diff --git "a/data_all_eng_slimpj/shuffled/split2/finalzzmbxl" "b/data_all_eng_slimpj/shuffled/split2/finalzzmbxl" new file mode 100644--- /dev/null +++ "b/data_all_eng_slimpj/shuffled/split2/finalzzmbxl" @@ -0,0 +1,5 @@ +{"text":"\\section{INTRODUCTION}\n\\label{secz:intro}\nThe pion--nucleus interaction on and below the $\\Delta_{33}$ resonance\nhas been extensively studied, and microscopic models of elastic scattering\ndo a reasonable job \\cite{john92} of describing the data.\nThe pion--nucleus interaction at energies above the $\\Delta_{33}$ resonance\nhas received \\cite{rokn88,will89,marl84,hipi} much less attention. In this\nenergy region (300 MeV $\\le T_\\pi \\le $ 1 GeV), the pion has a much shorter\nwavelength. For example, the wavelength at resonance is about 4 fm (about\nthe size of the nucleus)\nwhile at 1 GeV the wavelength is 1 fm (about the size of a single nucleon).\nThe shorter wavelength implies that elastic and inelastic data at the\nhigher energies will provide sensitivity to the details of the spatial\ndependence of the ground--state and transition densities. The data may also\nprove sensitive to modifications \\cite{sieg84,brow88} of the properties of\nthe nucleon in the nuclear medium, or reveal significant contributions from\nexchange currents \\cite{jiang92}.\n\nMoreover, the pion--nucleon two--body interaction becomes much weaker as\none goes to energies above the $\\Delta_{33}$ resonance.\nThe two--body total cross section becomes less than 30 mb, which is about 15\n\\% of that on the $\\Delta_{33}$ resonance. The first implication of this\nweaker amplitude is that the pion is able to penetrate deeper into the\nnucleus. A simple estimate \\cite{erns79} from total reaction cross\nsection studies shows that a projectile can penetrate into a target to\na radius that is equal to the impact parameter at which the profile\nfunction equals one mean free path. In Fig.\\ \\ref{fig1} the profile\nfunction for $^{12}$C, $^{40}$Ca and $^{208}$Pb are pictured, and the\narrows indicate approximately how far into a nucleus a pion of the\nlabeled energy can penetrate. The pion in the energy region from 500\nMeV to 1 GeV is one of the most penetrating of the strongly interacting\nparticles.\n\nThe second implication of the weaker two-body cross section is\nthat multiple scattering theory for the optical potential becomes\nincreasingly convergent. A simple estimate of the convergence\nis obtained by comparing a typical second--order term in the optical\npotential to the first--order term. For the case of short--range\ncorrelations, one obtains \\cite{erns77}\n\\begin{equation}\nR\\equiv {U^{(2)}\\over U^{(1)}}=\\sqrt{\\sigma}{\\ell_c\\over k}\\rho\\,\\,\\,,\n\\end{equation}\nwhere $\\sigma$ is the total two--body cross section, $\\ell_c$ is the\ncorrelation length, and $\\rho$ is the nuclear density at which the\npion interacts (See Fig. 1). The\nfactor of $1\/k$, where $k$ is the incident pion momentum, reflects the\nsuppression of the pion propagator in $U^{(2)}$ with increasing energy. On\nresonance, we find $R\\approx 0.1$, with $R < 1$ only\nbecause $\\rho$ is so small there. At 500 MeV, we\nfind $R\\approx 0.04$ and at 1 GeV, $R\\approx 0.02$.\nThus differences between the data and a carefully calculated result\nutilizing a first--order optical potential would be a strong indication\nof the presence of unconventional phenomena such as mesonic current\ncontributions, modified nucleon properties in the medium, or other yet to be\nthought of effects.\n\nOne option for calculating high--energy\npion scattering is a momentum--space optical--potential approach. This\npresents the\nopportunity \\cite{gieb88} to\ncalculate the scattering from a lowest--order optical potential\nincluding the following features 1.) exact\nFermi--averaging integration, 2.) fully covariant kinematics \\cite{erns80},\nnormalizations, and phase--space factors, 3.) invariant\namplitudes \\cite{gieb88,erns90} and 4.) finite--range, physically motivated\ntwo--body off--shell \\cite{erns90} amplitudes. Within the multiple\nscattering theory developed in Refs.\\ \\cite{john92,john83} this is an exact\ncalculation of the lowest--order optical potential. A brief review of\nthis approach will be given in Sec.\\ \\ref{sec:formal}.\n\nA practical difficulty arises when one wants to use the momentum--space\noptical model to study high--energy pion--nucleus scattering. As the\nprojectile energy increases, more partial waves are needed in the\npion--nucleon\ntwo--body amplitudes. Below 300 MeV, only S--waves and P--waves are needed.\nAbove 300 MeV, D--waves become important. F--waves become significant\nabove 500 MeV and G--waves and H--waves above 700 MeV. At the same time\nthe number of pion--nucleus partial waves is increasing at a rate\nproportional to the pion momentum. At the high energies the momentum--space\napproach becomes prohibitively computer--intensive, and one would like to\nsearch for a simpler alternative.\n\nThe semi--classical theory immediately comes to mind as an alternative to\nthe momentum--space approach. It is often used at high energies not only\nbecause it is easier to compute, but also because the simpler character\nof the theory facilitates obtaining physical insights into the reaction.\nHowever, the semi--classical theory is\nexpected to be a good approximation only for local potentials and only when\nthe wavelength is sufficiently short. Because the exact pion--nucleus\noptical potential appears to be highly\nnonlocal, we examine the role that nonlocalities play in\nhigh--energy pion--nucleus scattering. This is done numerically in momentum\nspace and is presented in Sec. \\ref{subsec:opt}. We undertake this study\nas a first step in obtaining a more quantitative measure of the validity\nof the semi--classical theory than currently exist in the literature.\n\nIn Sec. \\ref{subsec:eikon} we present our semi--classical model \\cite{semi}.\nIn\norder to have a quantitative as well as a simple model for confronting data, we\ninclude the Coulomb interaction and the first Wallace correction\n\\cite{wallace}.\nWe establish the validity of the semi--classical model both by examining\nthe size of the Wallace correction and by comparing it to the model--exact\nresults obtained in momentum space. Comparisons of the eikonal model to\nmodel--exact calculations are made in Sec.\\ \\ref{sec:results}.\n\nA previous similar investigation can be found in Ref.\\ \\cite{seki}.\nThere a factorized approximation \\cite{pipit} was used in the\nmomentum--space calculations, so that the full nonlocality of the\noptical potential was not considered. Furthermore, Glauber multiple scattering\ntheory\\ \\cite{glauber} was proposed, whereas we will examine an even\nsimpler model, that of a local optical potential with the scattering\nsolved via an eikonal approximation. A more detailed discussion that\ncompares and contrasts our results with previous work is given in\nSec.\\ \\ref{sec:comp}.\nA summary, conclusions and future prospects are presented in Sec.\\\n\\ref{sec:conc}.\n\\section{SCATTERING THEORY}\n\\label{sec:formal}\nAt the higher--energies where the pion--nucleon amplitude\nbecomes a smooth function of energy and is much weaker than on resonance,\nthe first--order optical potential may well be adequate for describing\nall of the conventional nuclear physics phenomena that enter into the\nreaction dynamics. Whether this is so remains for further investigation.\nIndependent of this question, there is the question of whether simpler\napproaches to the calculation of the lowest--order optical potential\nand the scattering from this potential are reasonably quantitative. It is\nthis question upon which we concentrate here. Below we first review the\nmomentum--space optical potential which serves as our `model--exact'\ncalculation. In momentum--space, we examine the importance of the\nnonlocalities caused by both the resonance propagation and the finite\nrange of the two--body amplitude. Finding that these both can be accurately\napproximated, we present a simple eikonal model.\n\nBoth the optical and the eikonal models use the same target\nwave functions, which are obtained from Hartree--Fock calculations\\\n\\cite{hf,bein75} with spurious center--of--mass motion removed as discussed in\n\\ref{sec:dcm}. They also use the same on--shell pion--nucleon two--body\namplitudes,\nwhich are constructed from Arndt's\\ \\cite{arndt} and H\\\"ohler's\\\n\\cite{hohler} phase--shifts.\n\\subsection{Optical potential}\n\\label{subsec:opt}\nA complete description of the momentum--space optical potential that we use\ncan be found in Ref.\\ \\cite{gieb88}, and a detailed discussion of\nthe covariant kinematics that we use can be found in Refs.\\ \\cite{erns80}.\nThe formal multiple--scattering theory in which this work is embedded is\ngiven in Refs.\\ \\cite{john92,john83}. Here we provide only a brief\noverview.\n\nThe first--order optical potential in the impulse approximation can be\nwritten as,\n\\begin{eqnarray}\n\\langle \\vec k'_\\pi\\vec k'_A\\,\\vert\\,U_{\\rm mo}(E)\\,\\vert\\,\n\\vec k_\\pi\\vec k_A\\rangle\n=\\sum_\\alpha \\int&&\\,{d^3 k_{A-1}\\over 2\\bar E_{A-1}}\\,{d^3 k'_N\\over 2\\bar\nE'_N}\\,{d^3 k_N\\over 2\\bar E_N}\\,\n\\langle\\Psi'_{\\alpha ,\\vec\nk'_A}\\,\\vert\\,\\vec k'_N\\vec k_{A-1}\\rangle\\nonumber \\\\\n&&~~~~~~\\times\\,\\langle \\vec k'_\\pi\n\\vec k'_N\\,\\vert\\,t(E)\\,\\vert\\,\\vec k_\\pi \\vec k_N \\rangle\\langle \\vec\nk_N\\vec k_{A-1}\\,\\vert\\,\\Psi_{\\alpha ,\\vec k_A}\\rangle\\,\\,,\n\\label{eq:optical}\n\\end{eqnarray}\nwhere $\\vec k_\\pi$, $\\vec k_N$ and $\\vec k_{A-1}$ are the momenta of the\npion, the struck nucleon and the $A-1$ residual nucleons in the\npion--nucleus center--of--momentum frame\nrespectively. $E$ is the incident energy of the pion and the nucleus\n(in the pion--nucleus center--of--momentum frame also), and $\\alpha$\nis a set of quantum numbers that specifically label the nuclear bound state.\n\nThe target wave function $\\langle \\vec k_N\\vec k_{A-1} \\,\\vert\\,\n\\Psi_{\\alpha ,\\vec k_A}\\rangle$ defined covariantly \\cite{gieb88} contains\na momentum--conserving delta function $\\delta (\\vec k_A - \\vec k_N - \\vec\nk_{A-1})$ and is a function of the relative momentum between the nucleon\nand the $A-1$ residual nucleons. The pion--nucleon t-matrix similarly\ncontains a momentum--conserving delta function and is a function\nof the relative momentum between the pion and the nucleon, $\\vec \\kappa$,\n\\begin{equation}\n\\langle \\vec k'_N \\vec k'_\\pi \\,\\vert\\,t(E) \\,\\vert\\, \\vec k_N \\vec\nk_\\pi \\rangle =\n\\delta (\\vec k'_N + \\vec k'_\\pi - \\vec k_N-\\vec k_\\pi)\n\\langle \\vec \\kappa' \\,\\vert\\, t[\\omega_\\alpha (E,\\vec k_\\pi\n,\\vec k_N )] \\,\\vert\\, \\vec \\kappa \\rangle\\,.\n\\end{equation}\nOne of the these three momentum--conserving delta functions leads to overall\nmomentum\nconservation; the others allow one to perform two of the three\nintegrals in Eq.~\\ref{eq:optical}. The remaining integral is the Fermi\naveraging integration which\nmust be performed numerically. The details of how we perform the integration\ncan be found in \\cite{gieb88}.\n\nThe energy at which one evaluates the two--body t--matrix, $\\omega_\\alpha\n(E, \\vec k_\\pi, \\vec k_N )$, must be carefully chosen \\cite{erns85} if a\nconvergent perturbation theory is to result for calculations at and below the\n$\\Delta{33}$ resonance region. The energy $\\omega_\\alpha$ is defined\ncovariantly by\nfirst defining the energy available to the pion--nucleon two--body subsystem\n\\begin{equation}\nE_{\\pi N}=E -\\sqrt{(\\vec k_\\pi+\\vec k_N)^2+m_{A-1}^2}\\,\\,,\n\\end{equation}\nand then defining the invariant center-of-momentum energy for this system\n\\begin{equation}\n\\omega_\\alpha^2=E_{\\pi N}^2-(\\vec k_\\pi+\\vec k_N)^2\\,\\,.\n\\label{eq:cmen}\n\\end{equation}\nThe mass of the $A-1$ system, $m_{A-1}$, differs from the mass of the\n$A$--body target, $m_A$, by a nucleon mass and a binding energy,\n$m_A=m_{A-1}+m_N+E_b$. This energy must then be shifted by a `mean-spectral\nenergy', $E_{ms}$, which is a calculated number \\cite{erns85} that\napproximately accounts for the interaction of the intermediate $\\Delta_{33}$\n(or the intermediate nucleon or other hadronic resonance) with the residual\nnucleus. This first--order potential produces results for energies\nnear the $\\Delta_{33}$ resonance that are in remarkable agreement with the\ndata. This is because \\cite{john92} the sum of the second--order\neffects (Pauli exclusion, true absorption and correlation corrections)\ncancel amongst themselves. The use of invariant normalizations,\ninvariant phase\nspace and invariant amplitudes produces the phase-space factors that are\npresent in Eq. \\ref{eq:optical}.\n\nThe optical potential so defined is then inserted into the Klein--Gordon\nequation which is solved numerically to produce our model--exact calculation.\n\n\\subsection{Eikonal model}\n\\label{subsec:eikon}\nThe eikonal model we propose to examine results from first replacing\nthe optical potential of Eq.\\ \\ref{eq:optical} by a local potential and\nthen solving for the scattering amplitude arising from the use of the local\npotential and the eikonal propagator. In order to include the Coulomb\ninteraction, we divide the scattering amplitude as\n\\begin{equation}\nF(q)=F_{pt}(q)+F_{CN}(q)\n\\end{equation}\nwhere $q$ is the momentum transferred to the pion, $\\vec q=\\vec k'_\\pi-\\vec\nk_\\pi$, $F_{pt}$ is the point Coulomb scattering amplitude,\nand $F_{CN}$ is the scattering amplitude calculated from the sum of the\nstrong nuclear\npotential plus a Coulomb correction, which is the difference between the\npoint Coulomb interaction and the Coulomb interaction of a finite charge\ndistribution. The eikonal approximation gives for $F_{CN}(q)$\n\\begin{equation}\nF_{CN}(q)=ik_\\pi \\int_0^\\infty\nbdbJ_0(qb)e^{i\\chi_{pt}(b)}\\Gamma_{CN}(E,b),\n\\end{equation}\nwhere $k_\\pi$ is the momentum of the incident pion in the\npion--nucleus center--of--momentum frame, $b$ is the impact parameter,\n$\\chi_{pt}$ is the point Coulomb phase (in the eikonal approximation), and\n$\\Gamma_{CN}$ is the profile function defined below.\n\nIn order to incorporate the distortion and energy shift caused by the\nCoulomb interaction, we follow the prescription given in Ref.\\ \\cite{coul}\nand write the profile function as\n\\begin{eqnarray}\n\\Gamma_{CN}(E,b)&&=1-{\\rm exp}\\lbrace\ni\\chi_{CN}(E,b)\\rbrace\\nonumber \\\\\n&&=1-{\\rm exp}\\lbrace i\\chi_N\\lbrack E,\nb(1+EV_c(b)\/k_\\pi^2)\\rbrack\n+i\\chi_c(b)-i\\chi_{pt}(b)\\rbrace,\n\\end{eqnarray}\nwhere $V_c$ is the Coulomb potential of a uniformly\ncharged sphere,\n$\\chi_c$ is the phase shift caused by that potential, and the nuclear phase\nshift $\\chi_N$ is written as\n\\begin{equation}\n\\chi_N(E,b)=-{1\\over 2k_\\pi}\\,\\int_{-\\infty}^\\infty dz\\,\nU_{\\rm eik}\\lbrack E-V_c(r),r\\rbrack,\n\\end{equation}\nwhere\n\\begin{equation}\nr=(b^2+z^2)^{1\/2},\n\\end{equation}\nand $U_{\\rm eik}$ is the strong potential obtained as follows.\nFirst we construct a local potential $U_o$ from\nthe pion--nucleon two--body scattering amplitude and the target density by\n\\begin{equation}\nU_o(E,r)= -4\\pi Z\\lbrack f_p(0)\\rho_p +{f'_p(0)\\over\n2k_\\pi^2}\\nabla^2\\rho_p\\rbrack\n-4\\pi N\\lbrack f_n(0)\\rho_n+{f'_n(0)\\over\n2k_\\pi^2}\\nabla^2\\rho_n\\rbrack,\n\\label{eq:eiku}\n\\end{equation}\nwhere $Z$ and $N$ are the proton number and the neutron number\nrespectively. The scattering amplitude $f_p(0)$ [$f_n(0)$] is the\npion--proton [pion--neutron] scattering amplitude in the forward direction\nin the pion--nucleus center--of--momentum frame. The finite range of the\npion--nucleon interaction is included through lowest nonvanishing order\nand produces the additional terms in Eq.\\ \\ref{eq:eiku} which are\nproportional to the derivatives of the scattering amplitude as shown in\nAppendix \\ref{sec:dir}.\n\nIn order to investigate the importance of the lowest--order corrections to\nthe eikonal propagator as derived by Wallace \\cite{wallace}, next we\nreplace the potential $U_o (E,r)$ by the potential\n\\begin{equation}\nU_{\\rm eik}(E,r)=U_o+{U^2_o\\over 4k_\\pi^2}\\Big( 1+{2b^2\\over r}{d\\over dr}\n{\\rm ln}U_o\\Big).\n\\label{eq:wallace}\n\\end{equation}\nThe details of the derivation of this correction and higher order\ncorrections can be found in Ref.~\\cite{wallace}.\n\\section{ RESULTS}\n\\label{sec:results}\nIn this section we investigate the importance of some of the ingredients of\nour model--exact calculation of high--energy pion--nucleus scattering. We\nwill then examine how well our simple eikonal theory\nreproduces the results of the model--exact theory.\n\n\\subsection{Contributions from nonlocalities in the optical potential}\n\\label{subsec:nonloc}\nThe most difficult part of evaluating the optical potential in the\nmodel--exact theory is performing the Fermi--averaging integration. This\nis known to be very important in the region of the $\\Delta_{33}$ resonance,\nwhere proper accounting for the recoil and propagation of the pion--nucleus\nresonance requires a careful treatment of the dependence of the\nenergy $\\omega_\\alpha$ in Eq.~\\ref{eq:cmen} on the momentum of the nucleon.\nResonance propagation manifests itself as a nonlocality in the optical\npotential, and this represents the first source of nonlocality whose\nimportance we want to assess for high--energy pions.\n\nTo investigate this we calculate elastic scattering of $\\pi^-$ from\n$^{40}$Ca first with full Fermi averaging and then\nutilizing a closure approximation.\nThe closure approximation that we use results from taking\n\\begin{equation}\n\\omega_\\alpha^2= W_{\\pi N}^2-k_\\pi^2\\,\\,\\,{\\rm and}\\,\\,\\,W_{\\pi\nN}=E-\\sqrt{k_\\pi^2+m_{A-1}^2}\n\\label{eq:closure}\n\\end{equation}\nThis is equivalent to setting the momentum of\nthe struck nucleon to zero in the pion--nucleus c.m. frame.\n\nOn the delta resonance, it has been found \\cite{dhole,liurev,erns83}\nthat the closure approximation is totally\ninadequate and even the `optimally-factorized' approximation \\cite{erns83}\nis not quantitative. At these higher energies the\nconclusion is not {\\it \\'a priori} clear. The pion--nucleon amplitude\nis, partial wave by partial wave, rather energy dependent as the individual\npartial waves are resonant. The total amplitude, however, exhibits\nonly two very broad smooth peaks on a large background.\n\nThe resulting differential cross sections for pions at 500 MeV scattering\nfrom $^{40}$Ca are shown in Fig.\\ \\ref{fig2}. We see that the correct\ntreatment of the recoil of the two--body pion--nucleon system only affects\nthe depth of the minimum but not to a very significant degree.\n\nThe behavior of the pion--nucleon scattering amplitude off--shell is taken\nfrom the doorway--resonance model of Ref.\\ \\cite{erns90}.\nAmplitudes which are resonant are separable in their dependence on the\nrelative momenta, hence maximally nonlocal. This is the second\nsource of nonlocality whose importance we want to investigate. In the\ndoorway--resonance model, in each angular momentum, spin and isospin channel,\n\\begin{equation}\n\\langle \\kappa'\\,\\vert\\,t_{JLI} (\\omega_\\alpha )\\,\\vert\\, \\kappa\n\\rangle={v(\\kappa')\\over v(\\kappa_o)}\\,\\langle\\kappa_o\n\\,\\vert\\, t_{JLI} (\\omega_\\alpha )\\,\\vert\\,\\kappa_o\\rangle\\,\n{v(\\kappa)\\over v(\\kappa_o)},\n\\end{equation}\nwhere $\\kappa_o$ is the on--shell momentum, i.e. the momentum\ncorresponding to the center--of--momentum energy $\\omega_\\alpha $.\nThis nonlocality, which is also factored in the delta--hole \\cite{dhole}\nmodel, was found quite\nimportant\\cite{erns83} for energies on and below the $\\Delta_{33}$\nresonance. For numerical convenience, we take the form factor to be a\nGaussian, $v(\\kappa )=\\exp (-\\kappa^2\/\\beta^2)$, and we vary $\\beta$ from\n500 MeV to 4 GeV to investigate the importance of this nonlocality. The\ndifferential cross sections for $\\pi^-$--$^{40}$Ca scattering calculated\nwith this variation in the form--factor range are shown in\nFig. \\ref{fig3}. We see that the dependence on the off--shell range is\nweak at this energy.\n\nThere exists some confusion in the literature concerning the\npossible values for $\\beta$. First, if one does not include\ninvariant phase space factors explicitly, they will appear effectively in\nwhat one might wish to call the form factor. This difference is then one of\nsemantics. However, in deciding what might be a reasonable range over which\nto vary the form factor, one should treat the kinematic factors explicitly.\nIn the $P$--wave channels, an artificial and incorrect increase\n(in coordinate space) in the range of the form factor will also result if\nthe pion--nucleon pole term \\cite{erns78} is not included in the model of\nthe pion--nucleon amplitude. For our purposes here, the difference of\ninterest is in examining how the cross section changes in going from the\nphysically motivated doorway--model form factors to the higher momentum\ncutoff. The 4 GeV cutoff produces an approximately zero--range interaction\nin coordinate space. Figure \\ref{fig3} demonstrates that the nonlocalities\ncontained in a finite range interaction do not alter significantly the\npredicted elastic cross sections for these high energy pions.\n\n\\label{subsec:comp}\nWe have found that the nonlocalities in the optical potential at the high\nenergies do not have an important effect. This suggests that the simple\neikonal approach outlined above could prove to be an adequate model\nfor quantitative work at these energies. To investigate this, we\npicture the differential cross sections predicted by each model\nfor $\\pi^\\pm$--$^{12}$C and $\\pi^\\pm$--$^{40}$Ca at 800 MeV\/c in\nFig.\\ \\ref{fig4} and Fig.\\ \\ref{fig5}. The data are from Ref.\\\n\\cite{marl84}. As stated earlier, we construct the target density from the\nsame wave\nfunctions as were used in the momentum--space calculation and use the\nsame on--shell pion--nucleon scattering amplitude. For $^{40}$Ca, both the\nlocation of the minima and the\nmagnitude of the cross section at the forward angles are in good agreement.\nThe agreement is, however, not as quantitative for\n$^{12}$C where the minima are slightly shifted. This difference is\npresumably caused in part by the fact that the kinematics\nof target recoil enter explicitly in the momentum--space formulation\nbut not in the eikonal approach.\nThe eikonal model is thus more quantitative for\nthe heavier nuclei. We are examining the eikonal model to see if it\ncan be further improved for the light nuclei.\n\n\\subsection{Important features in the eikonal approximation}\n\\label{subsec:imeik}\nIn this section we investigate the role of the Coulomb interaction\nand the first Wallace correction, which are both necessary if one wants\nto use the eikonal model to quantitatively approximate the results of the\nsolution of the model--exact theory.\n\nFirst, we find that the Coulomb interaction plays an important role in\ndetermining the depth of the minima in the differential cross section. To\nsee exactly how important Coulomb--nuclear interference is, we\ncompare the calculation of the $\\pi^-$--$^{40}$Ca differential cross\nsection at a series of energies as shown in Fig.\\ \\ref{fig6} with and\nwithout the Coulomb interaction. The Wallace correction is not included in\neither the solid or the dashed cureves. We find that the first minimum\nin the differential cross sections becomes very deep when the Coulomb\ninteraction is included, falling well below $10^{-2}$ mb\/sr in the region\nof 500 to 600 MeV. Without the Coulomb interaction, the deepest minimum is\nmuch shallower and occurs at 780 MeV. The inclusion of the Coulomb\ninteraction in the theory is necessary if a quantitative comparison with the\ndata for $\\pi^-$--$^{40}$Ca is to be made. Coulomb--nuclear interference is\ndestructive for $\\pi^-$ and constructive for $\\pi^+$, which produces a less\ndramatic effect for $\\pi^+$--$^{40}$Ca scattering in this energy region.\n\nWallace has shown in Ref.~\\cite{wallace} that the semi-classical\napproximation can be improved if higher order corrections (known as\n``Wallace'' corrections) are included. We examine in Fig.~\\ref{fig7} the\nimportance of the first Wallace correction, Eq.~\\ref{eq:wallace}, for\n$\\pi^-$--$^{40}$Ca scattering. We see in Fig.~\\ref{fig7}, in the absence of\nthe Coulomb interaction, that the Wallace correction is dramatic on\nresonance, $T_\\pi= 180$ MeV, where it fills in the deep minimum. At the\nhigher energies, the correction interferes constructively until about 700\nMeV filling in the minima somewhat. At 780 MeV, however, the Wallace\ncorrection interferes destructively with $U_o$ and produce\na much deeper minimum.\n\nIn Fig.~\\ref{fig8} we turn on both the Coulomb interaction and the Wallace\ncorrection, we find a fascinating interplay between the two. Notice that at\n400 MeV, for example, there two corrections are out of phase and tend to\ncancel. At 680 MeV, however, they are in phase and interfere destructively\nwith $U_o$ producing the very deep minimum. By 780 MeV they\nare both out of phase with each other but in phase with the lowest order\nterm and thus fill in the minima. Both the Coulomb and the Wallace\ncorrection are necessary if there is to be a deep minima at 680 MeV, and not\nat the neighboring energies. We note that the data of Ref.~\\cite{marl84} was\ntaken at a pion lab momentum of 800 MeV\/c which is very near $T_\\pi= 680$\nMeV. The depth of the first minimum of the differential cross section for\n$\\pi^-$ scattering over the energy range 400 to 800 MeV should prove a very\nsensitive test of the existence, or lack thereof, of higher order\ncorrections in the reaction dynamics.\n\n \\section{COMPARISON TO OTHER WORK}\n\\label{sec:comp}\nOther work on high--energy pion--nucleus elastic scattering, which utilizes\nthe full Glauber theory, can be found in Refs.\\ \\cite{seki,oset}.\nIn particular, a comparison between the Glauber theory and the optical\nmodel on $\\pi^\\pm$--$^{12}$C has been made in Ref.\\ \\cite{seki} and the two\napproaches\nwere found to be in good agreement at the higher energies. We earlier\nnoted that the\noptical potential model there did not include the full Fermi averaging,\nand it was not clear to what extent the agreement found\nwas caused by the approximate treatment of the\nnon--localities in the optical potential. Our work clarifies that.\nWe also note that the Glauber theory they used is considerably more\ncomplicated than our eikonal model. To our knowledge, the Wallace\ncorrections have not been included in any of the numerical tests\nof the eikonal model in this energy region. Hence, the important\ninterplay between the Wallace and Coulomb corrections that we find has not\nbeen\npreviously noted. We also found that the center--of--mass\ncorrections in the target wave functions are important for $^{12}$C but\nnot for $^{40}$Ca, which is in agreement with Ref.\\ \\cite{seki}. Our work\nstrongly\nsupports their conclusion that pion scattering at the higher energies becomes\ntheoretically and calculationally simpler to treat at a quantitative\nlevel.\n\\section{ SUMMARY, CONCLUSIONS, AND FUTURE PROSPECTS}\n\\label{sec:conc}\nWe have compared the scattering from the fully microscopic and\nnonlocal lowest--order optical potential to the scattering from a simple\nlocal potential in the eikonal approximation. We have investigated\nthe importance of several features of the full model.\nWithin the framework of our\n``model--exact'' optical model, these include Fermi\naveraging and the off--shell behavior of the pion--nucleon\nscattering amplitude.\nFor the eikonal theory, we examine the importance of the first Wallace\ncorrection and find that this correction is responsible for\na noticeable improvement of the eikonal theory in comparison to the\n``model--exact'' theory. In all cases, the Coulomb--nuclear interference\nis important and must be included if one is to compare with data.\n\nThis semi--classical eikonal theory put forth here is very simple, even\nsimpler than the Glauber theory utilized by others.\nIt appears to be quantitatively\nvalid at high energy, at least for the first several minima in the\ndifferential elastic cross section. This makes possible a much simpler\nreaction theory than has been needed at energies at and below\nthe $\\Delta_{33}$ resonance.\n\nAt high energies and\/or for heavy nuclei, use of the full\nmomentum--space optical model becomes very time--consuming and sometimes\ncomputationally impossible.\nIn contrast, the eikonal theory is relatively simple and fast on the\ncomputer. It therefore becomes a matter of considerable practical\nimportance to realize that the physics is faithfully reproduced by the\nsimple version of the theory out to the position of the second minimum\nin the differential cross section.\n\nAn examination of the pion--nucleon cross section shows the\ncomplex interplay\nof resonances. Ultimately, it is interesting to explore how resonances behave\nin the nucleus and how the specific features of the optical potential\nare suited for nuclear structure studies. In this pursuit, it will be very\nhelpful to\nhave a simple version of the theory as developed here. We will pursue these\nquestions in subsequent work.\n\nThe momentum space optical model has been extended to treat kaon-nucleus\nscattering \\cite{chen92}. Results show qualitative agreement with previous\nwork \\cite{sieg84}. Modifications to the kaon-nucleon amplitudes\nin the\nnuclear medium are needed to eliminate discrepancies between the theory and\nthe data. On the quantitative level, results from the momentum space\napproach give larger discrepancies between theory and data than do those\nfrom the coordinate space\ncalculations used in Ref.\\ \\cite{sieg84}. Our eikonal model, which is also a\ncoordinate space approach and been shown to be a good approximation\nto the ``model-exact'' momentum space theory for high-energy pion\nscattering, can also be modified to treat the kaon scattering. Results from\nthe eikonal model may reveal the sources of the discrepancies between the\nmomentum space and the previous coordinate space calculations.\nWork on the kaon-nucleus scattering utilizing the eikonal model is in\nprogress.\n\\newpage\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\n\\label{introd}\n\nThe phase diagram of strongly interacting matter in the temperature ($T$) - \nbaryon density plane remains a challenge for theoretical physics. Although there\nis little doubt that it features a low-temperature hadronic phase, with \nbroken chiral symmetry, and a high-temperature deconfined phase,\nwith restored chiral symmetry, the question about the precise location\nand the exact nature of the transition between these two phases is still open. \nYet, the answer to this question has many phenomenological implications:\nthe region of the phase diagram with high $T$ and small baryon density is\nrelevant for the physics of the early Universe, whereas the region of \nlow $T$ and high baryon density is interesting for the astrophysics of \nsome compact objects, but other corners of the phase diagram are not less \ninteresting (see Ref.~\\cite{Stephanov:2007fk} for an overview). \n\nRelativistic heavy-ion collisions provide us with a unique opportunity\nto infer properties of the transition: depending on the energy of the ion beams\nand on the mass number of the colliding ions, the fireball generated in the \ncollision could fulfill the temperature and baryon density conditions under\nwhich the deconfined phase appears as a transient state, before the\nsystem freezes out into hadrons, which are then detected. Thermal-statistical \nmodels, assuming approximate chemical equilibrium at the chemical freeze-out\npoint, are able to describe the particle yields at a given collision energy\nin terms of two parameters only, the freeze-out temperature $T$ and the baryon \nchemical potential $\\mu_B$. The collection of freeze-out parameters extracted\nfrom experiments with different collision energy lie on a curve in \nthe $(T,\\mu_B)$-plane, extending up to $\\mu_B\\lesssim$ 800 MeV (see Fig.~1 of \nRef.~\\cite{Cleymans:2005xv}, or Ref.~\\cite{Becattini:2012xb} for a recent \nre-analysis of experimental data). \n\nQuantum ChromoDynamics (QCD) is widely accepted as the theory of strong \ninteractions and, as such, must encode all the information needed to\nprecisely draw the phase diagram in the $(T,\\mu_B)$-plane. As a matter of\nfact, only some corners of it can be accessed by first-principle applications\nof QCD, in the perturbative or in the nonperturbative regime. Here we \nfocus on the lattice approach of QCD, based on the idea of discretizing\nthe theory on a Euclidean space-time lattice and simulating it by Monte Carlo \nnumerical simulations as a statistical system, with Boltzmann weight \ngiven by $\\exp(-S_E)$, where $S_E$ is the QCD Euclidean action. The region\nof the phase diagram where $\\mu_B\/(3 T)\\lesssim 1$ is within the reach of this \napproach and one can therefore address, at least inside this region, the \nproblem of determining the shape taken by the QCD pseudocritical line \nseparating the hadronic from the deconfined phase.\n\nThere is no {\\it a priori} argument for the coincidence of the QCD \npseudocritical line with the chemical freeze-out curve: if the deconfined \nphase is realized in the fireball, in cooling down the system first \nre-hadronizes, then reaches the chemical freeze-out. This implies that the \nfreeze-out curve lies below the pseudocritical line in the $\\mu_B$-$T$ plane. \nIt is a common working hypothesis that the delay between chemical freeze-out and \nrehadronization is so short that the two curves lie close to each other and \ncan therefore be compared.\nUnder the assumptions of charge-conjugation invariance at $\\mu_B=0$ and \nanalyticity around this point, the QCD pseudocritical line, as \nwell as the freeze-out curve, can be parameterized, at low baryon densities, \nby a lowest-order expansion in the dimensionless quantity $\\mu_B\/T(\\mu_B)$, as\n\\begin{equation}\n\\frac{T(\\mu_B)}{T_c(0)}=1-\\kappa \\left(\\frac{\\mu_B}{T(\\mu_B)}\\right)^2 \\, + \\, \\ldots \\;,\n\\label{curv}\n\\end{equation}\nwhere $T_c(0)$ and $\\kappa$ are, respectively, the pseudocritical temperature \nand the curvature at vanishing baryon density.\n\nDirect Monte Carlo simulations of lattice QCD at nonzero baryon density\nare hindered by the well known ``sign problem'': $S_E$ becomes complex and \nthe Boltzmann weight loses its sense. Several ways out of this problem\nhave been devised (see Ref.~\\cite{Philipsen:2005mj,*Schmidt:2006us,*deForcrand:2010ys,*Aarts:2013bla} for a review): \nredesigning the Monte Carlo updating algorithms for a complex action~\\cite{Karsch:1985cb,*Aarts:2009uq,*Aarts:2011ax,*Aarts:2011zn,*Seiler:2012wz,*Aarts:2013uxa,*Sexty:2013ica}, \nreweighting from the ensemble at $\\mu_B=0$~\\cite{Barbour:1997bh,*Fodor:2001au,*Fodor:2001pe,*Fodor:2004nz}, \nTaylor expanding the relevant observables around $\\mu_B=0$ and calculating the \nfirst coefficients of the series by simulations at $\\mu_B=0$~\\cite{Gottlieb:1988cq,*Choe:2001cq,*Choe:2002mt,*Allton:2002zi,Allton:2003vx,*Allton:2005gk,*Ejiri:2005uv,Endrodi:2011gv,*Borsanyi:2012cr}, \nusing the canonical formulation~\\cite{Alford:1998sd,*Hasenfratz:1991ax,*Kratochvila:2005mk,*Alexandru:2005ix,deForcrand:2006ec},\nusing the density of states method~\\cite{Bhanot:1986kv,*Karliner:1987cu,*Azcoiti:1990ng,*Ambjorn:2002pz}\nand simulating the theory at imaginary chemical potentials and performing\nthe analytic continuation to real ones~\\cite{Lombardo:1999cz,*deForcrand:2002ci,*deForcrand:2003hx,*D'Elia:2002gd,*D'Elia:2004at,*Azcoiti:2005tv,*deForcrand:2006pv,*deForcrand:2008vr,*deForcrand:2007rq,Cea:2010md,Cea:2010fh,*Cea:2012vi,Cea:2012ev,*Wu:2006su,*Nagata:2011yf,*Giudice:2004se,*Giudice:2004pe,*Papa:2006jv,*Cea:2007wa,*Cea:2007wa,*Cea:2010bp,*Cea:2009ba,*Cea:2010bz,*Karbstein:2006er,Cea:2014xva,Bonati:2014rfa,Bonati:2015bha,Bellwied:2015rza}.\n\nThe numerical evidence gathered so far in QCD with $n_f=2+1$ and physical \nor almost physical quark masses points to a scenario with a smooth\ncrossover between the hadronic and the deconfined (or chirally symmetric)\nphase at $\\mu_B=0$, with a pseudocritical temperature $T_c(0)$ of about \n155~MeV~\\cite{Aoki:2006br,*Aoki:2006we,*Aoki:2009sc,Bazavov:2011nk,*Bhattacharya:2014ara}. This crossover behavior should persist in some neighborhood of\n$\\mu_B=0$, up to the onset of a first-order transition at some value of \n$\\mu_B>0$. \n\nThe state-of-the-art of lattice determinations of the curvature $\\kappa$,\nup to the very recent papers of Ref.~\\cite{Bonati:2015bha,Bellwied:2015rza}, \nis summarized in Fig.~10 of Ref.~\\cite{Bonati:2014rfa}: depending on the\nlattice setup and on the observable used to probe the transition, the value \nof $\\kappa$ can change even by almost a factor of three. The lattice \nsetup dependence stems from the kind of adopted discretization, the lattice \nsize, the choice of quark masses and chemical potentials, the procedure\nto circumvent the sign problem. This dependence would totally disappear if, \nideally, all groups would use the same lattice setup. A contribution to the\nunderstanding of the impact of the lattice setup dependence is provided \nin the Appendix~B of Ref.~\\cite{Bonati:2014rfa}. The dependence on the\nprobe observable is, instead, irreducible: since a smooth crossover is taking\nplace rather than a true phase transition, one cannot define a {\\it bona fide} \norder parameter whose behavior would permit to uniquely locate the transition\npoint; instead, for any adopted surrogate observable, a different transition\npoint should be expected, at least in principle.\n\nOn the side of the determinations of the freeze-out curve, two recent \ndeterminations~\\cite{Cleymans:2005xv,Becattini:2012xb} of $\\kappa$, both\nbased on the thermal-statistical model, but the latter of them including the\neffect of inelastic collisions after freeze-out, give two quite different\nvalues of $\\kappa$, each seeming to prefer a different subset of lattice\nresults (see Fig.~3 of Ref.~\\cite{Cea:2014xva} for a snapshot of the\nsituation). \n\nThe aim of this work is to contribute to a better understanding of the \nsystematics underlying lattice determinations of the curvature $\\kappa$, by \ncorroborating our previous determination~\\cite{Cea:2014xva} with \nan extrapolation to the continuum limit and by comparing it with\nexperimental analyses of the freeze-out curve.\n\nOur lattice setup is as follows. We simulate the HISQ\/tree action of the \nMILC collaboration with 2+1 staggered fermions on lattices with temporal\nextension $L_t=6$, 8, 10 and 12 and aspect ratio equal to four. We work on \nthe line of constant physics (LCP) as determined in Ref.~\\cite{Bazavov:2011nk},\nwith the strange mass set at the physical value and the light quark mass fixed \nat $m_l=m_s\/20$. \nAs discussed in Ref.~\\cite{Bazavov:2011nk}, this amounts to tune the strange quark mass \nuntil the mass of the fictitious $\\eta_{s\\bar{s}}$ meson matches the lowest order perturbation theory estimate\n$m_{\\eta_{s\\bar{s}}} = \\sqrt{2 m_K^2 - m_\\pi^2}$. Consequently within our simulations the pion mass is \n$m_\\pi \\simeq 160 \\, \\text{MeV}$.\n\nWe perform simulations at imaginary quark chemical potentials, \nassigning the same value to the three quark species, $\\mu_l=\\mu_s\\equiv \\mu$, \nthen extrapolate to real chemical potentials. Our probe observables are the \ndisconnected susceptibility of the light quark chiral condensate and its \nrenormalized counterpart.\nSimulating the theory at imaginary chemical potentials poses no restriction\non the lattice size or in the choice of the couplings. However, the periodicity \nin ${\\rm Im}(\\mu_l)\/T$ of the partition function~\\cite{Roberge:1986mm} implies \nthat the information gathered outside a narrow interval of imaginary chemical\npotentials is redundant. For the setup with $\\mu_l=\\mu_s\\equiv \\mu$, this \ninterval can be chosen as the region $0\\leq {\\rm Im}(\\mu)\/T\\leq \\pi\/3$. A safe\nextrapolation of the critical line to real chemical potentials requires that\nit exhibits a smooth dependence on imaginary chemical potential over this\ninterval, a condition which must be checked to be satisfied {\\it a posteriori} \nby our data. \n\nThe preference to the disconnected susceptibility of the light quark chiral \ncondensate has multiple motivations~\\cite{Bazavov:2011nk}. First of all, for \nsmall enough quark masses, its contribution to the chiral susceptibility \ndominates over the connected one, which is harder to compute; then, it shows a \nstrong sensitivity to the transition; finally, it is exempt from additive \nrenormalization, undergoing only a multiplicative one. This translate into\na very precise determination of the critical couplings at imaginary $\\mu$,\nwhich is the main prerequisite of a safe extrapolation to the real values of\n$\\mu$.\n\nThere are two main limitations in our setup. The first is that we work\nwith a physical strange quark mass, but with light quarks a bit heavier\nthan physical ones. Numerical results in $n_f=2$ indicate a mild dependence \nof the curvature on the quark mass (see the discussion in Sec.~III of \nRef.~\\cite{Cea:2012ev}). If the same applies here, as we believe, our result\nfor $\\kappa$ will only slightly underestimate (in absolute value) the true \nphysical curvature. The second limitation is that, for the sake of comparison\nwith the freeze-out curve, our setup of chemical potentials could not be the\none which better reproduces the initial conditions of heavy ion collision.\nIn fact, strangeness neutrality would rather impose $\\mu_s \\alt \\mu_l$.\nIn general, the setup $\\mu_l=\\mu_s= \\mu_B\/3$ approximates strangeness neutrality at low temperatures, while\nthe $\\mu_s= 0$ setup is relevant for high enough temperatures. \\\\\n It is natural to expect that the effect of taking $\\mu_s=\\mu_l$ instead of $\\mu_s=0$\nbecomes less and less evident when $\\mu_l\/T$ approaches zero, so that the \ncurvature $\\kappa$ at zero baryon density should not differ too much in the\ntwo cases. The numerical analysis of Refs.~\\cite{ Bonati:2014rfa,Bonati:2015bha,Bellwied:2015rza} has shown \nthat this effect is invisible within the accuracy of the lattice setup\nadopted there.\n\nThe paper is organized as follows: in Sec.~II we give some further details\nof our numerical simulations. In Sec.~III we show our numerical results\nfor $\\kappa$. Finally, in Sec.IV we draw our conclusions.\n\n\\section{Simulation details and numerical results}\n\nWe perform simulations of lattice QCD with 2+1 flavors of rooted staggered \nquarks at imaginary quark chemical potential.\nWe have made use of the HISQ\/tree action~\\cite{Follana:2006rc,Bazavov:2009bb,Bazavov:2010ru} \nas implemented in the publicly available MILC code~\\cite{MILC}, \nwhich has been suitably modified by us in order to introduce an imaginary quark\nchemical potential $\\mu = \\mu_B\/3$. \nThat has been done by multiplying all forward and backward \ntemporal links entering the discretized Dirac operator \nby $\\exp (i a \\mu )$ and $\\exp (- i a \\mu )$, respectively: in this way,\nthe fermion determinant is still real and positive, so that standard\nMonte Carlo methods can be applied.\nAs already remarked above, in the present study we have \n$\\mu=\\mu_l=\\mu_s$. This means that the Euclidean partition function of the\ndiscretized theory reads\n\\begin{equation}\n\\label{Z}\nZ=\\int [DU] e^{-S_{\\rm gauge}} \\prod_{q=u,d,s} {\\rm det}(D_q[U,\\mu])^{1\/4} \\;,\n\\end{equation}\nwhere $S_{\\rm gauge}$ is the Symanzik-improved gauge \naction and $D_q[U,\\mu]$ is the staggered Dirac operator, modified as \nexplained above for the inclusion of the imaginary quark chemical\n(see Ref.~\\cite{Bazavov:2009bb} and appendix~A of Ref.~\\cite{Bazavov:2010ru} for the precise definition of\nthe gauge action and the covariant derivative for highly improved staggered fermions). \n\nAll simulations make use of the rational hybrid Monte Carlo (RHMC) \nalgorithm. The length of each RHMC trajectory has been set to \n$1.0$ in molecular dynamics time units.\n\nWe have simulated the theory at finite temperature, and for several values of \nthe imaginary quark chemical potential, near the transition temperature, \nadopting lattices of size $16^3\\times 6$, $24^3\\times 6$, $32^3 \\times 8$, \n$40^3 \\times 10$ and $48^3 \\times 12$. \nWe have discarded typically not less than one thousand trajectories for each \nrun and have collected from {4k to 8k} trajectories for measurements.\n\nThe pseudocritical point $\\beta_c(\\mu^2)$ has been determined as the value \nfor which the renormalized disconnected susceptibility of the light quark chiral condensate\ndivided by $T^2$ exhibits a peak. \\\\\nThe bare disconnected susceptibility is given by:\n\\begin{equation}\n\\label{chi_dis}\n\\chi_{l, \\rm disc} =\n{{n_f^2} \\over 16 L_s^3 L_t}\\left\\{\n\\langle\\bigl( {\\rm Tr} D_q^{-1}\\bigr)^2 \\rangle -\n\\langle {\\rm Tr} D_q^{-1}\\rangle^2 \\right\\}\\;,\n\\end{equation}\n Here $n_f=2$ is the number of light flavors\nand $L_s$ denotes the lattice size in the space direction. The renormalized chiral susceptibility\nis defined as: \n\\begin{equation}\n\\label{chi_ren}\n\\chi_{l, \\rm ren} = \\frac{1}{Z_m^2} \\; \\chi_{l, \\rm disc} .\n\\end{equation}\nThe multiplicative renormalization factor $Z_m$ can be deduced\nfrom an analysis of the line of constant physics for the light quark masses.\nMore precisely, we have~\\cite{Bazavov:2010ru}:\n\\begin{equation}\n\\label{zeta}\nZ_m(\\beta) \\; = \\; \\frac{m_l(\\beta)}{m_l(\\beta^*)} \\; ,\n\\end{equation}\nwhere the renormalization point $\\beta^*$ is chosen such that:\n\\begin{equation}\n\\label{beta*}\n\\frac{r_1}{a(\\beta^*)} \\; = \\; 2.37 \\; ,\n\\end{equation}\nwhere the function $a(\\beta)$ is discussed below.\nIn Fig.~\\ref{fig_zetam} is shown the multiplicative renormalization \nfactor $Z_m$ determined in the case when $r_1$ is used to set the\nscale (see below).\n\\begin{figure}[tb]\n\\includegraphics*[width=1.\\columnwidth]\n{.\/zetam_plot.eps}\n\\caption{The multiplicative renormalization factor $Z_m$ in the case of \n$r_1$-scale. The renormalization point is $\\beta^*=6.54706$.}\n\\label{fig_zetam}\n\\end{figure}\nTo precisely localize the peak in $\\chi_{l, \\rm ren}\/T^2$, a Lorentzian fit has \nbeen used. For illustrative purposes, \nin Fig.~\\ref{fig_chiral_light_suscep} we display our determination of the \npseudocritical couplings at $\\mu\/(\\pi T)=0.2i$ for all lattices\nconsidered in this work. The complete collection of results for\nthe disconnected susceptibility of the light quark chiral condensate obtained\nin this work is presented in the Appendix.\n\n\\begin{figure}[tb]\n\\includegraphics*[width=1.\\columnwidth]\n{.\/Figure_1.eps}\n\\caption{The real part of the renormalized susceptibility of the light quark \nchiral condensate over $T^2$ on the lattices $16^3\\times 6$,\n$32^3\\times 8$, $40^3\\times 10$ and $48^3\\times 12$ at $\\mu\/(\\pi T)=0.2i$. \nFull lines give the Lorentzian fits near the peaks. The temperature\nhas been determined from the $r_1$ scale.}\n\\label{fig_chiral_light_suscep}\n\\end{figure}\n\nTo get the ratios $T_c(\\mu)\/T_c(0)$, we fix the lattice spacing through the \nobservables $r_1$ and $f_K$, following the discussion in the Appendix~B of \nRef.~\\cite{Bazavov:2011nk}. \\\\\n For the $r_1$ scale the lattice spacing\nis given in terms of the $r_1$ parameter as:\n\\begin{equation}\n\\label{scale-r1}\n\\frac{a}{r_1}(\\beta)_{m_l=0.05m_s}=\n\\frac{c_0 f(\\beta)+c_2 (10\/\\beta) f^3(\\beta)}{\n1+d_2 (10\/\\beta) f^2(\\beta)} \\; ,\n\\end{equation}\nwith $c_0=44.06$, $c_2=272102$, $d_2=4281$, $r_1=0.3106(20)\\ {\\text{fm}}$. \\\\\nOn the other hand, in the case of the $f_K$ scale we have:\n \\begin{equation}\n\\label{scale-fk}\n a f_K(\\beta)_{m_l=0.05m_s}=\n\\frac{c_0^K f(\\beta)+c_2^K (10\/\\beta) f^3(\\beta)}{\n1+d_2^K (10\/\\beta) f^2(\\beta)} \\; ,\n\\end{equation}\nwith $c_0^K=7.66$, $c_2^K=32911$, $d_2^K=2388$, $r_1f_K \\simeq 0.1738$. \nIn Eqs.~(\\ref{scale-r1}) and (\\ref{scale-fk}), $f(\\beta)$ is the two-loop beta \nfunction,\n\\begin{equation}\n\\label{beta function}\nf(\\beta)=(b_0 (10\/\\beta))^{-b_1\/(2 b_0^2)} \\exp(-\\beta\/(20 b_0))\\;,\n\\end{equation}\n$b_0$ and $b_1$ being its universal coefficients. \\\\\n\nOur results are summarized in Table~\\ref{summary}.\nFor all lattice sizes but $24^3\\times6$ (where we have only one value of $\\mu$),\nthe behavior of $T_c(\\mu)\/T_c(0)$ can be nicely fitted with a linear function \nin $\\mu^2$,\n\\begin{equation}\n\\label{linearfit}\n\\frac{T_c(\\mu)}{T_c(0)} = 1 + R_q \\left(\\frac{i \\mu}{\\pi T_c(\\mu)}\\right)^2 \\;,\n\\end{equation}\nwhich gives us access to the curvature $R_q$ and, hence, to the curvature\nparameter $\\kappa=-R_q\/(9\\pi^2)$ introduced in Eq.~(\\ref{curv}). \nOn the $24^3\\times6$ lattice the linearity in $\\mu^2$ has been assumed to hold, \nin order to extract $R_q$ from the only available determination at \n$\\mu\/(\\pi T)=0.2i$.\n\n\\begin{table}[tb]\n\\setlength{\\tabcolsep}{0.9pc}\n\\centering\n\\caption[]{Summary of the values of the ratio $T_c(\\mu)\/T_c(0)$ for \nthe imaginary quark chemical potentials $\\mu$ considered in this work.\nThe data for $\\mu=0$ on the $24^3\\times6$, $32^3\\times68$ and $48^3\\times12$ \nlattices have been estimated from the disconnected chiral susceptibilities \nreported respectively on Tables~X, XI and XII of Ref.~\\cite{Bazavov:2011nk}.\nThe datum for $\\mu=0$ on the $40^3\\times10$ lattice has been estimated from \nthe disconnected chiral susceptibilities reported on Table~XI of \nRef.~\\cite{Bazavov:2014pvz}.\nThe values of $T_c(\\mu)\/T_c(0)$ evaluated fixing the lattice scale by\n$r_1$ and $f_K$ are reported, respectively, in the third and in the \nfourth column of the table.}\n\\begin{tabular}{cllll}\n\\hline\n\\hline\nlattice & $\\mu\/(\\pi T)$ & $T_c(\\mu)\/T_c(0)$ & $T_c(\\mu)\/T_c(0)$ \\\\\n & & \\ ($r_1$ scale) & \\ ($f_K$ scale) \\\\\n\\hline\n16$^3\\times 6$ & 0.15$i$ & 1.038(13) & 1.043(14) \\\\\n & 0.2$i$ & 1.063(15) & 1.070(15) \\\\\n & 0.25$i$ & 1.085(16) & 1.095(18) \\\\\n\\hline\n24$^3\\times 6$ & 0.2$i$ & 1.061(9) & 1.067(10) \\\\\n\\hline\n32$^3\\times 8$ & 0.15$i$ & 1.054(7) & 1.059(8) \\\\\n & 0.2$i$ & 1.066(10) & 1.071(11) \\\\\n & 0.25$i$ & 1.117(10) & 1.126(10) \\\\\n\\hline\n40$^3\\times 10$ & 0.15$i$ & 1.023(23) & 1.024(24) \\\\\n & 0.2$i$ & 1.075(14) & 1.079(15) \\\\\n & 0.25$i$ & 1.102(15) & 1.107(15) \\\\\n\\hline\n48$^3\\times 12$ & 0.15$i$ & 1.013(31) & 1.013(33) \\\\\n & 0.20$i$ & 1.051(14) & 1.052(15) \\\\\n & 0.25$i$ & 1.094(26) & 1.097(25) \\\\\n\\hline\n\\hline\n\\end{tabular}\n\\label{summary}\n\\end{table}\n\nFor the sake of the extrapolation to the continuum limit, in \nFig.~\\ref{curvature_vs_Nt} we report our determinations of $R_q$ on the \nlattices $24^3\\times 6$, $32^3\\times 8$, $40^3\\times 10$, $48^3\\times 12$, \nand from the two different methods to set the scale, {\\it versus} $1\/L_t^2$.\n\n\nWithin our accuracy, cutoff effects on $R_q$ are negligible, so that a \nconstant fit works well over the whole region ($\\chi^2_r \\simeq 0.99$), thus \nincluding also the smallest $24^3\\times6$ lattice. Taking into account\nthe uncertainties due to the continuum limit extrapolation, \n\\begin{equation}\n\\label{curvature}\n\\kappa = 0.020(4) \\;.\n\\end{equation}\nOur estimate of the uncertainties for the curvature given in Eq.~(\\ref{curvature}) takes into account both the error in the fit minimization\nand the choice of the minimization function.\nWe stress, however, that if we exclude from the fit the value on the lattice\nwith the smallest $L_t$, {\\it i.e.} the rightmost points in \nFig.~\\ref{curvature_vs_Nt}, the extrapolation to the continuum becomes largely\nundetermined. Indeed, with the values of $R_q$ obtained in the present work\n(see Table~\\ref{Rq}), the fit with a constant is rather stable even if\n$L_t=6$ is excluded, but the fit with a linear function in $1\/L_t^2$ in the \nlatter case gives a much smaller value of the curvature $\\kappa$, though\nwith a large uncertainty (see Table~\\ref{fit}).\n\\begin{table*}[tb]\n\\setlength{\\tabcolsep}{0.9pc}\n\\centering\n\\caption[]{Summary of determinations of the curvature $R_q$ for all values of \n$L_t$ considered in this work, and from the two different methods to set the \nscale.}\n\\begin{tabular}{lcccc}\n\\hline\n\\hline\n$L_t$ & $6$ & $8$ & $10$ & $12$\\\\\n\\hline\n$R_q$ ($r_1$ scale) & $-1.466(306)$ & $-1.902(192)$ & $-1.685(294)$ \n& $-1.337(410)$ \\\\\n$\\kappa$ ($r_1$ scale) & $0.017(3)$ & $0.021(2) $ & $0.019(3)$ & $0.015(5)$ \\\\\n\\hline\n$R_q$ ($f_K$ scale) & $-1.646(336)$ & $-2.041(206)$ & $-1.769(309)$ \n& $-1.394(415)$ \\\\\n$\\kappa$ ($r_K$ scale) & $0.019(4)$ & $0.023(2)$ & $0.020(3)$ & $0.016(5)$ \\\\\n\\hline\n\\hline\n\\end{tabular}\n\\label{Rq}\n\\end{table*}\n\\begin{table*}[tb]\n\\setlength{\\tabcolsep}{1.5pc}\n\\centering\n\\caption[]{Summary of the fit of the curvature $\\kappa$ with the \nfunction $\\kappa(L_t) = \\kappa + A\/L_t^2$, with $A$ taken equal to zero or left\nfree. The first column specifies the values of $L_t$ included in the fit,\nthe last column the reduced $\\chi^2$. The uncertainties on the fit parameters\nare obtained with 70\\% confidence level.}\n\\begin{tabular}{lllc}\n\\hline\n\\hline\n$L_t$ included & \\hspace{0.6cm} $\\kappa$ & \\hspace{0.6cm} $A$ & $\\chi^2_{\\rm r}$ \n\\\\\n\\hline\n6, 8, 10, 12 & 0.01991(114) & 0. & 0.76 \\\\ \n6, 8, 10, 12 & 0.02014(449) & \\hspace{-0.35cm} $-$0.015(259) & 0.89 \\\\ \n\\hline\n8, 10, 12 & 0.02048(127) & 0. & 0.80 \\\\ \n8, 10, 12 & 0.01182(748) & 0.669(559) & 0.14 \\\\ \n\\hline\n\\hline\n\\end{tabular}\n\\label{fit}\n\\end{table*}\n\\begin{figure}[tb]\n\\includegraphics*[width=1.\\columnwidth]\n{.\/Figure_2.eps}\n\\caption{Determinations of the curvature $R_q$ on the lattices $24^3\\times 6$, \n$32^3 \\times 8$, $40^3 \\times 10$, $48^3 \\times 12$, and from the two different \nmethods to set the scale, {\\it versus} $1\/L_t^2$. Data points related with \nthe $f_K$ scale setting have been slightly shifted along the horizontal axis \nfor better readability. The dashed horizontal line gives the result of \nthe fit to all data with a constant; the solid horizontal lines indicate the \nuncertainty on this constant (95\\% confidence level).}\n\n\\label{curvature_vs_Nt}\n\\end{figure}\n\n\\section{Conclusions and discussion}\n\nWe have studied QCD with $n_f = 2+1$ flavors discretized in the HISQ\/tree \nrooted staggered fermion formulation and in the presence of an imaginary \nbaryon chemical potential, with a physical strange quark mass and a \nlight-to-strange mass ratio $m_l\/m_s = 1\/20$, and $\\mu=\\mu_l=\\mu_s$.\n\nWe have estimated, by the method of analytic continuation, the continuum limit\nof the curvature of the pseudocritical line in the temperature - baryon \nchemical potential, defined in Eq.~(\\ref{curv}). The observable adopted\nto identify, for each fixed $\\mu$, the crossover temperature has been\nthe disconnected part of the renormalized susceptibility of the light quark \nchiral condensate, in units of the squared temperature. This observable is\nconvenient for many reasons: it dominates, for small enough quark masses, the\nwhole light chiral susceptibility, which would be much harder to implement;\nit undergoes only a multiplicative renormalization; it is strongly sensitive \nto the transition, thus allowing precise determinations of the pseudocritical \ntemperatures.\n\nWe have found that, within the accuracy of our determinations, cutoff effects\non the curvature are negligible already on the lattice with temporal size\n$L_t=6$. Our determination of the curvature parameter, $\\kappa$=0.020(4),\nis indeed compatible with the value quoted in our previous \npaper~\\cite{Cea:2014xva}, $\\kappa$=0.018(4), without the extrapolation to\nthe continuum.\n\n\\begin{figure}[tb]\n\\includegraphics*[height=0.295\\textheight,width=1.\\columnwidth]\n{.\/Figure_3.eps}\n\\caption{$T_c(\\mu_B)$ versus $\\mu_B$ (units in GeV). \nExperimental values of $T_c(\\mu_B)$ are taken\nfrom Fig.~1 of Ref.~\\cite{Cleymans:2005xv} (black circles) and from\nFig.~3 of Ref.~\\cite{Alba:2014eba} (green triangles), for the \nstandard hadronization model and for the susceptibilities of conserved charges respectively. \nThe dashed line is a parametrization corresponding to\n$T_c(\\mu_B) = T_c(0) - b \\mu_B^2$ with $T_c(0)=0.154(9)\\, \\text{GeV}$ and \n$b=0.128(25)\\,{\\text{GeV}}^{-1}$ . \nThe solid lines represent the corresponding error band.}\n\\label{Tcmu}\n\\end{figure}\n\nIt is interesting to extrapolate the critical line as determined in this work \nto the region of real baryon density and compare it with the freeze-out\ncurves resulting from a few phenomenological analyses of relativistic\nheavy-ion collisions. This is done in Fig.~\\ref{Tcmu}, where we report\ntwo different estimates. The first is from the analysis of \nRef.~\\cite{Cleymans:2005xv}, based on the standard statistical hadronization \nmodel, where the freeze-out curve is parametrized as \n\\begin{equation}\n\\label{cleymans}\nT_c(\\mu_B) = a - b \\mu_B^2 - c \\mu_B^4\\;,\n\\end{equation}\nwith $a=0.166(2) \\ {\\text{GeV}}$, $b=0.139(16) \\ {\\text{GeV}}^{-1}$, \nand $c=0.053(21) \\ {\\text{GeV}}^{-3}$.\nThe second estimate is from Ref.~\\cite{Alba:2014eba} and is based on the analysis\nof susceptibilities of the (conserved) baryon and electric charges.\nIn fact, our critical line is in nice agreement with all the freeze-out points of\nRefs.~\\cite{Cleymans:2005xv,Alba:2014eba}. In particular, using our estimate\nof the curvature, Eq.~(\\ref{curvature}), we get \n$b=0.128(25) \\ {\\text{GeV}}^{-1}$, in very good agreement with the\nquoted phenomenological value. The significance of the comparison\npresented in Fig.~\\ref{Tcmu}, with special reference to the question\nwhether the pseudocritical line lies indeed above the freeze-out curve, can be \nincreased at the (nonnegligible) price of reducing the uncertainties on \n$T_c(0)$ and on $\\kappa$. \\\\\n\nSome {\\it caveats} are in order here. We do not expect our critical line \nto be reliable too far from $\\mu=0$: as a rule of thumb, we can trust it up \nto real quark chemical potentials of the same order of the modulus\nof the largest imaginary chemical potential included in the \nfit~(\\ref{linearfit}), {\\it i.e.} $|\\mu|\/(\\pi T)=0.25$. This translates\nto real baryon chemical potentials in the region $\\mu_B \\lesssim 0.4\\,\\text{GeV}$.\nMoreover, the effect of taking $\\mu_s=\\mu_l$ instead of $\\mu_s < \\mu_l$\nshould become visible on the shape of the critical line as we move away\nfrom $\\mu=0$ in the region of real baryon densities, thus reducing\nfurther the region of reliability of our critical line. So, from a prudential\npoint of view, the agreement shown in Fig.~\\ref{Tcmu} could be considered\nthe fortunate combination of different kinds of systematic effects. We cannot\nhowever exclude the possibility that the message from Fig.~\\ref{Tcmu} is to be \ninterpreted in positive sense, {\\it i.e.} the setup we adopted and the\nobservable we considered may catch better some features of the crossover\ntransition, thus explaining the nice comparison with freeze-out data.\nIndeed, our result for the continuum extrapolation of the curvature $\\kappa$ is\nin fair agreements with the recent estimates in Ref.~\\cite{Bonati:2015bha}, where\nboth setup $\\mu_s=\\mu_l$ and $\\mu_s=0$ were adopted,\nand Ref.~\\cite{ Bellwied:2015rza}, where the strangeness neutral trajectories\nwere determined from lattice simulations by imposing $\\langle n_S \\rangle=0$.\n\n\\section*{Acknowledgments}\nThis work was in part based on the MILC collaboration's public lattice gauge \ntheory code. See {\\url{http:\/\/physics.utah.edu\/~detar\/milc.html}}.\nThis work has been partially supported by the INFN SUMA project.\nSimulations have been performed on BlueGene\/Q at CINECA \n(Projects Iscra-B\/EXQCD and INF14\\_npqcd), on the BC$^2$S cluster in Bari and on the CSNIV \nZefiro cluster in Pisa.\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}} +{"text":"\\section{Introduction}\nLet $I\\subset k[x_0,\\dots,x_N]$ be a homogeneous ideal. For $r\\geq 0$, the $r$-th symbolic power of $I$ is defined to be \\[I^{(r)}=\\bigcap_{p\\in\\mbox{Ass}(R\/I)}(I^rR_p\\cap R).\\] Symbolic powers of ideals are interesting for a number of reasons, not least of which is that, for a radical ideal $I$,the $r$-th symbolic power $I^{(r)}$ is the ideal of all polynomials vanishing to order at least $r$ on $V(I)$ (by the Zariski-Nagata theorem).\n\nContainment relationships between symbolic and ordinary powers are a source of great interest. As an immediate consequence of the definition, $I^r\\subseteq I^{(r)}$ for all $r$. However, the other type of containment, namely that of a symbolic power in an ordinary power is much harder to pin down. It has been proved by Ein-Lazarsfeld-Smith \\cite{ELS} and Hochster-Huneke \\cite{HH} that $I^{(m)}\\subseteq I^r$ for all $m\\geq Nr$, but as of yet there are no examples in which this bound is sharp.\\\\\n\nIt was conjectured by Harbourne in \\cite[Conjecture 8..4.3]{PSC} (and later in \\cite[Conjecture 4.1.1]{HaHu} in the case $e=N-1$) that $I^{(m)}\\subseteq I^r$ for all $m\\geq er-(e-1)$, where $e$ is the codimension of $V(I)$. While this conjecture holds in a number of important cases, some counterexamples have also been found. Notably, the main counterexamples come from singular points of hyperplane arrangements \\cite{BNAL}. One particular family is known in the literature under the name of Fermat configurations of points cf. $\\mathbb P^2$ \\cite{BNAL, MS}. These have been recently generalized to Fermat-like configurations of lines in $\\mathbb P^3$ in \\cite{MS}. The $\\rm{Ceva}(n)$ arrangement of hyperplanes in $\\mathbb{P}^N$ is defined by the linear factors of \\[F_{N,n}=\\prod_{0\\leq i3$, assume that $F_{N-1,n}\\not\\in I_{N-1,n}^2$ and consider the evaluation homomorphism $\\pi:S_N\\rightarrow S_{N-1}$ defined by $\\pi(x_N)=1$ and $\\pi(x_i)=x_i$ for $i \\leq N-1$. Then:\n\\[\\pi(I_{N,n})\\subseteq C_{N-1,n}\\cap\\left(\\bigcap_{0\\leq i \\gamma_{31}\/\\left(\\gamma_{31}+\\gamma_{21}\\right)$ to ensure that $C_{\\textrm{2-half}}^{+'}$ is positive.). Now the optimal $\\rho_x$ is $1$ since $C_{\\textrm{1-half}}^{+'}$ is a monotonically increasing function over $\\rho_x$. The optimal $\\beta$ is given as follows by maximizing $C_{\\textrm{1-half}}^{+'}\\big\\vert_{\\rho_x = 1}$, which stands for the value of $C_{\\textrm{1-half}}^{+'}$ given $\\rho_x=1$ where the resulting function is concave over $\\beta$:\n\\begin{equation}\n\\beta^*=\\max \\left(\\beta_2, \\beta_2^* \\triangleq \\frac{1}{2}+\\frac{1}{2}\\frac{\\gamma_{31}-\\gamma_{32}(1-\\lambda)}{\\sqrt{\\left(\\gamma_{31}-\\gamma_{32}(1-\\lambda)\\right)^2+4\\gamma_{31}\\gamma_{32}(1-\\lambda)^2}}\\right) .\n\\end{equation}\n\nIf $\\beta_1 \\leq \\beta \\leq \\beta_2$, we let $C_{\\textrm{1-half}}^{+'}(\\rho_x) = C_{\\textrm{2-half}}^{+'}(\\rho_x)$ in order to solve the min-max problem since there must exist an unique crossing point. The optimal $\\rho_x$ can be obtained as:\n\\begin{equation}\n\\rho_x^* = \\frac{\\sqrt{\\gamma_{21}\\lambda\\beta\/(1-\\lambda)}-\\sqrt{\\gamma_{32}(1-\\beta)}}{\\sqrt{\\gamma_{31}\\beta}}.\\label{eq_half_rho_x}\n\\end{equation}\nBy substituting \\eqref{eq_half_rho_x} into \\eqref{eq_half_lower_deri}, we can maximize $C_{\\textrm{2-half}}^{+'}$ for $\\beta \\in [\\beta_1,\\beta_2]$, which leads to the following optimal $\\beta$:\n\\begin{equation}\n\\beta^*=\\label{eq_half_beta}\\min \\left(\\beta_2,\\beta_3^* \\triangleq \\frac{1}{2}+\\frac{1}{2}\\frac{\\gamma_{31}+\\gamma_{32}(1-\\lambda)}{\\sqrt{(\\gamma_{31}+\\gamma_{32}(1-\\lambda))^2+4\\gamma_{21}\\gamma_{32}\\lambda(1-\\lambda)}}\\right).\n\\end{equation}\nWhen $\\beta_3^*$ is optimal for a given $\\lambda$, the lower bound on the minimum energy per bit can be derived by substituting $\\beta_3^*$ and the corresponding $\\rho_x^*$ into \\eqref{eq_half_lowerbound} and \\eqref{eq_half_lower_deri}:\n\\begin{eqnarray}\\label{eq_half_lowerbound_3}\n\\nonumber\\left(\\frac{E_b}{N_0}\\right)_{\\min}^{-} &=& \\min_{0<\\lambda\\leq 1} \\frac{\\log_e 2}{C_{\\textrm{2-half}}^{+'}}\\\\\n&=& \\min_{0<\\lambda\\leq 1} \\frac{4\\log_e 2}{\\gamma_{31}-\\gamma_{32}(1-\\lambda) + \\sqrt{\\left(\\gamma_{31}+\\gamma_{32}(1-\\lambda)\\right)^2+4\\gamma_{21}\\gamma_{32}\\lambda(1-\\lambda)}},\n\\end{eqnarray}\nwhere the denominator is a concave function over $\\lambda \\in (0,1]$. Therefore, the optimal $\\lambda$ can be obtained by maximizing the denominator in \\eqref{eq_half_lowerbound_3}:\n\\begin{equation}\\label{eq_opt_lambda}\n\\lambda_{\\textrm{opt}} = \\frac{\\gamma_{31}+\\gamma_{32}-2\\gamma_{21}-\\sqrt{\\gamma_{31}^2+\\gamma_{31}\\gamma_{32}+\\gamma_{21}\\gamma_{32}}}{\\gamma_{32}-4\\gamma_{21}}\n\\end{equation}\nCorrespondingly, we can calculate the lower bound on the minimum energy per bit by substituting $\\lambda_{\\textrm{opt}}$ into \\eqref{eq_half_lowerbound_3}:\n\\begin{equation}\\label{eq_half_lowerbound_4}\n\\left(\\frac{E_b}{N_0}\\right)_{\\min}^{-} = \\frac{2(\\gamma_{32}-4\\gamma_{21})\\log_e 2 }{\\gamma_{31}\\gamma_{32}-2\\gamma_{21}\\gamma_{31}+\\gamma_{21}\\gamma_{32}-2\\gamma_{21}\\sqrt{\\gamma_{31}^2+\\gamma_{31}\\gamma_{32}+\\gamma_{21}\\gamma_{32}}}.\n\\end{equation}\n\nNow we summarize the solution by combining all the cases that we discussed above.\n\\begin{enumerate}\n\\item For $\\lambda \\in (0,\\gamma_{31}\/(\\gamma_{31}+\\gamma_{21})]$, $C_{\\textrm{2-half}}^{+'}$ is not ensured to be larger than $C_{\\textrm{1-half}}^{+'}$ for all $\\rho_x \\in [0,1]$. Accordingly, the lower bound is $\\left(\\frac{E_b}{N_0}\\right)_{\\min}^{-}$ in \\eqref{eq_half_lowerbound_4} if $\\lambda_{\\textrm{opt}}\\in (0, \\gamma_{31}\/(\\gamma_{31}+\\gamma_{21})]$. Otherwise, the lower bound is $\\frac{2\\log_e 2}{C_{\\textrm{1-half}}^{+'}}\\big |_{\\lambda=\\gamma_{31}\/(\\gamma_{31}+\\gamma_{21}),\\beta_{3}^*,\\rho_x^*}$.\n\\item For $\\lambda \\in (\\gamma_{31}\/(\\gamma_{31}+\\gamma_{21}), 1]$, one-dimensional searching can be utilized to obtain the lower bound. Specifically, for a given $\\lambda \\in (\\gamma_{31}\/(\\gamma_{31}+\\gamma_{21}), 1]$, we obtain an intermediate lower bound by taking the minimum among the ones derived in the following four cases. Then we search over all possible $\\lambda \\in (\\gamma_{31}\/(\\gamma_{31}+\\gamma_{21}), 1]$ to have the lower bound for case 2).\n \\begin{itemize}\n \\item If $\\beta_2\\leq\\beta_2^*,\\ \\beta_3^*$, the corresponding lower bound is $\\frac{2\\log_e 2}{C_{\\textrm{1-half}}^{+'}}\\big |_{\\beta_2^*,\\rho_x=1}$;\n \\item If $\\beta_3^*<\\beta_2\\leq\\beta_2^*$, the corresponding lower bound is $\\frac{2\\log_e 2}{\\max\\left(C_{\\textrm{1-half}}^{+'}|_{\\beta_2^*,\\rho_x=1}, C_{\\textrm{1-half}}^{+'}|_{\\beta_3^*,\\rho_x^*}\\right)}$;\n \\item If $\\beta_2^*<\\beta_2\\leq\\beta_3^*$, the corresponding lower bound is $\\frac{2\\log_e 2}{C_{\\textrm{1-half}}^{+'}}\\big |_{\\beta_2,\\rho_x=1}$;\n \\item Otherwise, $\\beta_2^*,\\ \\beta_3^*<\\beta_2$, the corresponding lower bound is $\\frac{2\\log_e 2}{C_{\\textrm{1-half}}^{+'}}\\big |_{\\beta_3^*,\\rho_x^*}$.\n \\end{itemize}\n\\item The global lower bound can be calculated as the minimum value between that for case 1) of $\\lambda \\in (0,\\gamma_{31}\/(\\gamma_{31}+\\gamma_{21})]$ and that for case 2) of $\\lambda \\in (\\gamma_{31}\/(\\gamma_{31}+\\gamma_{21}), 1]$.\n\\end{enumerate}\n\n\n\n\n\n","meta":{"redpajama_set_name":"RedPajamaArXiv"}}