_0$ are fixed in our calculations, the dependence of the coefficient $K_{\\rm{sat},2}$ on $x_{\\rm{SNM}}^{\\rm{HMT}}$ and\/or\n$x_{\\rm{PNM}}^{\\rm{HMT}}$ mainly reflects the dependence of the curvature coefficient $K_{\\rm{sym}}$ and the skewness parameter $J_0$ on them.\nThe $K_{\\rm{sat},2}$ shown in FIG.\\,\\ref{fig_Ksat2_frac} is in good agreement with the estimate of $K_{\\rm{sat},2}\\approx-550\\pm100\\,\\rm{MeV}$ from analyzing different types\nof experimental data currently available\\,\\cite{Col14}, as shown using the cyan bands.\nFor example, if $x_{\\rm{PNM}}^{\\rm{HMT}}\\approx1.5\\%$ is fixed, then the constraint on $K_{\\rm{sat},2}$ leads to about $x_{\\rm{SNM}}^{\\rm{HMT}}\\gtrsim24\\%$. Similarly, if $x_{\\rm{SNM}}^{\\rm{HMT}}\\approx28\\%$ is fixed, then we can roughly find the constraint that $x_{\\rm{PNM}}^{\\rm{HMT}}\\lesssim12\\%$.\n\n\n\\subsection{SRC\/HMT Effects on the Proton Fraction $x_{\\textmd{p}}$, Core-crust Transition Density $\\rho_{\\textmd{t}}$ and pressure $P_{\\textmd{t}}$ as well as the Mass-radius Relation of Cold Neutron Stars at $\\beta$-equilibrium}\\label{sb_NS}\n\nAs an example of applying the Gogny-like EDFs encapsulating the SRC\/HMT in astrophysical studies, we investigate here the density profile of proton fraction $x_{\\textmd{p}}(\\rho)=[1-\\delta(\\rho)]\/2$, the core-crust transition density $\\rho_{\\rm{t}}$ and transition pressure $P_{\\rm{t}}$ as well as the mass-radius correlation in cold neutron stars at $\\beta$-equilibrium.\nThe transition density $\\rho_{\\rm{t}}$ is the nucleon number\ndensity that separates the liquid core from the inner crust in\nneutron stars. It plays an important role in determining many\nproperties of neutron\nstars\\,\\cite{Hor01,Pro06,Duc08a,Duc08b,XuJ09}.\nOne simple and widely used approach to determine the core-crust transition density $\\rho\n_{\\rm{t}}$ is the thermodynamical method. Normally, neutron star matter has to obey the intrinsic stability\ncondition\\,\\cite{Lat04,XuJ09,NBZ19,Kub07}\n\\begin{align}\n\\mathcal{U}_{\\rm{ther}}(\\rho)=&2\\rho \\frac{\\partial E(\\rho\n,x_{\\rm{p}})}{\\partial \\rho }+\\rho ^{2}\\frac{\\partial\n^{2}E(\\rho ,x_{\\rm{p}})}{\\partial \\rho ^{2}}\\notag\\\\\n& -\\left. \\left( \\frac{\\partial ^{2}E(\\rho ,x_{\\rm{p}})}{\\partial\n\\rho\n\\partial x_{\\rm{p}}}\\rho \\right) ^{2}\\right\/ \\frac{\\partial ^{2}E(\\rho ,x_{\\rm{p}})%\n}{\\partial x_{\\rm{p}}^{2}}\\notag\\\\\n=&\\rho^2\\left(\\frac{\\partial^2E(\\rho,x_{\\rm{p}})}{\\partial x_{\\rm{p}}^2}\\right)^{-1} \n\\left[\\frac{\\partial\\mu_{\\rm{n}}}{\\partial\\rho_{\\rm{n}}}\\frac{\\partial \\mu_{\\rm{p}}}{\\partial\\rho_{\\rm{p}}}-\\left(\\frac{\\partial\\mu_{\\rm{n}}}{\\partial\\rho_{\\rm{p}}}\\right)^2\\right]>0. \\label{Vther}\n\\end{align}\nIn the above, the density profile $x_{\\rm{p}}=x_{\\rm{p}}(\\rho)$ is determined by the $\\beta$-equilibrium and charge neutrality condition of neutron star matter consisting of neutrons, protons and electrons (the core-crust transition occurs around $\\rho_0\/3\\mbox{$\\sim$}\\rho_0\/2$ where muons generally do not emerge). It is uniquely determined by the density dependence of nuclear symmetry energy\\,\\cite{Lat04}.\nOnce the above stability condition is broken, the speed of sound in the uniform outer core becomes imaginary and small density oscillations will then grow exponentially, indicating the onset of a transition from the uniform liquid core to the crust. \nIn (\\ref{Vther}), $E(\\rho ,x_{\\rm{p}})$ is the energy per nucleon in\nthe $\\beta $-equilibrated neutron star matter. The corresponding pressure is obtained as $P=P_{\\rm{N}}+P_{\\rm{e}}$ with\n$P_{\\rm{N}}$ and $P_{\\rm{e}}$ denoting contributions from nucleons and electrons, respectively.\nThe electron pressure $P_{\\rm{e}}$ could be obtained through the FFG model\\,\\cite{XuJ09}, i.e., \n\\begin{equation}\\label{dd_e}\n\\varepsilon_{\\rm{e}}(\\rho,\\delta)=\\eta_{\\rm{e}}\\Phi_{\\rm{e}}(t_{\\rm{e}}),\n\\end{equation}\nhere \\begin{equation}\n\\eta_{\\rm{e}}={m_{\\rm{e}}}\/{8\\pi^2\\lambda_{\\rm{e}}^3},~~\n\\lambda_{\\rm{e}}={1}\/{m_{\\rm{e}}},~~t_{\\rm{e}}=\\lambda_{\\rm{e}}\\left(3\\pi^2\\rho_{\\rm{e}}\\right)^{1\/3}\n,\\end{equation} and,\n\\begin{equation}\\label{dd_e1}\n\\Phi_{\\rm{e}}(t_{\\rm{e}})=t_{\\rm{e}}\\left(1+2t_{\\rm{e}}^2\\right)\\sqrt{1+t_{\\rm{e}}^2}-\\ln\\left(t_{\\rm{e}}+\\sqrt{1+t_{\\rm{e}}^2}\\right)\n,\\end{equation} \nwhere $m_{\\rm{e}}$ and $\\rho_{\\rm{e}}$ are the electron mass and density respectively.\nConsequently, $P_{\\rm{e}}(\\rho,\\delta)=\\rho_{\\rm{e}}\\mu_{\\rm{e}}-\\varepsilon_{\\rm{e}}(\\rho,\\delta)\n$, with $\n\\mu_{\\rm{e}}=[{k_{\\rm{e}}^2+m_{\\rm{e}}^2}]^{1\/2}\\approx k_{\\rm{e}}$ and $k_{\\rm{e}}=(3\\pi^2\\rho_{\\rm{e}})^{1\/3}$.\n\nShown in FIG.\\,\\ref{fig_ab_rhot} are our results for the core-crust transition density $\\rho_{\\rm{t}}$ (upper panel) and pressure $P_{\\rm{t}}$ (lower panel) as functions of the slope parameter $L(\\rho_{\\rm{r}})$ at a sub-saturation density $\\rho_{\\rm{r}}=0.11\\,\\rm{fm}^{-3}$.}\nIt is clear that the SRC\/HMT affects significantly the\ntransition density $\\rho_{\\rm{t}}$, e.g., a reduction as high as about\n58\\% with the HMT-exp parameter set compared to the FFG model. More quantitatively, the crust-core transition density in the FFG model is found to be about\n$\\rho_{\\rm{t}}\\approx0.086\\,\\rm{fm}^{-3}$, while that in the HMT\nmodels is about $\\rho_{\\rm{t}}\\approx0.079\\,\\rm{fm}^{-3}$\n(HMT-SCGF) and $\\rho_{\\rm{t}}\\approx0.036\\,\\rm{fm}^{-3}$ (HMT-exp), respectively.\nCorrespondingly, the transition pressure in the FFG model and the\nHMT-SCGF (HMT-exp) model is about\n$0.28\\,\\rm{MeV}\/\\rm{fm}^{3}$ and $0.26\\,\\rm{MeV}\/\\rm{fm}^{3}$\n($0.17\\,\\rm{MeV}\/\\rm{fm}^{3}$), respectively. \n\nAs discussed in the\nabove sections, the symmetry energy $E_{\\rm{sym}}(\\rho)$ at sub-saturation densities is reduced by the SRC\/HMT while its magnitude and slope at $\\rho_0$ are fixed at their currently known empirical values. Thus, the $E_{\\rm{sym}}(\\rho)$ is effectively hardened within the sub-saturation density region (see the inset of FIG.\\,\\ref{fig_ab_Esym}), leading to an enhancement of the slope\nparameter $L(\\rho)$ for $\\rho\\lesssim\\rho_0$. It is known that a larger $L(\\rho_{\\rm{r}})$ with $\\rho_{\\rm{r}}\\lesssim\\rho_0$\\,\\cite{Zha13} leads to a\nsmaller core-crust transition density as studied carefully in, e.g.,\nRefs.\\,\\cite{XuJ09}. The observed SRC\/HMT effects on the crust-core transition density and pressure are qualitatively consistent with expectations based on previous studies. \n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[height=2.7cm]{rhot.eps}\\\\\n \\hspace{0.cm}\n \\includegraphics[height=2.7cm]{Pt.eps}\n \\caption{Correlation between the core-crust transition density (upper) and the transition pressure (lower) in $\\beta$-stable neutron star matter\n and the slope parameter $L(\\rho_{\\rm{r}})$ at $\\rho_{\\rm{r}}=0.11\\,\\rm{fm}^{-3}$.}\n \\label{fig_ab_rhot}\n\\end{figure}\n\nAs we discussed earlier, the fundamentally important quantity that is growing from the FFG to HMT-SCGF then HMT-exp is the strength of the underlying isospin dependence of SRC\/HMT, namely, \nthe difference $\\delta x=x_{\\rm{HMT}}^{\\rm{SNM}}-x_{\\rm{HMT}}^{\\rm{PNM}}$ is increasing. Our results shown in FIG.\\,\\ref{fig_ab_rhot} indicates an anti-correlation between the $\\delta x$ and the transition density $\\rho_{\\rm{t}}$.\nIt provides an important connection to the study of the thickness of inner crusts\\,\\cite{XuJ09} as well as the related observational phenomena\\,\\cite{Gear11,Wen12,Newton12,Newton14,Hooker15,Zhou21}, such as the crustal oscillations and glitches as well as the $r$-mode stability window of rapidly rotating neutron stars. Our findings here may also have important implications on the heat transport process in neutron stars\\,\\cite{Cac08}.\n\nWithin the minimal model of neutron stars assuming no hadron-quark phase transition as well as baryon resonance and hyperon production, the compositions of neutron stars are \ndetermined by the requirement of charge neutrality and $\\beta$-equilibrium conditions. From the nucleon specific energy $E(\\rho,\\delta)$ we can calculate numerically the proton fraction $x_{\\rm{p}}(\\rho)$ for $\\beta$-stable matter and examine its dependence on the strength of SRC\/HMT. More specifically, for neutrino free $ \\beta $-stable matter, the chemical\nequilibrium for the reactions $\n\\rm{n}\\rightarrow\\rm{p}+\\rm{e}^{-}+\\overline{\\nu}_{\\rm{e}}$ and\n$\\rm{p}+\\rm{e}^{-}\\rightarrow \\rm{n}+\\nu _{\\rm{e}}$ requires $ \\mu\n_{\\rm{e}}=\\mu_{\\rm{n}}-\\mu_{\\rm{p}}\\approx\n4E_{\\rm{sym}}(\\rho)\\delta+8E_{\\rm{sym,4}}(\\rho)\\delta^3+12E_{\\rm{sym},6}(\\rho)\\delta^5+\\cdots$, here $E_{\\rm{sym},6}(\\rho)\\equiv 120^{-1}\\partial^6E(\\rho,\\delta)\/\\partial\\delta^6|_{\\delta=0}$ is the sixth-order symmetry energy.\nFor relativistically degenerate electrons, we have $ \\mu _{\\rm{e}}=[\nm_{\\rm{e}}^{2}+(3\\pi ^{2}\\rho x_{\\rm{e}})^{2\/3}] ^{1\/2}\\approx\n( 3\\pi ^{2}\\rho x_{\\rm{e}}) ^{1\/3} $, where\n$m_{\\rm{e}}\\approx0.511\\,$MeV is the electron mass, and\n$x_{\\rm{p}}=x_{\\rm{e}}$ because of charge neutrality. Above a certain density where $\\mu _{\\rm{e}}$ exceeds the muon\nmass $m_{\\mu }\\approx105.7\\,$MeV, the reactions\n$\\rm{e}^{-}\\rightarrow \\mu ^{-}+\\nu\n_{\\rm{e}}+\\overline{\\nu}_{\\mu }$, $\\rm{p}+\\mu ^{-}\\rightarrow \\rm{n}+\\nu _{\\mu }$ and $%\n\\rm{n}\\rightarrow\\rm{p}+\\mu ^{-}+\\overline{\\nu}_{\\mu }$ are\nenergetically allowed. Then, both electrons and muons are present\nin $\\beta $-stable matter. This alters the $\\beta$-stability condition\nto $\\mu_{\\rm{e}}=\\mu _{\\mu }=[ m_{\\mu }^{2}+(3\\pi ^{2}\\rho x_{\\mu\n})^{2\/3}] ^{1\/2}$ with $x_{\\rm{p}}=x_{\\rm{e}}+x_{\\mu }$. \nGenerally, when the nucleon momentum distribution function $n_{\\v{k}}^J(\\rho,\\delta)$ has a low- (high-) momentum depletion (tail), the nucleon chemical potential is not given by \n$k_{\\rm{F}}^{\\rm{n\/p},2}\/2M+U_{\\rm{n\/p}}(\\rho,\\delta,k_{\\rm{F}}^{\\rm{n\/p}})$, i.e., it could not be obtained by simply letting the momentum $|\\v{k}|$ to be the Fermi momentum in the single-nucleon potential. \nIn fact, determining the relation between the nucleon chemical potential and the single-nucleon potential for a general $n_{\\v{k}}^J(\\rho,\\delta)$ is known as a fundamental problem in nuclear many-body theories\\,\\cite{AGD1960}.\nWe thus have to calculate numerically the chemical potential $\\mu_J$ for a nucleon $J$ from the energy density $\\varepsilon$ of neutron star matter via\n$\n \\mu_J=\\partial\\varepsilon(\\rho,\\delta)\/\\partial\\rho_J.\n$\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[width=4.cm]{abMDI-xp.eps}\\quad\n \\includegraphics[width=4.05cm]{abMDI-sound.eps}\n \\caption{The proton fraction $x_{\\rm{p}}$ (left) and the square of the speed of sound $s^2$ (right) in neutron star matter within the FFG and HMT-exp model, respectively.\n The black circle and the cyan star correspond to the central densities of the neutron star in the two models.}\n \\label{fig_abMDI-xp}\n\\end{figure}\n\nShown in the left panel of FIG.\\,\\ref{fig_abMDI-xp} is the proton fraction $x_{\\rm{p}}$ in neutron star matter as a function of density within the FFG and HMT-exp model, respectively.\nSince the HMT-SCGF result is similar to the one using the FFG model, we thus in the following focus on the comparison between the predictions using the FFG and the HMT-exp models.\nObviously, the most important observation is that due to the reduction of the high-density symmetry energy in the HMT-exp model the proton fraction $x_{\\rm{p}}$ in this model starts to decrease quickly above some critical density. \nIt thus prompts the role played by the symmetry energy in the core of neutron stars as $\\delta$ in the HMT-exp model is much closer to 1 than that in the FFG model.\nTherefore, it may have important ramifications on properties of neutron stars. For example, it is well known that the $x_{\\rm{p}}$ determines the cooling mechanisms of protoneutron stars and the associated neutrino emissions\\,\\citep{LPPH}. \nMore quantitatively, the critical proton fraction $x^{\\rm{DU}}_{\\rm{p}}$ enabling the direct URCA process (DU) for fast cooling is given by\n$x^{\\rm{DU}}_{\\rm{p}}=1\/[1+(1+x_{\\rm{e}}^{1\/3})^3]$\nwith the $x_{\\rm{e}}\\equiv \\rho_{\\rm{e}}\/\\rho_{\\rm{p}}$ between 1 and 0.5 leading to $x^{\\rm{DU}}_{\\rm{p}}$ between 11.1\\% to 14.8\\%\\,\\citep{Kla06}, see the red dashed lines in the left panel of FIG.\\,\\ref{fig_abMDI-xp}.\nThe black circle and the cyan star in the left panel of FIG.\\,\\ref{fig_abMDI-xp} correspond to the central densities of neutron stars in the two models studied here. It is seen that around these central densities the direct URCA is allowed in the FFG model but forbidden in the HMT-exp model. \n\nThe speed of sound squared $s^2=\\d P\/\\d\\varepsilon$ is a measure of the stiffness of nuclear EOS. Shown in the right panel of FIG.\\,\\ref{fig_abMDI-xp} are the predicted $s^2$ by the two models considered. It is seen that the EOS in the HMT-exp model is much softer than the FFG model prediction at densities above about $2.5\\rho_0$. Since the ANM EOS $E(\\rho,\\delta)\\approx E_0(\\rho)+E_{\\rm{sym}}(\\rho)\\delta^2+\\cdots$, the reduced symmetry energy but increased $\\delta^2$ at high densities together may effectively soften the EOS. However, its net effects depend on how strongly the SNM EOS $E_0(\\rho)$ is stiffened by the SRC\/HMT. The results shown here indicate clearly that the SRC\/HMT effect on the symmetry energy is winning against that on the SNM EOS. \nThis result is expected to have some dynamical effects on properties of neutron stars. Since the symmetry energy and SNM EOS in different density regions have different effects on properties of neutron stars, the SRC\/HMT effects through these two terms of the ANM EOS have to be studied quantitatively as we shall do next. \n\nThe mass-radius correlation of neutron stars is obtained from integrating the Tolman-Oppenheimer-Volkoff (TOV) equations\\,\\cite{Misner1973}\n\\begin{align}\n\\frac{\\d P(r)}{\\d r}=&-\\frac{[\\varepsilon(r)+P(r)][M(r)+4\\pi r^3P(r)]}{r[r-2M(r)]},\\\\\n\\frac{\\d M(r)}{\\d r}=&4\\pi r^2\\varepsilon(r),\n\\end{align}\nwhere $r$ is the radial distance from the center of the star, and\n$M(r)$ is the mass enclosed within $r$ (adopting natural unit in which $c=G=1$).\nThe pressure of $\\beta$-stable and charge neutral npe$\\mu$ matter is given by $\nP(\\rho,\\delta)=P_{\\rm{N}}(\\rho,\\delta)\n+P_{\\rm{e}}(\\rho,\\delta)+P_{\\mu}(\\rho,\\delta)$,\nwhere $P_{\\rm{N}}(\\rho,\\delta) $ is the nucleon pressure. \nThe lepton pressure is further given as $\nP_\\ell(\\rho,\\delta)=\\rho_\\ell\\mu_\\ell-\\varepsilon_\\ell(\\rho,\\delta)\n$, with $\n\\mu_\\ell=[{k_\\ell^2+m_\\ell^2}]^{1\/2}$ and $k_\\ell=(3\\pi^2\\rho_\\ell)^{1\/3}$ for $\\ell=\\rm{e},\\mu$, here $\\varepsilon_{\\ell}(\\rho,\\delta)=\\eta_{\\ell}\\Phi_{\\ell}(t_{\\ell})$ (see (\\ref{dd_e1}) for the definition of $\\Phi_{\\ell}$).\nFor the whole system, we have the self-consistency relation between the total pressure $P$ and the total energy density $\\varepsilon$, i.e., \n$\nP=\\rho^2{\\d(\\varepsilon\/\\rho)}\/{\\d\\rho}$.\nThe inner crust of neutron stars with densities ranging between $\\rho_{\\text{out}}=2.46\\times\n10^{-4}\\,\\rm{fm}^{-3}$ corresponding to the neutron dripline and the\ncore-crust transition density $\\rho _{\\text{t}}$ is the region where\nsome complex and exotic structures\\,---\\,the ``nuclear pasta'' may exist. Because of our very poor knowledge\nabout this region we adopt the polytropic EOSs expressed in\nterms of the pressure $P$ as a function of the total energy density\n$\\varepsilon$ as\n$P=c+d\\varepsilon^{4\/3}$\\,\\cite{XuJ09,Hor03}. The constants $c$ and\n$d$ are determined by the quantities at $\\rho\n_{\\text{t}}$ and $\\rho _{\\text{out}}$\\,\\cite{XuJ09}, i.e.,\n\\begin{equation}\nc=\\frac{P_{\\rm{out}}\\varepsilon_{\\rm{t}}^{4\/3}-P_{\\rm{t}}\\varepsilon_{\\rm{out}}^{4\/3}}{\\varepsilon_{\\rm{t}}^{4\/3}-\\varepsilon_{\\rm{out}}^{4\/3}},\n~~\nd=\\frac{P_{\\rm{t}}-P_{\\rm{out}}}{\\varepsilon_{\\rm{t}}^{4\/3}-\\varepsilon_{\\rm{out}}^{4\/3}}\n.\\end{equation}\nFor the outer\ncrust\\,\\cite{BPS71,Iida1997}, we use the conventional Baym-Pethick-Sutherland (BPS) EOS for the region with $6.93\\times 10^{-13}\\,\\rm{fm}^{-3}\\lesssim\\rho \\lesssim\\rho _{\\text{out}}$ and the\nFeynman-Metropolis-Teller (FMT) EOS for $4.73\\times 10^{-15}\\,\\rm{fm}^{-3}\\lesssim\\rho \\lesssim6.93\\times\n10^{-13}\\,\\rm{fm}^{-3}$, respectively.\nThe total neutron star matter EOSs in the two models are shown in FIG.\\,\\ref{fig_abMDI-beta}.\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[width=6.5cm]{abMDI-beta.eps}\n \\caption{The EOS of neutron star matter in the FFG model and the HMT-exp model. See the text for details.}\n \\label{fig_abMDI-beta}\n\\end{figure}\n\nThe resulting mass-radius correlation of neutron stars is shown in FIG.\\,\\ref{fig_abMDI-MR} for the FFG and the HMT-exp model, respectively. Firstly, we notice that both models give a radius of about 12\\,km for canonical neutron stars of mass 1.4$M_{\\odot}$ consistent with many analyses of the GW170817 observation, see, e.g., Refs.\\,\\cite{EPJA-review,Baiotti,Capano20,David,Kat20,AngLi} for reviews. \nThis is mostly by design as our EDF parameters are fixed such that all the EOS characteristics at $\\rho_0$ are consistent with existing constrains from both astrophysical observations and terrestrial experiments. \nSecondly, it is interesting to see that the two models indeed predict significantly different maximum masses and the corresponding radii. More quantitatively, the maximum mass of neutron stars in the FFG model is about $M_{\\rm{NS}}^{\\max}\\approx1.98M_{\\odot}$, with the corresponding radius being about 9.9\\,km. While the $M_{\\rm{NS}}^{\\max}$ in the HMT-exp model is about $1.72M_{\\odot}$ with a radius of about 10.3\\,km. \nThe reduction of the maximum mass $M_{\\rm{NS}}^{\\max}$ due to the SRC\/HMT is thus about 13\\%. \nThe point corresponding to the maximum mass on the $P(\\varepsilon)$ plane is also shown in FIG.\\,\\ref{fig_abMDI-beta}, by the black circle (cyan star) for the FFG (HMT-exp) model.\nThirdly, since the HMT largely affects the transition density $\\rho_{\\rm{t}}$ (see FIG.\\,\\ref{fig_ab_rhot}), the radii of low-mass neutron stars in the FFG model and the HMT-exp model show obvious differences.\n\nThe central density $\\rho_{\\rm{cen}}$ corresponding to $M_{\\rm{NS}}^{\\max}$ is about $\\rho_{\\rm{cen}}^{\\rm{FFG}}\\approx7.3\\rho_0$ ($\\rho_{\\rm{cen}}^{\\rm{HMT}\\mbox{-}\\rm{exp}}\\approx7.9\\rho_0$) for the FFG (HMT-exp) model (see the points shown in FIG.\\,\\ref{fig_abMDI-xp}).\nAlthough the EDF constructed is non-relativistic in nature, it could be effectively applied to neutron stars and fulfill the principle of causality (set by the speed of sound $s^2=\\d P\/\\d\\varepsilon=1$).\nIn particular, we find that the square of the speed of sound in the FFG model at $\\rho_{\\rm{cen}}^{\\rm{FFG}}$ is about $s^2_{\\rm{FFG}}\\approx0.84$, while that for the HMT-exp model is about $s_{\\rm{HMT}\\mbox{-}\\rm{exp}}^2\\approx0.35$, see the right panel of FIG.\\,\\ref{fig_abMDI-xp} for the density dependence of the $s^2$ in the two model. Thus, due to the softening of the symmetry energy above a critical density and the non-relativistic nature of the Gogny-like EDF, the high-density $s^2$ in the HMT-exp model is \nmuch smaller than that in the FFG model. In fact, it may not exceed 1 at even larger densities when the FFG becomes acausal (the density violating the principle of causality in the FFG model is about $9.8\\rho_0$).\n\n\\begin{figure}[h!]\n\\centering\n \n \\includegraphics[width=6.5cm]{abMDI-MR.eps}\n \\caption{The mass-radius relation for neutron stars in the FFG and the HMT-exp models.}\n \\label{fig_abMDI-MR}\n\\end{figure}\n\nTo this end, it is interesting to note that some earlier studies on the mass-radius relation of neutron stars adopting the nonlinear RMF models found that the SRC\/HMT generally increases the maximum mass $M_{\\rm{NS}}^{\\max}$\\,\\cite{Cai16b,Lou22a,Lu2022,Hong22,Lou22,Lu21,Souza20}. In these models the SRC\/HMT induced increase in SNM pressure dominates over the reduction of symmetry energy. \nOne main reason is that in the nonlinear RMF models the symmetry energy itself plays a minor role in determining the maximum mass $M_{\\rm{NS}}^{\\max}$ as pointed out already a long time ago\\,\\cite{Ser79}. Moreover, the softening of the $E_{\\rm{sym}}(\\rho)$ in the RMF models is weaker than that in the current (non-relativistic) Gogny-like EDF. Of course, this then raises an interesting question: how can this kind of EDFs encapsulating the SRC\/HMT effects support the currently observed most massive neutron star PSR J0740+6620 which has a mass of $2.08^{+0.07}_{-0.07}M_\\odot$\\,\\cite{Fon21}? In fact, even without considering the SRC\/HMT effects, it has been a challenging problem to find mechanisms to support heavy neutron stars (with masses around and above $2M_{\\odot}$) within non-relativistic energy density functionals, see, e.g., Refs.\\,\\cite{Rios-G,Zha16,ZhouY2019,Gon18,Gon19,Lou20}. Our results presented above indicate that this problem may become more challenging and certainly deserves further investigations.\n\n\\setcounter{equation}{0}\n\\section{Summary and Outlook}\\label{sec6}\n\nIn summary, we investigated SRC\/HMT effects on the EOS and single-nucleon potential in cold neutron-rich matter within the Gogny-like EDF. The SRC effects are incorporated into the EDF by using a single-nucleon\nmomentum distribution function $n_{\\v{k}}^J(\\rho,\\delta)$ that has an isospin-dependent high-momentum (low-momentum) tail (depletion) with its parameters determined by the available SRC experiments. We introduced a parametrization as a surrogate for the momentum-dependent kernel of the Gogny-like EDF to facilitate derivations of analytical expressions for all relevant physical quantities describing the EOS and single-nucleon potential in ANM. The surrogate is shown to be effective for problems with momentum (density) scale smaller than about 1 GeV\/c (several times the saturation density). The resulting expressions for all terms in the ANM EOS and single-nucleon potential will facilitate further explorations of nuclear interactions in dense neutron-rich matter as well as their impacts on properties of neutron stars and heavy-ion reactions induced by high-energy radioactive beams. \nIn particular, \n\\begin{enumerate}\n\\item[(a)]We have derived the analytical expressions for the EOS of SNM $E_0(\\rho)$, the corresponding pressure $P_0(\\rho)$, the incompressibility coefficient $K_0(\\rho)$, the nucleon effective k-mass $M_0^{\\ast}(\\rho)$, the isoscalar potential $U_0(\\rho,|\\v{k}|)$, the symmetry energy $E_{\\rm{sym}}(\\rho)$ together with its slope parameter $L(\\rho)$, and the isovector (symmetry) potential $U_{\\rm{sym}}(\\rho,|\\v{k}|)$.\nNumerical constructions based on these analytical expressions are given through three models for the nucleon momentum distribution, namely the FFG model without the SRC-induced HMT in the $n_{\\v{k}}^J(\\rho,\\delta)$, the\nHMT-SCGF\/HMT-exp model with different properties of the HMT, to fulfill certain empirical constraints on the EOS of ANM. All three models contain proper potential parts.\n\\item[(b)] While the same set of available empirical constraints on the ANM EOS at $\\rho_0$ are all maintained in the three models, the EOS $E_0(\\rho)$ of SNM at high densities is made slightly harder once the SRC\neffects are considered. In addition, the symmetry energy in the presence of HMT becomes softer at large densities even to start decreasing above a critical density. Although qualitatively consistent with the nonlinear RMF model predictions, the reduction of the high-density symmetry energy $E_{\\rm{sym}}(\\rho)$ is much stronger while the enhancement of the SNM pressure $P_0$ is much weaker in the Gogny-like EDF with SRC\/HMT. Consequently, as a result of the competition between these two effects, the maximum mass of a cold neutron star is found to be reduced compared with the prediction of the Gogny-like EDF without considering the SRC\/HMT.\n\\item[(c)] The single-nucleon potential $U_0(\\rho,|\\v{k}|)$ in SNM at $\\rho_0$ in all three models is consistent with the isoscalar nucleon optical potential from earlier analyses of various experimental data. The nucleon isovector (symmetry) potential $U_{\\rm{sym}}(\\rho,|\\v{k}|)$ in ANM is largely enhanced due to the corresponding enhancement of the potential symmetry energy $E_{\\rm{sym}}^{\\rm{pot}}(\\rho)$.\nThese are the direct consequences of the momentum-dependence of the nucleon potential in the Gogny-like EDF due to the finite ranges of nuclear interactions. Such momentum-dependence is absent in the nonlinear RMF approach.\n\\end{enumerate}\n\nThe formulas, table of coupling constants, parameterizations and quantitative demonstrations of the relevant physical quantities given in this work provide a useful and convenient starting point to further investigate the nature of strong interactions at short distances and novel effects of SRC\/HMT on properties of dense neutron-rich matter. In particular, the generally density- and momentum-dependent single-nucleon potential in neutron-rich matter is expected to play a significantly role in understanding the dynamics and observables of heavy-ion reactions induced by high-energy radioactive beams, the cooling mechanisms of protoneutron stars as well as properties of cold neutron stars at $\\beta$-equilibrium and their mergers. For instance, with the coupling parameters given in Table \\ref{tab_para}, one can use the single-nucleon potential of Eq. (\\ref{Gen-U}) in simulating heavy-ion reaction with the $f_J(\\v{r},\\v{k})$ self-consistently generated by the BUU transport model. After comparing the predicted observables with experimental data, one can then make better informed predictions for the EOS of hot, dense and neutron-rich matter. The latter is a necessary input for simulating mergers of two neutron stars based on principles of general relativity \\cite{Ra1,Ra2}. In particular, it is important for understanding the high-frequency spectrum of gravitational waves from post-mergers that carry critical information about the nature of possible phase transitions and the separation boundary\/gap between neutron stars and blackholes. Interestingly, it has already been found that the density dependence of nuclear symmetry energy play a significant role in answering many intriguing questions in simulating neutron star mergers \\cite{Most21}. \n\nAs discussed earlier, our approach has some caveats and some challenging issues remain to be resolved. Fortunately, multimessenger nuclear astrophysics with high-precision X-ray and gravitational wave detectors as well as advanced rare isotope beam facilities are expected to provide us with various kinds of new and more accurate observational data. Combined analyses of these data using the same Gogny-like EDF with SRC\/HMT may help us address some of the remaining issues. For example, the strong enhancement of the symmetry potential $U_{\\rm{sym}}(\\rho,|\\v{k}|)$ due to the SRC\/HMT especially at supra-saturation densities, e.g., $\\rho\\approx2\\rho_0\\mbox{$\\sim$}3\\rho_0$, is expected to affect significantly the dynamics and some observables of nuclear reactions with high-energy radioactive beams. While the corresponding reduction of the total symmetry energy reduces significantly the maximum mass of neutron stars that the resulting EOS can support. A joint analysis of these effects using a unified Gogny-like EDF with SRC\/HMT is expected to be fruitful. Besides the traditionally used forward predictions with the different model parameter sets, the expected new data and existing ones together will enable us to get new insights from inferences of relevant model parameters using machine leaning techniques, such as the Bayesian model selection \\cite{MAL}. These studies are on the top of our working agenda. \n\n\\section*{Acknowledgement} We would like to thank Lie-Wen Chen, Xiao-Tao He, Che Ming Ko and Xiao-Hua Li for useful discussions over the years on some of the issues studied in this work.\nThis work is supported in part by the U.S. Department of Energy, Office of Science,\nunder Award No. DE-SC0013702, the CUSTIPEN (China-\nU.S. Theory Institute for Physics with Exotic Nuclei) under\nUS Department of Energy Grant No. DE-SC0009971.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}